You are on page 1of 818

9782_9789814719926_tp.

indd 1 22/8/16 3:42 PM


May 2, 2013 14:6 BC: 8831 - Probability and Statistical Theory PST˙ws

This page intentionally left blank


K E

9782_9789814719926_tp.indd 2 22/8/16 3:42 PM


Published by
World Scientific Publishing Co. Pte. Ltd.
5 Toh Tuck Link, Singapore 596224
USA office: 27 Warren Street, Suite 401-402, Hackensack, NJ 07601
UK office: 57 Shelton Street, Covent Garden, London WC2H 9HE

Library of Congress Cataloging-in-Publication Data


Thomas, G. H. (Gerald Harper), 1942–
Geometry, language, and strategy / Gerald H. Thomas.
v. 1. [Without special title] -- v. 2. The dynamics of decision processes
p. cm. -- (Series on knots and everything ; v. 37, v. 59)
Includes bibliographical references and index.
ISBN-10 981-256-617-1 (v. 1 : hdbk. : alk. paper) -- ISBN-13 978-981-256-617-1 (v. 1)
ISBN 978-981-4719-92-6 (v. 2 : hardcover)
1. Game theory. 2. Statistical decision. 3. Management science. I. Title.
II. Series.

QA269 .T578 2006


519.3--dc22
2005057872

British Library Cataloguing-in-Publication Data


A catalogue record for this book is available from the British Library.

Wolfram Mathematica® is registered trademark of Wolfram Research, Inc.

Copyright © 2017 by World Scientific Publishing Co. Pte. Ltd.


All rights reserved. This book, or parts thereof, may not be reproduced in any form or by any means,
electronic or mechanical, including photocopying, recording or any information storage and retrieval system
now known or to be invented, without written permission from the publisher.

For photocopying of material in this volume, please pay a copying fee through the Copyright Clearance
Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to photocopy is not
required from the publisher.

Printed in Singapore

EH - Geometry, Language-V59-V2.indd 5 22-08-16 3:27:47 PM


Dedication

“…and have always in view not only the present but also the coming
generations1…”
Leo and Aaron
Maggie and Kate
Aleksandra, Michael and Teodora

1
Seven generation sustainability (Wikipedia, 2015).
May 2, 2013 14:6 BC: 8831 - Probability and Statistical Theory PST˙ws

This page intentionally left blank


Foreword

In the Foreword to Gerald Thomas’ previous book “Geometry, Language


and Strategy” I wrote: “This is a most unusual book, breaking new
ground in the theory of games. Thomas begins with the ideas of classical
game theory, as formulated by von Neumann and Morgenstern. He takes
this approach through a turn that gives rise to a model that has many
players, each with their own game parameters and a global structure that
partakes of the mathematics of differential geometry. As Thomas says in
his Introduction ‘The best borrowings are probably of a mathematical
nature, since Mathematics is superb at speaking deeply without knowing
what it is talking about. For purposes of considering dynamics, I believe
the ideal mathematical language is geometry.’ It is the nature of
mathematics to have multiple interpretations and applications but it
requires artistry and a conceptual shift to find the patterns that match in a
new field.
Once there is a multiplicity (imagine indeed a continuum) of players
(strategies) in the game, then it is natural to posit a metric that measures
the relative distance between two strategies. This book creates, using
geometry and physical concepts, a language appropriate to the discussion
of the dynamic behavior of strategies.
The metric in its global properties and its particulars is variable. It is
affected by the very strategies that it describes. The space (metric) tells
the strategies how to move, while the strategies tell the space how to
curve (distort). The situation is exactly analogous to that of general
relativity where space tells matter how to move and matter tells space
how to curve. Thomas takes this analogy seriously and develops a theory
of games that is based in the mathematical and conceptual structure of

vii
viii Geometry, Language and Strategy—Vol. 2

general relativity and the allied subjects of electromagnetism and gauge


theory. The result is a fascinating read that will enlighten game theorists,
relativists and all readers with an active curiosity about the structure of
the world and the enterprise and exploration that is the making of
mathematical models.”
In this second volume and sequel to the first book, we find a much
expanded and deepened look at the new subject of game theory, decision
theory and geometry. There is a wealth of new material in this second
volume and many places for the reader to begin specific study. The new
book is self-contained and does not require its predecessor, but a reader
may wish to have the first book as a companion and guide to this wider
treatment. The first part of the new book gives a detailed and clear
introduction to the differential geometric formulation of games and
decisions that is basic to the study. It is an in-depth development of
decision process theory. In the second part, the book tests the properties
of the theory by analyzing numerous scenarios of the following major
types: steady state scenarios, reflection and transmission scenarios,
resonant scenarios and transient scenarios. Here we see, through this
choice of terminology, the essential character of the work. Thomas
understands that there is wisdom about process and the making of
decisions in the intellectual tools of the physicist and the engineer. By
taking these concepts and generalizing them to the social and economic
domain he makes a great step in creating a transdisciplinary context and
effective arena for physical and social science. In social science the
models are not separate from the ongoing processes that are being
modeled. It will be of great interest to see how these models of Gerald H.
Thomas become part of the landscape of social and economic action.

Louis H. Kauffman
December 16, 2015
Preface

This project is an inquiry on the extent to which decision processes can


be considered deterministic. For many people, the answer is that
decisions are not, though a stochastic approach can be fruitful. I have
taken the approach that some aspects of decision making can be
considered deterministic; I rely not on the usual stochastic approach but
on a deterministic and physical model of decisions, a model that
incorporates both the competitive aspect of decisions as well as the
temporal aspects of a causal world. Not everyone agrees with this
hypothesis, yet ultimately agreement or disagreement rests not on our
beliefs but on whether the hypothesis provides an explanation of known
data. Further, a useful theory is ultimately one that gives insight into
behaviors that have not yet happened; insight that is both qualitative and
quantitative.
It may be helpful to indicate what the reader may find in this book.
The exposition of this theory requires making choices about what
material should be included and what examples should be chosen. I wish
I could argue that the theory of decision making has provided that
guidance. In point of fact, so far I haven’t been able to make such a
bootstrap. My choices have been motivated by my own experiences as
well as what I have learned from a vast number of people that I have met
over time.
Based on my experiences as a research physicist I have chosen what I
think are the important mathematical and physical topics. Based on my
experiences as an engineering professor I have chosen problems that I
believe provide pedagogical support. And based on my experiences from
industry I have chosen examples that I think are important and
instructive. But by far I owe the greatest debt to those that have shared

ix
x Geometry, Language and Strategy—Vol. 2

their experiences with me. This has occurred over many years and I am
not able to provide a detailed accounting. But I am able to point to a few
people who have been especially helpful for this phase of the project.
I am grateful to Stephen Wolfram and members of his staff at
Wolfram Research, Inc. (WRI) for ongoing support and especially from
Maryka Baraka and others from the Partnership Program Support at
Wolfram Research. Lou Kauffman strongly supported this project and
provided in this new phase, access and feedback from his students.
Gordon Kane provided feedback on the physics and pedagogy based on a
draft of the manuscript. Jack Behrend read and provided edits of
countless early drafts for which I am grateful. I received useful and
supportive feedback on my philosophical stance from Steve Rosen. Lisa
Maroski and Helen Kessler have held me accountable for my project and
goals as well as providing feedback on specific sections. They have
honed my understanding of Systems Dynamics as a deterministic
approach applied to a variety of real world applications. Keelan Kane
collaborated with me on some classic and paradoxical problems in the
literature on the prisoner’s dilemma, which led to chapter 7. Col. Edward
Sobiesk introduced me to the new field of Network Science, and the
interest at West Point in applications. I have benefited from the support
of the Milwaukee School of Engineering (MSOE) during the course of
this project. My Colleagues at MSOE have patiently listened to my
attempts to clarify my ideas. Most recently, I am particularly grateful to
Bharathwaj Muthuswamy and Jovan Jevtic and our weekly meetings to
discuss a variety of research topics including this one. Finally I am
grateful to my daughter Liisa M. Thomas for her legal advice and to my
wife Kathleen for her patience throughout the course of this project and
her commitment to seven generation sustainability.

G. H. Thomas
Contents

Foreword ................................................................................................................... vii


Preface ....................................................................................................................... ix
List of Tables ........................................................................................................... xxi
List of Figures ........................................................................................................ xxiii
Part 1 Physical and Game Theory Foundations..................................................... 1
Chapter 1 Introduction............................................................................................. 3
1.1 Pedagogy ...................................................................................................... 5
1.2 Decision Process Theory Goals .................................................................... 9
1.3 Introduction ................................................................................................ 12
1.4 Foundational Game Theory ........................................................................ 14
1.5 Newtonian Mechanics ................................................................................ 17
1.6 Newtonian Mechanics, Lagrange and Game Theory .................................. 20
1.7 Conceptual Basis and Payoffs .................................................................... 23
1.8 Electromagnetic Fields as Payoffs .............................................................. 27
1.9 Outcomes .................................................................................................... 34
1.10 Exercises..................................................................................................... 35
Chapter 2 Dynamics of Decision Processes ........................................................... 37
2.1 Time, Effort and Utility .............................................................................. 39
2.2 Geometric Forms ........................................................................................ 40
2.3 Source-Free Field Equations ...................................................................... 42
2.4 Sourced Field Equations ............................................................................. 45
2.5 Dynamic Payoff Equations ......................................................................... 47
2.6 Continuous Change versus Uncertainty...................................................... 50
2.7 Holonomic Coordinate Bases ..................................................................... 53
2.8 General Coordinate Bases .......................................................................... 55
2.9 Parallel Translations ................................................................................... 57
2.10 Geometric Structures .................................................................................. 61

xi
xii Geometry, Language and Strategy—Vol. 2

2.11 Decision Process Theory ............................................................................ 64


2.12 Outcomes .................................................................................................... 71
2.13 Exercises..................................................................................................... 73
Chapter 3 Inertial Behaviors ................................................................................. 75
3.1 Inertial Media ............................................................................................. 78
3.2 Inertial Cohesion ........................................................................................ 81
3.3 Unifying Inertial Media with Payoff Fields ................................................ 85
3.4 Distance ...................................................................................................... 88
3.5 Principle of Least Action ............................................................................ 93
3.6 Orientation Flux Fields ............................................................................... 95
3.7 Organizational Dynamics—Energy Momentum Stress Tensor Tab ......... 100
3.8 Hamiltonian Formulation ......................................................................... 103
3.9 Quantum of Action ................................................................................... 109
3.10 Outcomes .................................................................................................. 113
3.11 Exercises................................................................................................... 116
Chapter 4 Persistent Behaviors ........................................................................... 119
4.1 Coordinate and Non-Coordinate Bases and Potentials ............................. 121
4.2 Normal-Form Coordinate Basis—Hidden Symmetries ............................ 125
4.3 Normal-Form Coordinate Basis—Harmonic Gauge ................................ 128
4.4 Normal-Form Coordinate Basis—Field Equations................................... 130
4.5 Isometry and Hidden Symmetry ............................................................... 136
4.6 Co-Moving Orthonormal Coordinate Basis.............................................. 138
4.7 Decomposition of the Orientation Potentials   ................................. 144

4.8 Player Fixed Frame Model ....................................................................... 146


4.9 Persistency under the Player Fixed Frame Model .................................... 149
4.10 Outcomes .................................................................................................. 155
4.11 Exercises................................................................................................... 157
Chapter 5 Deconstructing the Theory ................................................................. 163
5.1 Player Fixed Frame Model ....................................................................... 165
5.1.1 Dynamic Components ..................................................................... 167
5.1.2 Persistent Components .................................................................... 169
5.2 Field Equations—Co-Moving Frame ....................................................... 171
5.2.1 Time Evolution of Vorticity and Expansion ................................... 173
5.2.2 Codacci Equations .......................................................................... 174
5.2.3 Remaining Equations ...................................................................... 175
Contents xiii

5.2.4 Summary and Final Codacci Equation ............................................ 176


5.2.5 Conservation Laws.......................................................................... 180
5.3 Field Equations—Player Fixed Frame Model .......................................... 181
5.3.1 Expansion Equations....................................................................... 181
5.3.2 Six Symmetry Classes..................................................................... 182
5.3.3 Time Evolution of the “Mixed” Expansion Coefficient .................. 184
5.3.4 Codacci Equations .......................................................................... 185
5.3.5 Vorticity Equations ......................................................................... 186
5.3.6 Remaining Potentials and their Evolution ....................................... 187
5.3.7 Conservation Laws.......................................................................... 189
5.4 Player Fixed Frame Model—Distinctions ................................................ 189
5.4.1 Distinctions—Persistency ............................................................... 191
5.4.2 Distinctions—Electromagnetic Fields............................................. 194
5.4.3 Distinctions—Tidal Fields .............................................................. 196
5.4.4 Distinctions—Inertial Fields ........................................................... 198
5.5 Outcomes .................................................................................................. 202
5.6 Exercises................................................................................................... 204
Chapter 6 Steady State Harmonics ..................................................................... 209
6.1 Steady State Scalars .................................................................................. 211
6.2 Streamline Solutions................................................................................. 214
6.3 Models for the Player Current .................................................................. 217
6.4 Inactive Stress Equations .......................................................................... 220
6.5 Active Stress Equations ............................................................................ 223
6.6 Frame-Wave Equation .............................................................................. 225
6.7 Centrally Co-Moving Hypothesis ............................................................. 231
6.8 Single Active Strategy Dynamic Solutions .............................................. 234
6.9 Outcomes .................................................................................................. 237
6.10 Exercises................................................................................................... 239
Part 2 Decision Engineering ................................................................................. 259
Chapter 7 Prisoner’s Dilemma: A Code of Conduct Application ..................... 261
7.1 Introduction .............................................................................................. 263
7.2 Egoists and Altruists ................................................................................. 265
7.3 Prisoner’s Dilemma—the Story in Normal Form ..................................... 268
7.4 Equivalent Formulations of the Initial Conditions.................................... 273
7.5 Persistency, Variability and Code of Conduct .......................................... 277
xiv Geometry, Language and Strategy—Vol. 2

7.6 Known Behaviors ..................................................................................... 283


7.7 Known Strategic Flows ............................................................................ 285
7.8 Known Payoff Fields and Orientation Potentials...................................... 286
7.9 Symmetric Co-Moving Frame Values ...................................................... 289
7.10 Numerical Values and Inertial Effects ...................................................... 292
7.11 Outcomes .................................................................................................. 294
7.12 Exercises................................................................................................... 296
Chapter 8 Prisoner’s Dilemma—Steady State Geometry ................................. 299
8.1 Inertial Stress Behaviors ........................................................................... 302
8.2 Model Results ........................................................................................... 303
8.3 Sensitivity Analysis .................................................................................. 305
8.4 Strain Behaviors ....................................................................................... 308
8.5 Player Bonding Results ............................................................................ 310
8.6 Shear Results ............................................................................................ 311
8.7 Player Interest Results .............................................................................. 314
8.8 Persistent Behaviors ................................................................................. 315
8.8.1 Electric Field Components f j  .................................................... 317
8.8.2 Player Engagement Components e ............................................. 317
8.8.3 Inactive Metric Components  jk .................................................... 318

8.8.4 Inactive Flow Components E j ..................................................... 319

8.8.5 Inactive Transformation Components E j .................................... 320
8.8.6 Inactive Determinant  jk ............................................................. 322
8.9 Outcomes .................................................................................................. 323
8.10 Exercises................................................................................................... 326
Chapter 9 Prisoner’s Dilemma: Determinism and Chaos ................................. 329
9.1 Determinism and Chaos ........................................................................... 332
9.2 Time Bonding ........................................................................................... 336
a
9.3 Strategy Flows E  ................................................................................ 340
ab
9.4 Time, Distance and g ........................................................................... 343
9.5 Sensitivity Analysis for Dynamic Behaviors ............................................ 348
9.6 Baseline .................................................................................................... 349
9.7 Low Stakes, High Stakes .......................................................................... 351
9.8 Normal Form Behaviors ........................................................................... 354
9.9 Recurrence Plots and Chaos ..................................................................... 357
9.10 Outcomes .................................................................................................. 360
9.11 Exercises................................................................................................... 362
Contents xv

Chapter 10 Robinson Crusoe Economics ............................................................ 369


10.1 Decision Process Taxonomy .................................................................... 370
10.2 Identify the Story—Robinson Crusoe ...................................................... 373
10.3 Put Story into Normal Form ..................................................................... 376
10.4 Identify the Natural Units ......................................................................... 377
10.5 Co-Moving Coordinate Basis ................................................................... 378
10.6 Known Behaviors ..................................................................................... 380
10.7 Steady State Behaviors ............................................................................. 382
10.8 Frame-Waves............................................................................................ 385
10.9 Limits to Choice ....................................................................................... 389
10.10 Travelling Frame-Waves .......................................................................... 393
10.11 Standing Frame-Waves ............................................................................ 395
10.12 Limits to Growth ...................................................................................... 398
10.13 Decision Process Values........................................................................... 400
10.14 Outcomes .................................................................................................. 401
10.15 Exercises................................................................................................... 403
Chapter 11 Short Historical Review of Game Theory ....................................... 407
11.1 Historical Review ..................................................................................... 408
11.2 Code of Conduct....................................................................................... 413
11.2.1 Adam Smith—The Invisible Hand................................................ 413
11.2.2 Prisoner’s Dilemma and the Tragedy of the Commons ................ 416
11.3 Tragedy of the Commons ......................................................................... 419
11.3.1 Blame Game ................................................................................. 419
11.3.2 Freeway Game .............................................................................. 420
11.3.3 Political Commentary Game ......................................................... 420
11.3.4 Standard of Behavior .................................................................... 421
11.4 Self-Interest versus Public-Interest ........................................................... 421
11.5 Strategic Control....................................................................................... 423
11.5.1 Standing in Line Game ................................................................. 423
11.5.2 Zero Sum Strategy Games ............................................................ 425
11.5.3 Accountability and Number of Agents.......................................... 426
11.6 Outcomes .................................................................................................. 427
11.7 Exercises................................................................................................... 427
Chapter 12 Decisions and Determinism .............................................................. 429
12.1 Dealing with Uncertainty ......................................................................... 431
xvi Geometry, Language and Strategy—Vol. 2

12.1.1 Strictly Determined Processes ............................................................. 431


12.1.2 Repetitive Games and Mixed Strategies .............................................. 432
12.1.3 Bayesian View..................................................................................... 433
12.1.4 Calculus View ..................................................................................... 435
12.2 Utility Theory ........................................................................................... 437
12.2.1 Preferences .......................................................................................... 438
12.2.2 Measurement ....................................................................................... 439
12.2.3 Cardinal and Ordinal Utility ................................................................ 441
12.2.4 Local Utility ........................................................................................ 442
12.2.5 Payoffs ................................................................................................ 444
12.3 Dynamic Law of Competition .................................................................. 445
12.3.1 Game Theory Limit ............................................................................. 446
12.3.2 Linear Programming............................................................................ 448
12.3.3 Decision Process Theory Solutions ..................................................... 451
12.3.4 Law of Competition ............................................................................ 452
12.4 Dynamic Law of Cooperation .................................................................. 454
12.4.1 Game Theory Limit ............................................................................. 454
12.4.2 Decision Process Theory Solutions ..................................................... 456
12.4.3 Law of Cooperation ............................................................................. 458
12.5 Dynamic Law of Organization ................................................................. 460
12.5.1 Organization and Contextual Frame .................................................... 461
12.5.2 Organizational Equilibrium ................................................................. 462
12.6 Outcomes .................................................................................................. 463
12.7 Exercises................................................................................................... 464
Chapter 13 Global Systems View......................................................................... 467
13.1 Global Connections .................................................................................. 469
13.2 Systems Dynamics Modeling ................................................................... 470
13.3 Physically Based Models .......................................................................... 474
13.4 The Three Laws—Reprise ........................................................................ 475
13.5 Energy-Momentum Convertibility ........................................................... 476
13.6 Opportunity .............................................................................................. 479
13.7 Inertia and Interest Opportunity ............................................................... 481
13.8 Opportunity Cost ...................................................................................... 483
13.9 Closing the Loop ...................................................................................... 485
13.10 Outcomes .................................................................................................. 488
13.11 Exercises................................................................................................... 489
Contents xvii

Part 3 Three Active Strategy Systems ................................................................. 493


Chapter 14 Vorticity and Active Strategies ........................................................ 495
14.1 Introduction .............................................................................................. 497
14.2 Decision Process Theory Strategies.......................................................... 499
14.3 Strategic Effort ......................................................................................... 501
14.4 Known Behaviors ..................................................................................... 503
14.5 Constructing an Ortho-Normal Frame ...................................................... 505
14.6 Surface Geometry ..................................................................................... 506
14.7 Geometry of Competitive Field ................................................................ 509
14.8 Stress and Strain Fields ............................................................................ 509
14.9 3+1 Active Strategy Equations ................................................................. 511
14.9.1 Pressure ......................................................................................... 512
14.9.2 Acceleration .................................................................................. 513
14.9.3 Player Bonding ............................................................................. 513
14.9.4 Player Shear Tensor ...................................................................... 513
14.9.5 Player Interest ............................................................................... 514
14.9.6 Player Payoff................................................................................. 515
14.9.7 Player Passion ............................................................................... 515
14.9.8 Composite Payoff.......................................................................... 516
14.9.9 Stress Components ........................................................................ 516
14.9.10 Inactive Frame Transformations ................................................. 517
14.9.11 Seasonal Vector Potential ........................................................... 517
14.9.12 Proper Time Equations................................................................ 519
14.9.13 Central Time Equations .............................................................. 520
14.9.14 Waves ......................................................................................... 521
14.10 2+1 Preview of Lattice Models ................................................................ 521
14.11 Locked Behavior ...................................................................................... 524
14.12 Player Bonding Effects ............................................................................. 527
14.13 3+1 Active Strategy Lattice Models ......................................................... 528
14.14 Institutional Vortex................................................................................... 531
14.15 Outcomes .................................................................................................. 532
14.16 Exercises................................................................................................... 536
Chapter 15 Game Equivalency ............................................................................ 545
15.1 Equivalency .............................................................................................. 547
15.2 Still Point and Free Fall ............................................................................ 549
xviii Geometry, Language and Strategy—Vol. 2

15.3 Engagement .............................................................................................. 552


15.4 Entitlement ............................................................................................... 554
15.5 Focus Direction ........................................................................................ 556
15.6 Symmetric Form Transformations............................................................ 558
15.7 Game Payoffs and Potential Field Gradients ............................................ 561
15.8 Game Value and Entitlement Payoffs....................................................... 562
15.9 Player Engagement Values ....................................................................... 565
15.10 Cooperative Payoffs ................................................................................. 567
15.11 Seasonal Player and Natural Resonances ................................................. 569
15.12 Scale Invariance ....................................................................................... 572
15.13 Three Active Strategy Models .................................................................. 575
15.14 Outcomes .................................................................................................. 577
15.15 Exercises................................................................................................... 580
Chapter 16 Two-Person Decision Processes ....................................................... 591
16.1 War Games versus Social Decision Processes ......................................... 592
16.2 Attack-Defense Model.............................................................................. 593
16.3 Locked Behaviors in f0AD Model ............................................................ 595
16.4 Decision Flow in f0AD Model .................................................................. 600
16.5 Gauge Considerations in f0AD Model ...................................................... 603
16.6 Work-Wealth-Wisdom Model .................................................................. 606
16.7 Streamlines in f0WWW Model ................................................................. 609
16.8 Forced f1WWW Model ............................................................................. 614
16.9 Travelling Waves ..................................................................................... 618
16.10 Standing Waves ........................................................................................ 622
16.11 Recurrence Plots ....................................................................................... 625
16.12 Outcomes .................................................................................................. 626
16.13 Exercises................................................................................................... 628
Chapter 17 Steady State Geometries in 3D......................................................... 649
17.1 Validity Regions ....................................................................................... 652
17.2 Transfer Functions .................................................................................... 658
17.3 Player Payoffs........................................................................................... 661
17.4 Player Passion........................................................................................... 668
17.5 Player Engagement ................................................................................... 671
17.6 Player Interest ........................................................................................... 674
17.7 Player Strategy Bias ................................................................................. 677
Contents xix

17.8 Decision Pressure, Player Bonding and Shear .......................................... 679


17.9 Composite Payoff ..................................................................................... 687
17.10 Seasonal Payoff ........................................................................................ 688
17.11 Outcomes .................................................................................................. 689
17.12 Exercises................................................................................................... 693
Chapter 18 Structural Distinctions for Decision Processes ............................... 705
18.1 Morphogenesis and Structural Stability.................................................... 707
18.2 Causality and Determinism ...................................................................... 709
18.3 Harmonic Steady State Fields................................................................... 715
18.4 High Frequency Harmonics ...................................................................... 718
18.5 Network Connections ............................................................................... 722
18.6 Individuality, Honoring Contracts and Sustainability............................... 725
18.7 Energy Storage—Capacitance, Inductance and Gravitas ......................... 727
18.8 Limits to Choice—High Bias Example .................................................... 730
18.9 Game Equivalency and Free Fall.............................................................. 734
18.10 Acceleration.............................................................................................. 739
18.11 Outcomes .................................................................................................. 743
18.12 Exercises................................................................................................... 746
Chapter 19 Social Distinctions for Decision Processes....................................... 759
19.1 Code of Conduct....................................................................................... 760
19.2 Player Entitlement .................................................................................... 765
19.3 Player Engagement ................................................................................... 770
19.4 Player Interest ........................................................................................... 775
19.5 Player Passion........................................................................................... 777
19.6 Mutual Player Support.............................................................................. 778
19.7 Hidden-in-Plain-Sight............................................................................... 781
19.8 Outcomes .................................................................................................. 782
19.9 Exercises................................................................................................... 784
Chapter 20 Afterword .......................................................................................... 795
Bibliography ........................................................................................................... 799
Index ....................................................................................................................... 807
May 2, 2013 14:6 BC: 8831 - Probability and Statistical Theory PST˙ws

This page intentionally left blank


List of Tables

Table 5.1. Player Fixed Frame Model—Persistent and Dynamic Components .... 170
Table 5.2. Time Equations .................................................................................... 175
Table 5.3. Gradient Equations on Surface Normal to Flow .................................. 178
Table 5.4. Player Fixed Frame Model—Electromagnetic Fields .......................... 195
Table 5.5. Player Fixed Frame Model—Tidal Fields ........................................... 197
Table 5.6. Player Fixed Frame Model—Inertial Components .............................. 201
Table 9.1. Player Stakes  and Decision  Choices for Sensitivity Analysis ...... 348
Table 16.1. Mesh Analysis in WWW Model .......................................................... 639

xxi
May 2, 2013 14:6 BC: 8831 - Probability and Statistical Theory PST˙ws

This page intentionally left blank


List of Figures

Figure 7.1. Pressure contours .............................................................................. 293


Figure 7.2. Pressure p   . ................................................................................... 293
Figure 8.1. Pressure p versus  . ....................................................................... 303
Figure 8.2. Acceleration q versus  ................................................................. 303
Figure 8.3. Inertial potential  q ......................................................................... 305
Figure 8.4. Pressure for Nash solution: zero game value..................................... 306
Figure 8.5. Pressure for common good solution: zero game value ...................... 306
Figure 8.6. Common good solution for    : zero game value...................... 307
1

Figure 8.7. Nash solution for    : non-zero game value ............................. 307
1

Figure 8.8. Player bonding gradient        (PDM) ................................. 311


Figure 8.9. Player bonding volume as a function of strategy  (PDM)......... 311
q
Figure 8.10. Form factor e with initial value normalized to unity (PDM) ......... 312
Figure 8.11. Reduced bond shear   q   (PDM) .............................................. 312
1 2

Figure 8.12. Reduced bond shear      (PDM) ............................................... 313


1 2

Figure 8.13. Player interest field    in the co-moving frame (PDM). ........ 316
Figure 8.14. Electric field f j    in the co-moving frame (PDM). ................... 316
Figure 8.15. Player engagement e   in the co-moving frame (PDM) ........ 316
Figure 8.16. Diagonal components ln  jj   in the normal form frame(PDM) .... 316
Figure 8.17.  ln  jj   in the normal form frame (PDM) .................................. 319
Figure 8.18. Flow components of the Killing vector E j   (PDM) ..................... 319
Figure 8.19. Prisoner 1 inactive Killing vector components E    (PDM) ........ 319
1

Figure 8.20. Prisoner 2 inactive Killing vector components E    (PDM) ........ 319
2

Figure 8.21. Player bonding  (PDM) ............................................................... 321


Figure 8.22. Determinant of the inactive metric  jk   (PDM) ......................... 323
Figure 9.1. Time Bonding g (PDM) ................................................................ 339
Figure 9.2. Time flow E t  ,  (PDM) .............................................................. 340

xxiii
xxiv Geometry, Language and Strategy—Vol. 2

Figure 9.3. Time flow E t  ,  (PDM) .............................................................. 341


Figure 9.4. Active strategy flow E u (PDM) ....................................................... 342
Figure 9.5. Active strategy contours E u (PDM) ................................................. 343
Figure 9.6. Time bonding g tt  ,  (PDM) ......................................................... 344
Figure 9.7. Distance-time central contours (PDM) .............................................. 344
Figure 9.8. Contours for central frame coordinates (PDM) ................................. 344
Figure 9.9. Strategy bonding g uu  ,  (PDM).................................................... 345
Figure 9.10. Null-geodesics constraints for high stakes low  ........................... 347
Figure 9.11. Baseline contours E t   ,  ............................................................. 349
Figure 9.12. Baseline contours g tt  ,  .............................................................. 350
Figure 9.13. Low stakes and low  contour plot for the metric g tt  ,  ....... 350
Figure 9.14. Low stakes and low  contour plot for the flow E t   ,  ........ 351
Figure 9.15. High stakes and low  time bonding g tt  ,  ................................ 352
Figure 9.16. High stakes and high  time bonding g tt  ,  .............................. 353
Figure 9.17. Baseline normal form basis flows ..................................................... 354
Figure 9.18. Baseline confess choice for prisoner 1 .............................................. 355
Figure 9.19. Baseline player interests in normal form basis .................................. 355
Figure 9.20. Baseline player interest for s1 code of conduct ................................ 356
Figure 9.21. High stakes-low  player interest for s1 code of conduct ........... 356
Figure 9.22. High stakes-low  distance time contours........................................ 357
Figure 9.23. Strategy flow recurrence plot two-frequencies PDM ........................ 359
Figure 9.24. Time flow recurrence plot two frequencies PDM ............................. 359
Figure 9.25. Proper-strategy vector E u E u  high stakes, low  ................... 364
Figure 9.26. High stakes, low  , null geodesic c .............................................. 364
Figure 9.27. Coordinate timelines and spacelines for high stakes and low  ....... 365
Figure 9.28. Pressure for high stakes and low  ................................................... 365
Figure 9.29. Proper-time   u , t  for high stakes and low  ................................ 365
Figure 9.30. Proper-distance   u , t  for high stakes and low  .......................... 365
Figure 9.31. Flow along E u for high stakes and low  ..................................... 366
Figure 9.32. Flow along E u for high stakes and high  .................................... 366
Figure 9.33. Proper-time   u , t  for high stakes and high  ............................... 366
Figure 9.34. Proper-distance   u , t  for high stakes and high  ........................ 367
Figure 9.35. Strategy flow recurrences for single harmonic PDM ........................ 367
Figure 9.36. Time flow recurrences for single harmonic PDM ............................. 367
Figure 9.37. Strategy flow recurrences for two-frequency PDM ........................... 368
Figure 10.1. Pressure p   ................................................................................... 382
List of Figures xxv

Figure 10.2. Acceleration q   ........................................................................... 382


Figure 10.3. Player bonding    ..................................................................... 383
Figure 10.4. Player vorticity    ...................................................................... 383
Figure 10.5. Player engagement e   ............................................................... 383
Figure 10.6. Frame transformation E j    .......................................................... 384
Figure 10.7. Player charge E j ............................................................................. 384
Figure 10.8. Player interest f j ......................................................................... 384
Figure 10.9. Left-travelling time wave and its approximation .............................. 387
Figure 10.10. Left-travelling strategy wave and its approximation ........................ 388
Figure 10.11. Boundary difference    .............................................................. 390
Figure 10.12. Strategy flows E a  , 0  ................................................................. 393
Figure 10.13. Time flow E t   ,  ........................................................................ 393
Figure 10.14. Strategy flow E u  ,  .................................................................... 394
Figure 10.15. Frame transformation E u  ,  ....................................................... 394
Figure 10.16. Time bonding g tt  ,  ..................................................................... 394
Figure 10.17. Time flow contour plot ..................................................................... 395
Figure 10.18. Strategy flow E u  ,  contours ..................................................... 396
Figure 10.19. Frame transformation E u  ,  ....................................................... 397
Figure 10.20. Time bonding g tt  ,  contours ...................................................... 398
Figure 10.21. Time bonding gtt for bond   20 ................................................. 399
Figure 10.22. Time bonding contours for gtt and   20 .................................... 399
Figure 10.23. Known behavior of F j tu  , 0  ......................................................... 400
Figure 10.24. Decision value F j tu  ,  ................................................................ 401
Figure 10.25. Distance time contours for   2 ................................................... 405
Figure 10.26. Distance time contours for   20 ................................................. 406
Figure 10.27. Time bonding field with low frequency   3 ................................ 406
Figure 12.1. Prisoner's dilemma  11   ............................................................... 457
Figure 12.2. Prisoner’s dilemma  12   .............................................................. 457
Figure 13.1. Systems dynamic predator-prey model ............................................ 472
Figure 13.2. Predator prey populations ................................................................. 473
Figure 13.3. Limits to growth ............................................................................... 490
Figure 13.4. Software Project ............................................................................... 491
Figure 14.1. Acceleration flow in Laplace's equation ........................................... 523
Figure 14.2. Acceleration flow in non-linear Laplace's equation ......................... 523
Figure 14.3. Pressure in 2  1 dimensions ........................................................... 525
Figure 14.4. Acceleration qz in 2  1 dimensions ............................................. 525
xxvi Geometry, Language and Strategy—Vol. 2

Figure 14.5. Pressure with x structure................................................................. 526


Figure 14.6. Acceleration qz with x structure .................................................. 526
Figure 14.7. Acceleration flow q ...................................................................... 526
Figure 14.8. Player-1 interest flow I1 .................................................................. 526
Figure 14.9. Composite payoff  ........................................................................ 527
Figure 14.10. Player 2 payoff B2 ........................................................................... 527
Figure 14.11. Payoff vector field B1 ...................................................................... 529
Figure 14.12. Payoff vector field B2 ...................................................................... 529
Figure 14.13. Composite payoff field ω ............................................................... 529
Figure 14.14. Characteristic potential a ................................................................. 530
Figure 14.15. Passion vector field P1 ..................................................................... 530
Figure 14.16. Passion vector field P2 ..................................................................... 530
Figure 14.17. Payoff vector field B3 with code of conduct ................................... 531
Figure 14.18. Composite payoff ω with code of conduct ..................................... 531
Figure 14.19. Passion vector field P3 with code of conduct................................... 531
Figure 14.20. Interest vector field I3 with code of conduct ................................... 532
Figure 14.21. Interest flow I1 for Ex. 10 ............................................................... 540
Figure 14.22. Player passion P1 for Ex. 10 ............................................................ 540
Figure 14.23. Player interest I1 for Ex. 11............................................................. 541
Figure 14.24. Player passion P1 for Ex. 11 ............................................................ 541
Figure 14.25. Player passion P1 for Ex. 12 ............................................................ 542
Figure 14.26. Player passion P2 for Ex. 12 ............................................................ 542
Figure 16.1. Attack-Defense f0AD model q z  0, 0, z  ........................................... 596
Figure 16.2. Attack-Defense f0AD model p  0, 0, z  ............................................ 596
Figure 16.3. Attack-Defense f0AD model  z  0,0, z  ............................................ 597
Figure 16.4. Attack-Defense f0AD model e  0,0, z  ............................................ 597
Figure 16.5. Attack-Defense f0AD model initial flow V a  0,0, z  ......................... 598
Figure 16.6. Attack-Defense f0AD model initial intercept U a  0, 0, z  ................. 599
Figure 16.7. Attack-Defense f0AD model composite payoff ω  0, 0, z  ................ 599
Figure 16.8. Attack-Defense f0AD model seasonal payoff f  0,0, z  .................... 600
Figure 16.9. Attack-Defense f0AD model frame transformation E a x  0,0,0,  .... 601
Figure 16.10. Attack-Defense f0AD model frame transformation E a y  0,0,0,  .... 602
Figure 16.11. Attack-Defense f0AD model frame transformation E a z  0,0,0,  .... 602
Figure 16.12. Attack-Defense f0AD model frame transformation E a  0,0,0,  .... 603
Figure 16.13. Initial phase space for characteristic potential for f0AD model ........ 604
List of Figures xxvii

Figure 16.14. Initial characteristic potential gradients  a z  x , y ,0 


for f0AD model ................................................................................. 605
Figure 16.15. f0WWW model initial flows ............................................................ 612
Figure 16.16. Frame rotation in the f0WWW model .............................................. 614
Figure 16.17. f1WWW model initial flows ............................................................ 615
Figure 16.18. f1WWW model energy flow velocity ............................................... 616
Figure 16.19. f1WWW model time flow velocity ................................................... 616
Figure 16.20. Transfer function phase gradient for travelling wave
inequality flow in f1WWW model ................................................... 620
Figure 16.21. Travelling wave inequality flow of forced harmonic component.... 621
Figure 16.22. Frame transformation for travelling wave in f1WWW model ......... 622
Figure 16.23. Transfer function phase gradient for the standing
wave in the f1WWW model ............................................................. 623
Figure 16.24. Standing wave inequality flow of forced harmonic
flow in f1WWW model .................................................................... 623
Figure 16.25. Seasonal phase gradient in f1WWW model ..................................... 624
Figure 16.26. Frame transformation for standing wave in f1WWW model ........... 625
Figure 16.27. Standing wave recurrence plot for inequality flow
in f1WWW model ............................................................................ 626
Figure 16.28. f0AD model streamline surface for x  y  0 .................................. 629
Figure 16.29. f0AD model streamline surface x  y  1 ......................................... 630
Figure 16.30. f0AD model frame transformation E u z  0, 0, z ,  ............................. 631
Figure 16.31. f0AD model frame transformation E u z  0,0, z,  ............................. 631
1

Figure 16.32. f0AD model frame transformation E u z  0,0, z ,  ............................ 632


2

Figure 16.33. f0AD model pressure contours along the streamlines ...................... 632
Figure 16.34. f0AD model acceleration potential contours along the streamlines.. 633
Figure 16.35. f0AD model streamline contours for free fall................................... 634
Figure 16.36. f0AD model defense (blue) charge ................................................... 634
Figure 16.37. f0AD model attack (red) charge ....................................................... 635
Figure 16.38. f0AD model code of conduct charge ................................................ 636
Figure 16.39. Gauge condition surface with mesh  4 .......................................... 638
Figure 16.40. Gauge condition surface with mesh  10 ........................................ 638
Figure 16.41. Pressure in the f1WWW model ........................................................ 639
Figure 16.42. Acceleration potential in the f1WWW model .................................. 640
Figure 16.43. Streamline “tube” in f1WWW model ............................................... 641
xxviii Geometry, Language and Strategy—Vol. 2

Figure 16.44. Normal coordinate frames as seen in the orthonormal


frame for f1WWW model ................................................................ 642
Figure 16.45. f2WWW model transfer function phase gradient ............................. 642
Figure 16.46. f2WWW model travelling wave for frame transformation
E t z   cos  ,  sin  , z ,  ................................................................... 643
Figure 16.47. Streamline “tube” in f2WWW model ............................................... 645
Figure 16.48. Pressure in the f2WWW model ........................................................ 646
Figure 16.49. Acceleration potential in the f2WWW model .................................. 646
Figure 16.50. Normal coordinate frames as seen in the
orthonormal frame for f2WWW model ............................................ 647
Figure 16.51. Travelling wave recurrence plot for the inequality
flow in the f2WWW model .............................................................. 648
Figure 17.1. Communication ellipsoid for f2AD model ....................................... 655
Figure 17.2. Communication hyperboloid for f2AD model.................................. 656
Figure 17.3. Energy-momentum ellipsoid for f2AD model .................................. 657
Figure 17.4. Offset energy-momentum ellipsoid for f2AD model........................ 658
Figure 17.5. Transfer phase gradient for f2AD model ......................................... 659
Figure 17.6. Transfer modulus for f2AD model ................................................... 661
Figure 17.7. Payoff field for “blue” player .......................................................... 662
Figure 17.8. Playoff field for “red” player ........................................................... 663
Figure 17.9. Initial boundary condition for “blue” player.................................... 664
Figure 17.10. Initial boundary for “red” player ..................................................... 664
Figure 17.11. “Blue” player transverse payoff ...................................................... 665
Figure 17.12. “Red” player transverse payoff........................................................ 666
Figure 17.13. Code of conduct transverse payoffs................................................. 667
Figure 17.14. Code of conduct initial boundary condition .................................... 667
Figure 17.15. Code of conduct payoff ................................................................... 668
Figure 17.16. “Blue” player passion density plot for f2AD model ......................... 669
Figure 17.17. “Red” player passion density plot for f2AD model .......................... 670
Figure 17.18. “Blue” engagement e1  e1 density plot ........................................ 672
Figure 17.19. “Red” engagement density plot ....................................................... 673
Figure 17.20. Code of conduct engagement density plot ....................................... 674
Figure 17.21. “Blue” player interest transverse vector field .................................. 675
Figure 17.22. “Red” player interest transverse vector field ................................... 676
Figure 17.23. “Blue” reduced player strategy bias field ........................................ 677
Figure 17.24. “Red” reduced player strategy bias field ......................................... 678
List of Figures xxix

Figure 17.25. Decision pressure for the f2AD model ............................................. 680
Figure 17.26. Decision process acceleration potential ........................................... 680
Figure 17.27. Player bonding scalar field .............................................................. 683
Figure 17.28. Player shear vector field 11 ....................................................... 684
Figure 17.29. Player shear vector field  22 ...................................................... 684
Figure 17.30. Player shear vector field  12 ...................................................... 685
Figure 17.31. Player shear vector field  13 ...................................................... 685
Figure 17.32. Player shear vector field  23 ...................................................... 686
Figure 17.33. Composite payoff ............................................................................ 687
Figure 17.34. Seasonal payoff vector field ............................................................ 689
Figure 17.35. Allowed region for βn, in the f2AD model ....................................... 694
Figure 17.36. βn for the f2AD model...................................................................... 697
Figure 17.37. The dual-*βn for the f2AD model ..................................................... 697
Figure 17.38. Coordinate streamlines for the f2AD model ..................................... 698
Figure 17.39. Code of conduct passion density plot for the f2AD model ............. 699
Figure 17.40. “Blue” passion vector field f2AD model .......................................... 700
Figure 17.41. “Red” passion vector field f2AD model ........................................... 700
Figure 17.42. Code of conduct passion vector field f2AD model........................... 701
Figure 17.43. “Blue” player interest vector field f2AD model ............................... 702
Figure 17.44. “Red” player interest vector field f2AD model ................................ 702
Figure 17.45. Code of conduct player interest vector field f2AD model .............. 703
Figure 17.46. Density plot of the source I  B for composite
payoff f2AD model .......................................................................... 704
Figure 18.1. Time sequence for a typical stock Jan-Dec 2015............................. 709
Figure 18.2. Recurrence plot for typical stock time sequence ............................. 710
Figure 18.3. f2AD model flow vectors ................................................................. 716
Figure 18.4. Coordinate plot u a for f2AD model ................................................. 717
Figure 18.5. Initial flow for f1-hf AD model ........................................................ 719
Figure 18.6. Travelling wave phase gradient along z for f1-hf AD model .......... 720
Figure 18.7. Persistence of Memory by Salvador Dali [2015]............................. 726
Figure 18.8. Player charge for “blue” when “blue” is aggressor for f0AD model 731
Figure 18.9. Phase space plot of player engagements for “blue” as
aggressor in f0AD model.................................................................. 732
Figure 18.10. Phase space plot of player engagements for “red” as
aggressor in f0AD model.................................................................. 733
Figure 18.11. Active acceleration Q   x , y , 0  for f0AD model .............................. 736
xxx Geometry, Language and Strategy—Vol. 2

Figure 18.12. Active acceleration Q  x , y ,  0.1 for the f0AD model .................... 737
Figure 18.13. f0AD model of the time metric factor .............................................. 741
Figure 18.14. Initial co-moving frame acceleration q  x , y ,0  for f0AD model .... 742
Figure 18.15. Player engagements for the f2AʹD model ......................................... 747
Figure 18.16. Coordinate phase space for the f2AʹD model ................................... 748
Figure 18.17. Frequency dependence of the f2AʹD model transfer function ........ 748
Figure 18.18. Phase behaviors for the f2AʹD model ............................................... 749
Figure 18.19. Recurrence plot for the f2AʹD model relative strategy component .. 749
Figure 18.20. f2AʹD model initial strategy flows................................................... 750
Figure 18.21. Time dependence of the f2AʹD model strategy flows ...................... 751
Figure 18.22. Player aggression flows for the f2AʹD model .................................. 751
Figure 18.23. Gradient aggression strategy flow for the f2AʹD model .................. 752
Figure 18.24. Network connections recurrence plot for aggression
preference strategy for the f2AʹD model .......................................... 752
Figure 18.25. Time metric factor for the f2AʹD model ........................................... 753
Figure 18.26. Initial acceleration gradient for the f2AʹD model ............................. 754
Figure 18.27. Acceleration potential for the f2AʹD model ..................................... 754
Figure 18.28. Initial active acceleration vector for the f2AʹD model...................... 755
Figure 18.29. Active acceleration slices for the f2AʹD model ................................ 756
Figure 18.30. Pressure for the f2AʹD model ........................................................... 756
Figure 18.31. Recurrence plot for stock price ratio (dB) between two similar
companies ........................................................................................ 757
Figure 18.32. Mean price of two similar companies .............................................. 758
Figure 18.33. Price ratio of two similar companies ............................................... 758
Figure 19.1. Code of Conduct charge behavior for the f2WWW model ............ 766
Figure 19.2. “Work” entitlement payoff in the f2WWW model .......................... 767
Figure 19.3. “Wealth” entitlement payoff in the f2WWW model ........................ 768
Figure 19.4. Code of conduct entitlement payoff in the f2WWW model ............. 771
Figure 19.5. Composite payoff in the f2WWW model ......................................... 772
Figure 19.6. Phase space plot of player engagements in the f2AʹD
model on initial boundary ................................................................ 773
Figure 19.7. Phase space plot of player engagements in the f2AʹD model on
z  0.02 boundary ........................................................................... 774
Figure 19.8. Code of conduct charge for f2AʹD model......................................... 784
Figure 19.9. Code of conduct charge for f2WWW model with extended
time interval ..................................................................................... 785
List of Figures xxxi

Figure 19.10. Entitled payoff for “blue” in the f2A´D model ................................. 785
Figure 19.11. Entitled payoff for “red” in the f2A´D model................................... 786
Figure 19.12. Entitled payoff for code of conduct in the f2A´D model ................ 786
Figure 19.13. Composite payoff in the f2A´D model ............................................. 787
Figure 19.14. Charge contours z  0.2 for the f2WWW model .............................. 787
Figure 19.15. Charge contours z   0.2 for the f2WWW model ........................... 788
Figure 19.16. Engagement payoff for “work” in the f2WWW model .................... 788
Figure 19.17. Engagement payoff for “wealth” in the f2WWW model ................ 789
Figure 19.18. Engagement payoff for code of conduct in the f2WWW model ...... 789
Figure 19.19. Player shear  11 in f2WWW model ................................................. 790
Figure 19.20. Player shear  22 in f2WWW model ................................................ 791
Figure 19.21. Player shear  12 in f2WWW model ................................................ 791
Figure 19.22. Player shear  13 in f2WWW model ................................................. 792
Figure 19.23. Player shear  23 in f2WWW model ................................................ 792
Chapter 1

Introduction

Decision making is a distinctly human process. We constantly make


decisions. What time will we get up in the morning? What activities will
we do during the day? What product will we make and sell? What
products will we buy? What will we eat for dinner? Who will we vote
for? How will we defend against terrorists? The list is endless, covering
the entire range of human activities. Our rather surprising ability is that
we make decisions without apparent effort. We sift through the possible
choices without seeming to enumerate them. We live with the
consequence of our decision and learn from our mistakes. We recognize
key patterns of decision making and take appropriate action. We are firm
believers that our decisions are optimal and our decision process
adequate for our needs.
Given this state of affairs, is there any need to improve our decision
making process? I assert that the answer is yes. There are times when our
decisions lead to disaster. In economics, apparently sound business
decisions can lead to recessions and depressions. In business dealings,
talented organizations behaving rationally still can build products that
don’t sell, can fall behind on their delivery schedules and can run out of
money. As individuals our successes can lead us to over indulge; we can
become over-weight or can run up large credit card debts.
We may be adept at responding quickly to certain situations such as
predator attack; situations that are local in time and space. We are not as
adept at recognizing situations that involve correlations of events over
large strategic distances or over long time durations. An engineering
analogy might be helpful. We might be good at stacking a few building
blocks on top of one another, but less adept at building towers of 20 or

3
4 Geometry, Language and Strategy—Vol. 2

200 or 2000 such blocks. Towers quickly thrown together may last a few
weeks but not a few centuries. We have developed engineering practices,
i.e. rules of behaviors, which provide guidelines for building large or
complex structures that are built to last. These guidelines help us build
structures that would have been impossible to imagine before
engineering. I grew up in Los Angeles that limited buildings to 12 stories
because of fear of earthquakes but today, the city allows significantly
taller skyscrapers that have a better chance of surviving than structures
built a hundred years ago.
I believe there is a need for an approach that treats decision making as
an engineering discipline to address those problems that involve large
distances or time frames. We will have to define distance more precisely,
but what I have in mind is a strategic distance as opposed to a purely
physical distance. I will argue that seemingly unrelated past choices may
in fact influence the same decision.
This is Vol. 2 of a project to treat decisions as deterministic
processes. Though others have started this conversation, in Vol. 1
[Thomas G. H., 2006], my contribution was the proposal that the
dynamic properties of decision processes are best described by the
mathematics of differential geometry. Illustrative examples using
streamlines showed that decision processes would exhibit not just
persistent behaviors but oscillatory behaviors analogous to weather
patterns with hurricanes, tornadoes and other vortical behaviors.
After completion of that project, it became obvious that there in fact
might be additional solutions of interest. Volume 1 extended the
essentially DC behavior of game theory to an AC version of discrete
steady state solutions. To understand issues with larger time frames and
strategic distances, there are further extensions. I view the discrete
solutions obtained in Vol. 1 as analogous to resonant phenomena in
physical systems. Thus one might expect that in addition, there would
also be transient and steady state AC solutions of any frequency that are
driven by the initial and boundary conditions of the system. It is the
intent of this volume to focus on such solutions and show that they do
exist. These solutions no longer correspond to purely local behaviors, but
may exhibit global characteristics that no longer mimic the purely local.
Introduction 5

To understand these additional solutions, readers will need additional


mathematical tools and background. Therefore that will be provided here
from the areas of physics, differential geometry and engineering. In
addition, we supply further examples from the literature to show that
these additional behaviors provide meaningful insight. As a further
extension of Vol. 1, we expand the vocabulary to cover these new
behaviors as a necessary step to develop an engineering discipline of
decision processes.
We organize this volume into three parts. In Part 1, Chaps. 1-6, we
layout the theoretical foundations starting with the descriptive basis of
game theory, which we also take as our descriptive basis. In Part 2,
Chaps. 7-13, we inquire into the nature of the theory with decision
engineering scenarios from the prisoner’s dilemma and Robinson Crusoe
economics and conclude with a general review of game theory and
economic approaches. In Part 3, Chaps. 14-19, we analyze non-trivial
numerical solutions for systems with three active strategies. This
includes decision processes with more than one person, each of which
can have one or more strategies. Such solutions involve larger distances
and intervals of time. They elucidate the engineering issues that arise
when analyzing decision processes based on our deterministic
perspective. This third part illustrates globally non-trivial AC behaviors
and shows the reader how the theory can be applied in a variety of
situations.

1.1 Pedagogy

One of the goals of this volume is to make a deterministic approach to


decision making accessible to any student who is willing to put in the
same effort expected of any student in an engineering discipline. Such
disciplines are based on the scientific method and emphasize an
understanding of measurements and the calculus needed for deterministic
theories. As with any scientific theory, it is crucial to compare results
against experimental facts. This volume provides the necessary tools to
carry that out.
6 Geometry, Language and Strategy—Vol. 2

The scientific method provides a common ground for discussion so


that the ideas become public rather than private. Arguments can be
carried out between people with substantial disagreements, with a
reasonable chance for a successful resolution. It is not a discipline where
theories are chosen based on individual desires. This common ground
approach has proved successful in science and we extend it to decision
making. We assume that the theory is a framework for all decision
making discussions without exception.
We thus propose more than a mathematical model or a narrow
theoretical set of ideas that are fine tuned to describe a single
phenomenon. We propose a systematic way of approaching a body of
phenomena: which lies first in the physical domain no less than the
phenomena of planetary orbits; which lies second in the biological
domain of physical sciences such as molecular biology; and which lies
third in the social and psychological domain that we assert is also
governed by scientific laws. The theory must therefore be of local
applicability and global scope. In short it must be a scientific theory
consistent with the physical sciences.
It is within this context that we approach decision engineering1 as a
calculus-based discipline; calculus-based since we are focused on how
behaviors change continuously over time and space. As such, this
volume can be used as the basis for an engineering course on decision
processes. A necessary prerequisite for this course is the same as an
upper division course in mechanical engineering, electrical engineering
or bio-engineering or their equivalents. Typically an engineer is required
to take two years of calculus and one year of the physical sciences and
sufficient courses in each of the engineering disciplines to have a
balanced view of engineering. For this book, we have assumed the same
prerequisites. It is desirable but not necessary that in addition, the
students have taken a course in electromagnetic fields and a business
course in the theory of games.
The core course might be taught in three 10-week quarters or two
semesters. We envision that the first quarter session of the course would

1
We indicate by bold-italics distinctions that we believe are new ways to speak about
decision processes.
Introduction 7

cover Chaps. 1-4, Part 1, the conceptual theoretical foundations of


decision process theory. The foundations are based strongly on an
electrical engineer’s view from the 21st Century modified to include the
post-Newtonian formulation of mechanics by Einstein and a modern
understanding of fields from quantum mechanics that replaces classical
probability ideas. An understanding is required, not a detailed working
knowledge of how to do computations in these difficult areas.
The second session of the course would cover Chaps. 5-6, the
theoretical framework and working models of decision process theory.
This session provides the technical essentials as well as reference
materials and can be covered in great detail (graduate level course) or as
an overview (undergraduate level course). It would also cover Part 2,
Chaps. 7-13, including detailed and solved examples of the Prisoner’s
Dilemma (Chaps. 7-9), Robinson Crusoe scenarios (Chap. 10) and
critical discussions of the basis of the theory (Chaps. 11-13).
The third session would include Part 3, Chaps. 14-19 on how to apply
and extend the decision process theory to realistic applications starting
with general two-person games (Chaps. 14-16). We suggest that the
numerical work be computed using Wolfram Mathematica® software
[Wolfram, 1992], which we have utilized throughout the book.
The third session would return and review Chaps. 11-13, key
concepts in the literature and in the process take on a new perspective
that game theory provides a set of known behaviors (Chap. 14) and
persistent behaviors (Chap. 4). The latter we associate with code of
conduct; which applies to the prisoner’s dilemma and changes the
foundational assumptions of our game theory starting point. The third
session would conclude with an exploration of stable structures that are
consequences of decision process theory (Chaps. 18-19), which highlight
the global aspects of the theory and provide a starting point for future
research. As a pedagogical guide, each chapter contains a specification of
the outcomes expected for the student, an exposition of the ideas of that
chapter and a list of problems or exercises that take the student further.
Though it is desirable for the student to have the engineering and
business prerequisites, every attempt is made to make the book self-
contained. We make no assumption that the reader is familiar with the
ideas of differential geometry or with their applications in physics. We
8 Geometry, Language and Strategy—Vol. 2

also assume that the reader is not familiar with the vast literature on
game theory, which is a rather specialized aspect of economic theory.
This presents challenges that are not unfamiliar to those in
engineering and science. On the one hand we must provide motivation to
the reader to make the effort to learn the new techniques and on the other
hand we must provide adequate tools and training for the reader to be
able to apply and use those techniques on interesting problems. The
payoff in the end is the possibility of being able to vastly improve the
readers understanding of real-world decision processes.
Thus the goal for this volume has been to select and bring relevant
components of the vast literature on differential geometry, including the
physics of gauge theories, to the reader interested in decision processes.
The approach is similar to bringing the ideas of physics to engineering
students who want to apply the ideas to realistic problems. We require
only that the student have sufficient sophistication in mathematics and
science. One way to gain that sophistication is to follow the course of
study that has proved successful in engineering schools.
For this reason we have suggested the requirements for a sophomore
level engineering student. The goal in this volume is to make the ideas
accessible to a student who has followed such a course or its equivalent.
We do this at three levels. We provide an exposition of the ideas, we
provide a set of problems to enhance the student’s understanding of the
material and we provide numerical examples with the help of
Mathematica. The latter provides an important visual tool for making the
ideas come alive.
We show how numerical solutions to the decision process equations
can be obtained for numerous illustrative examples. The equations are
Einstein-like partial differential equations and so the techniques and
solutions obtained also may be of interest to those in other fields such as
relativity, gauge theory and differential geometry. As the purpose of this
book is to provide the necessary mathematical and engineering training
to both understand and apply the approach, out of necessity it is cross
disciplinary and hopefully sheds light on the areas from which it
borrows.
We turn now to an introduction of our goals for decision process
theory.
Introduction 9

1.2 Decision Process Theory Goals

Decision process theory is meant to be a complete, comprehensive and


quantitative theory of the time dependence of decision processes. For the
quantitative aspects, we first make clear which are expected to be
deterministic and which are not. Such a distinction is needed even in
mechanics, where we separate out distance and time as quantities that are
subject to deterministic study and their measured value as something that
may vary within constraints due to the uncertainties of the measurement
process.
For decision process theory, the strategic choices in a suitable space
play the same role as distance in a physical theory. The uncertainty of
choice plays the same role as measurements. We make no claim to know
which choice a person will make any more than we claim in physics to
know what distance will be read on a ruler when measuring a room. In
physics, we believe that there is a distance independent of the
measurement process. In decision theory, we believe that there is a
strategic mixed strategy whose value exists outside of the valuation
process.
Thus we live with the uncertainty of choice as something outside of
the theory and deal with the systematics of distance and time as causal
aspects that are consequences of the theory. This approach is different
from much of modern game theory but agrees with the original
assumptions of Von Neumann & Morgenstern [1944].
The theory is also meant to be comprehensive. We apply this theory
to such diverse issues as the prisoner's dilemma in game theory and
ethical issues such as the tragedy of the commons using techniques
borrowed from electrical engineering, meteorology and general
relativity.
The theory is meant to be complete. The goal of the theory is to cover
all aspects of economic behaviors. To do this we need to examine all
aspects of the theory. In this volume, we make a major dent in that we
review the static and equilibrium behaviors described by game theory,
which are analogous to DC circuit behaviors. We extend such static
behaviors to steady state behaviors analogous to AC circuit behaviors.
10 Geometry, Language and Strategy—Vol. 2

The full treatment includes additional dynamic behaviors, including


transient dynamic behaviors.
The theory is meant to be computable. We borrow from engineering
in part since these disciplines are computable. Why is this so? We
believe it is because they are complete and quantitatively describe
deterministic processes. What insights do they provide? Can those
insights be applied to the realm of decision making? These are the core
questions for decision process theory.
We believe the answers are related to engineering’s intent to
extrapolate, in a useful and practical way, the behavior of events from the
local to the global both in space and in time. It is taking the qualitative to
the quantitative. In other words, engineering makes quantitative and
“correct” predictions from a limited starting point! It accomplishes this
using the scientific method along with the calculus-based mathematics of
Newton that makes concrete the continuous and topological
interconnection of events. Engineering works because is rests on a
deterministic set of theories that are constantly evolving using the
scientific method.
The scientific method provides the common base to collect,
understand and organize the phenomena around a theoretical framework.
The method provides a basis for reconciling failures of understanding
and feeding that back so the framework can be improved. This
framework has evolved to the point where it can make extrapolations
from the local to the global using accurate data based on detailed
measurements. The extrapolation is based on rules that have been vetted
by historical precedent. Thus engineering has codified how things are put
together, how long and whether they hold their shape and how changes
occur over time. In other words engineering is based on attributes of
forms that are both persistent over time and space, as well as on
attributes that exhibit organized variability over time. The strength of
engineering is a practical encapsulation of algebra, topology and
geometry, whose power and reach in mathematics and science has been
long argued by Bourbaki [1968].
We believe that it is possible to apply the engineering discipline to
decision making processes, despite knowing the reluctance of many to
accept the possibility that human processes are described by the same
Introduction 11

tools of science that describe physical processes. We believe there is a


solid scientific case that decision processes are deterministic. Decisions
are actions that occur in the physical world. These actions occur in
response to forces without need of postulating any other agents. What
may not be apparent is how the reasoning from the physical sciences
deals with human complexity. Yet many have already applied such
reasoning in game theory and economic behavior, as for example those
following the work of Von Neumann & Morgenstern [1944].
You may counter however that these works rest on a static view of
the world. Moreover, modern game theory predictions about future
behavior are based on an understanding of random events using
Bayesian probability theory [Bayes, 1764]. Recent work by Barabási
[2003] however on networks takes an opposite view. First of all networks
provide the topological view of whatever underlying mechanisms might
exist. The recent advances indicate that the networks are not random but
behave according to laws of nature that are not unlike those in the
physical sciences. It strongly supports the idea that behaviors are not
random but deterministic. Though that work focuses on networks it
provides support for our deterministic approach using the ideas of
differential geometry. Indeed, the theory of algebraic topology
[Eilenberg & Steenrod, 1952] provides details of how algebraic
topological properties (graphs and networks) are exhibited in topological
theories. Such network examples from the World Wide Web provide
wonderful examples of why the approach taken here might have
relevance to the real world.
The modern approach of game theory is one way to generalize from
the original static game theory approach. We suggest a different
approach. Our view is that game theory is a physical theory and physical
theories generalize from static theories to dynamic theories without
resorting to the Bayesian stochastic approach. In electrical engineering
we move from DC circuits with static behaviors to AC circuits with
steady state behavior, for example. The Bayesian approach is replaced by
the physical principle of least action. Our goal is to show in detail how
this is done.
We hope to demonstrate that decision making is an engineering
problem, one in which the decision process exhibits a characteristic
12 Geometry, Language and Strategy—Vol. 2

behavior that is intrinsic to the process, as well as a forced behavior set


by the initial conditions. This is in exact analogy to an electrical
engineering network with an applied load being forced by a source.
There too we see both a resonant or characteristic behavior (discrete
frequencies) along with a forced behavior (arbitrary frequencies). We too
find it useful to study such behaviors by considering the harmonic
steady state waves of the system, just as it is useful in electrical
engineering to use phasor harmonics.
The concept of stationary flow is one in which characteristic and
forced vortical behaviors play a central role. Equivalently we call such
behaviors steady state flow. Not only is there an electrical engineering
analogy, there is also a close atmospheric physics analogy to these
equations, namely the equations describing the fluid flow of the
atmosphere in which jet streams, tornadoes and hurricanes provide a
distinct aspect of the observed weather. They may in fact be steady state,
yet they are not simple extensions of linear local behaviors: such global
behaviors such as turbulence are not visible when examining only the
local behaviors.

1.3 Introduction

As an introduction to Part 1, in the rest of this chapter we lay out a


foundation for the game theory and physics concepts that will be needed.
In Sec. 1.4 we illustrate the type of interesting economic problems we
might study, such as the prisoner’s dilemma. They introduce the physical
foundation for the theory. In later chapters, we demonstrate complete
solutions that follow based on this foundation and find that our results
suggest answers to dilemmas that have been noted in the literature. For
example, our solution of the prisoner’s dilemma contradicts the game
theory intuition that the prisoners would be best off if they cooperated
with the authorities and confessed. We will be able to demonstrate under
what conditions this intuition is in fact supported, providing an example
in which we gain new insight over past theories.
In Sec. 1.4 we put into context that a large number of problems in
game theory are two person games in which each player has two strategic
Introduction 13

choices. Such problems are completely solved in game theory and are
also completely solved for the dynamic theory, which we demonstrate in
later chapters. In this chapter, we provide the context for economic
situations with any number of players and any number of strategic
choices.
In order to find solutions to such problems, in the next sections we
look at classical physics and how the approach might relate to decision
process theory. The equations of classical physics are derived from
differential geometry and we anticipate from Vol. 1 that the decision
process theory equations are as well. We start by looking at the physics
equations in the same way as electrical engineers. We provide a brief
survey of the relevant ideas starting from Newtonian mechanics (Sec.
1.5) to the theory of Maxwell (Sec. 1.8). The latter theory is particularly
relevant to decision process theory.
We find in later chapters that decision process theory equations are
similar to equations in general relativity, so our approach of harmonics
taken from electrical engineering (phasors) may be unorthodox even to
physicists and we think a novel approach to solving equations from
general relativity. Thus an introduction to physics may be helpful to the
more general reader.
We present decision process theory following the techniques of
Lagrange (Sec. 1.6). For the expert, this indicates what we intend; for the
non-expert we will say in sufficient detail what we mean so that you can
understand and use the concepts. We argue that it is useful to break
forces into those that are applied and those that are constraints, allowing
the number of variables to be drastically reduced. These techniques relate
to the key physical principle, the principle of least action, which we use
to create a causal theory. Without losing any generality, we can describe
processes using variables that are applied and matched to the problem.
For decisions we suggest that the key variables are the strategies
available to each participant in the decision.
Given the principle of least action, we bring together the key
concepts of decision making (such as strategies and utility, Secs. 1.7-1.8)
and show that they can be described by calculus-based rules that allow
the extrapolation from the local to the global, both in time and space. The
14 Geometry, Language and Strategy—Vol. 2

extrapolation relies on detailed local measurements (Chap. 2) from which


one gains insight into the global attributes of the decision process.
The key aspects of this process are firstly normative, which we
borrow from game theory: every decision can be put into a standard or
normal form so that its aspects can be measured. Secondly, every
decision involves choices some of which are persistent, unchanging and
inactive, the rest are variable and active. Thirdly, the decision processes
are either the local processes that we deal with so successfully or the
global processes that we deal with less well and often lead to unwanted
results. These attributes can be put together in a common decision
process theory in which individual choices result in collective decision
behaviors that are vortex-inducing. In this volume, we describe in
substantial detail how to extend and apply this theory to decision
processes from both a theoretical position and a practical and
computational position. Our goal is to explore in depth the underlying
mechanisms behind these vortex-inducing behaviors.

1.4 Foundational Game Theory

To make the conceptual basis for cause and effect relevant to decisions,
we frame decisions as events that occur in some space at a moment in
time. Here we review attempts to do this and find that foundational
aspects of game theory provide the perfect vehicle in which to express
our decision process theory.
We start with a foundational view of decision process theory and its
relationship to engineering and to mathematical and physical theories.
The foundational concept from game theory that we find appealing is
that every decision process can be represented in a standard way and put
into a normal form, an insight of Von Neumann & Morgenstern [1944].
They prove this under very general considerations, creating a theory of
games and economic behavior, where games in their context go beyond
recreational games to include real world economic behaviors. They
envisioned that this way of looking at decision processes provides the
basis for a quantitative way of dealing with social phenomena. See for
example, Myerson [1991], for an undergraduate course on game theory
Introduction 15

and Mas-Colell, Whinston, & Green [1995], for an undergraduate course


on microeconomics.
We note and will have reason to refer to other work, such as Nash
[1951], who improved and extended game theory and who also focused
on the equilibrium and static attributes of the decision process. If we
think of choices that change in time as flows, then we can say that these
theorists characterize the flows as being the static remnants after all
transient effects have died out. These theorists actually go further in that
they assume that the flows are the same everywhere; the flows are
equilibrium values akin to DC currents in a circuit.
In our view, in order to address questions such as recessions, cost
over-runs and personal bankruptcies, we need to understand flows akin
to AC currents as well as the transient behaviors; we need to understand
the way decision processes depend on time. In this sense our approach
diverges from the equilibrium focused literature. We should share some
common ground with the time based theories and models. We return to
this in Chap. 11 and provide an historical context.
In game theory, the normal form consists of pure strategies that
represent the full set of choices that need to be made for a game
(decision). Thus in a chess game, in principle you could lay out what
move you would make for every possible move your opponent might
make for the duration of the game. You would need an enormously long
piece of paper! The paper would capture a decision tree, with each
possible path down the tree representing a pure strategy. Your opponent
would make a corresponding list. The play of the game would consist of
each person picking one of the pure strategies. Game theory recognizes
however that when a game is played many times, there is a disadvantage
to one side if the other side can predict what choice you might make.
There is an advantage of playing a mix of strategies. In the static
context of game theory, it is envisioned that all transients have died out
and the player will have learned to pick a particular mixed strategy by
choosing the pure strategies with frequencies or weights that are known.
Under reasonable conditions, a game then has an equilibrium mixed
strategy for each player of the game in the sense that no player can do
better than that mixed strategy. At this point we go no further with game
theory since we need only their concept of pure strategies and a modified
16 Geometry, Language and Strategy—Vol. 2

view of their mixed strategies that we describe below. However, we note


that our theory must include game theory as a special case of equilibrium
behavior. Still, we must develop the framework in more detail before
attempting to understand what it might mean to have an equilibrium
strategy or what “doing better” implies.
We note that we ascribe a reality to the mixture of strategies a player
might pick. This is something that in fact might be part of a theory and
evolve according to a deterministic process. What we don’t know is
which strategy of that mix a player will choose at any given moment. It
is in this sense that this separation of chance is like separating out the
measurement process in physics when determining the length of an
object as opposed to asserting that there is a length that may change as
part of physical processes.
In creating a new decision process theory using the foundational
ideas from game theory, we thus make two additional assumptions. First,
we view mixed strategies not as fundamentally random choices, but as
the frequencies of successive decisions averaged over a sufficiently small
sampling time. This is consistent with the original assumptions of Von
Neumann & Morgenstern [1944], though not with much of the more
recent literature (see Chap. 7). The choices may or may not be made
randomly. In general every decision is the choice of a mixed strategy.
Second, we consider decisions as social phenomena carried out by
many players, each making what they consider to be independent
decisions. They are aware however, not only of the outcome for
themselves, but have their own view of the outcomes of their neighbors.
We believe it is plausible that decisions made in one strategic part of
space at one moment of time will therefore propagate not only forward in
time but outward from this point. This is a new concept since it implies
not only a causal relationship between events, but a neighbor or network
relationship.
The characteristics of the propagation should be specified by the
theory. This is analogous to the specification of the wave properties of a
lake or sea given knowledge of some initial disturbance at some point
and at some time. In the next two sections we make this analogy more
concrete.
Introduction 17

1.5 Newtonian Mechanics

The physical world consists of a variety of phenomena including


decisions. Before dealing with decision processes, it is instructive to
review how classical Newtonian mechanics deals with physical
phenomena, which we do in this section. We are particularly interested in
how classical theory deals with phenomena in which some aspects are of
interest and others, such as constraints or symmetries, must be accounted
for but the details are not of interest. We deal with how Newtonian
mechanics deals with complexity. We argue that we might deal with
complexity in the same way for decision processes.
Newton’s famous equation F  ma summarizes three separate laws.
First it says that a body at rest stays at rest. Second it says that bodies
that move with constant velocity (zero acceleration) continue to do so if
they are not subject to an outside force. Third it says that for every
action, there is an equal and opposite reaction. In other words any
imbalance in the forces results in acceleration.
Newton’s equation depends on the second order rate of change of the
position r   x y z  in terms of the x, y and z coordinates and the
 
vector F  Fx Fy Fz that represents the force along each of these
directions respectively:

F  mr . (1.1)

We represent the rate of change using a “single dot” so that r is the


velocity and the “double dots” r represents the second order rate of
change, namely the rate of change of the velocity, the acceleration.
Newtonian physics reduces every problem, no matter how complex,
to the use of this fundamental rule for each single body, where the
proportionality constant is the mass of the body. The mass represents the
inertia of a body. It is an observational fact that a massive body
accelerates more slowly from rest than a light body.
Mass has an experimental basis—it can be measured in units of the
mass of smaller bodies—yet it does not have a clear origin other than the
above equation. Inertia is a property of complex systems as well as
simple, since the complex systems are made from the simple ones.
18 Geometry, Language and Strategy—Vol. 2

Indeed, one might think of inertia as a part of the system that involves
many degrees of freedom about which we know little.
We rarely deal with a single (“rigid”) body in physics. The simplest
interesting case is two bodies attracted to each other by a gravitational
force. Newton observed that the force is proportional to the masses of
each body and inversely proportional to the square of their distances. In
this way he could show that the theory would agree with the observations
of Kepler that equal areas would be swept out in equal times.
A success of Newtonian mechanics is the accuracy of the predictions
that result from these two statements. A success of the scientific method
on which the theory rests is that as challenges to the theory have
occurred, the framework for discussion has remained or evolved in a
systematic way. For example it was discovered by Einstein that the
above examples contain a fatal flaw for Newtonian mechanics, since they
implicitly assume that the influence of one body is felt instantaneously
by the other body. More careful observations show that action at a
distance is not instantaneous, but is an effect that travels with a finite
speed. A related flaw is the notion of a rigid body, which again would
require the influence to travel instantaneously.
Despite the successes for the motion of two bodies under
gravitational influence, such as the motion of the earth around the Sun
and the belief that the laws hold generally, the Newtonian formulation as
it stands does not deal directly with many real world classical situations.
A car moving along the highway might be idealized as a single object,
but the road on which the car moves is an important contributor to the
forces acting on the car. Using Newton’s formulation, you must account
for all the forces, including those keeping the car from sinking into the
pavement. Calculations that include these effects are long and once
complete, have information that is not normally of great interest, such as
knowing the value of the pavement force.
It is at this point that we see a connection to decision theory. Here too
we have many forces to deal with, but have a suspicion that their
inclusion will yield information that may not be of great interest. One
might be tempted to argue that the forces are not knowable and use
probability arguments. However, we shall stick to the deterministic
approach using the ideas of Newton.
Introduction 19

Being able to formulate problems in a more effective way was


addressed over the years and centuries after Newton. The starting point
was to write out (as a thought experiment) all of Newton’s equations, one
for each body in the problem. This means idealizing in some cases a
body as being made up of infinitesimally small parts. For the car, this
would idealize the highway as being made up of lots of small bits, each
characterized by a mass, acceleration and the net force acting on that bit:

Fk  m k rk . (1.2)

This represents many (second order differential) equations and takes into
account all the forces.
The forces of interest are applied and external (the motive force
moving the car forward) and the forces that are not of interest are
constraints or internal (the gravitational forces holding the car to the
road, the atomic forces preventing the car from sinking into the road as
well as those holding the bits of the car together, and possibly symmetry
forces such as the car is on a planet that is spinning). These definitions
are not absolute, but relative to the interest of the person solving the
problem.
A rigid body is held together by internal forces and such forces
normally do no work (see below). A formulation that even partially
excludes such internal constraining forces is advantageous. To appreciate
this approach, it is helpful to introduce the concept of work, which is
defined as an integral made up of the sum of the product of the force
times incremental distances along a specific path. Work has the
dimensions of energy. Newtonian mechanics distinguishes two types of
energy.
First there is energy associated with motion as kinetic 1 2 m r  r
defined as one-half the product of the mass and the sum of the squares of
the velocity components.
Second there is energy associated with forces, as potential energy:
typically the variation of the potential energy with distance along a given
axis determines the force along that axis. In both cases energy has
dimensions of a force times a distance.
20 Geometry, Language and Strategy—Vol. 2

For so-called conservative systems, a consequence of Newtonian


mechanics is that the sum of the kinetic and potential energy is constant.
This is the case for two bodies interacting through the gravitational field.
The potential energy in this case is proportional to the mass and inversely
proportional to the distance. So how does this help us? We take this up in
the next section.

1.6 Newtonian Mechanics, Lagrange and Game Theory

The idea is to find a formulation of mechanics that would allow us to


apply those ideas to systems that are complex, such as decision process
theory. Such a formulation is Lagrange’s, which starts with Eq. (1.2) and
derives a new equivalent set of equations in which only the active and
external forces are present. For a standard treatment, and the one we
follow, see Goldstein [1959].
Lagrange provides a general way of formulating mechanics that has
proved useful even in areas where Newtonian mechanics breaks down as
suggested by Einstein. In the Lagrange formulation, first we consider not
the acceleration as given by the double rate of change of the coordinate,
but consider the acceleration in terms of the momentum defined as
p k  mk rk , the product of the mass and the velocity.
The momentum is first order in the rate of change of the coordinate
(the “velocity”) and the force is first order in the rate of change of the
momentum:

Fk  p k . (1.3)

Newton’s equations are second order in the derivative, whereas Lagrange


separates them into two sets of coupled first order equations.
This may not seem like an improvement, but mathematicians have
found this effective, so bear with us. They often use this device when
studying differential equations of arbitrary order: they reduce the
equations to a set of first order equations, where they have created a set
of theorems for proving existence and uniqueness of solutions. For us,
the practical consequence is that we can often solve such equations
numerically using application programs on our laptops. The general
Introduction 21

theory of solving such equations demonstrates that there will always be


unique solutions when the initial value of each first order equation is
specified. The real advantage however, is that we can more easily
separate out the active from the constraint forces.
Lagrange attacks the key question of constraints with the goal of
eliminating the variables that are of no interest. He considers the work
done on the system if the coordinates undergo a small displacement
[denoted by  rk ]:

F k
k  p k   rk  0 . (1.4)

There are applied forces and constraining forces. If the constraining


forces do no work2 and if the constraints are expressible as formulae
involving only the coordinates f  r1 , , rN , t   0 (called holonomic
constraints), then only the applied forces appear in the above
summation:

  F 
k
a
k 
 p k  rk  0 , (1.5)

where F a  k are the components of the applied force. The positions of the
objects are not independent but constrained.
The Lagrangian point of view is that any set of independent
coordinates q  (not necessarily equal to the above rk coordinates) can
be used to rewrite the above equation. Lagrange showed that the
equations can be written in terms of the kinetic energy T now expressed
in terms of the independent forces and the applied forces Q as they
relate to these independent coordinates:

d  T  T
   Q . (1.6)
dt  q  q

The derivation of the above equation involves standard mathematical


manipulations and relates these derived quantities to the original
Newtonian quantities:

2
The constraint force does no work because the force moves through no distance.
22 Geometry, Language and Strategy—Vol. 2

T  1 2  mk rk 2
rk . (1.7)
Q   F  a  k 
q

The resultant Lagrange equations are an exact restatement of Newtonian


mechanics involving only the active forces and the independent
coordinates.
The subtle shift is away from the original set of all forces, to the set
of applied forces that govern the motion of interest and the kinetic energy
of the system including all the components, written in terms of an
independent set of variables. We emphasize that this approach is really
equivalent to the Newtonian approach as applied to the applied forces. It
provides a solution for the constraint or internal forces in the special case
that the special independent variables are taken to be the original
coordinates, in which case it is assumed that the problem has been
specified in its entirety with no constraints.
The expression obtained by Lagrange may also be written in terms of
the kinetic and potential energy. For the important class of problems in
which there is a potential energy V , the applied force Q is given in
terms of the independent variables:

V d  V  V d  V 
F
a
k      Q     . (1.8)
rk dt  rk  q dt  q 

In this case the equation of motion is expressed in terms of the difference


between the kinetic and potential energy, called the Lagrangian
L  T V :

d  L  L
  0. (1.9)
dt  q  q

We see that all of the information about the problem is now reflected in
the Lagrangian that has dimensions of energy.
Further these equations of motion solve the problem posed, namely
the equations involve only the applied coordinates, though all aspects of
the kinetic and potential energy are taken into account. In Sec. 3.1, we
Introduction 23

establish that the Lagrange equations of motion can be formulated in


terms of a principle of least action.
In the case of a car moving along a highway, though the motion
occurs in three dimensions, the configuration space needed to describe
the situations is 3N where N is a truly large number representing the
number of bodies interacting (which is of the order 10 24 ). Thus the
practical application recognizes that the important question is just the
motion along the two-dimensional highway (or indeed just a single
dimension if the car stays in its lane). The above view demonstrates that
there is no loss of generality focusing on the two dimensional
coordinates in determining the motion as long as the kinetic and potential
energies represent their totals faithfully.
It is therefore plausible that we deal with decisions in the same way.
The foundational aspects of game theory give us the applied strategies
that we should consider, though there may be other variables to consider.
The applied strategies are the generalized coordinates of the Lagrange
approach. As long as we faithfully include all aspects of the kinetic and
potential energy, we have a specification that takes all the applied and
constraint aspects of the problem into account. We take this view and
apply it to our conceptual basis from foundational game theory, Sec. 1.4,
in the next section.

1.7 Conceptual Basis and Payoffs

If we take as our conceptual basis for decisions only the foundational


aspects of game theory, we make little if any use of the dynamic content
that has been unearthed by several decades of work on the ideas of Von
Neumann & Morgenstern [1944] and Nash [1951]. It is our view that the
scientific method has power in its ability to incorporate ideas and that
there are many game theory ideas that need to be incorporated into any
theory for decision processes.
Our analysis suggests that a common feature of game theory models
is that decision processes are characterized by payoffs. The outcome of a
decision can be quantified numerically as a value each player receives
(or pays) as a result of each player choosing a pure strategy. The payoffs
24 Geometry, Language and Strategy—Vol. 2

for mixed strategy choices are determined from these pure strategy
payoffs. We must find a way to incorporate payoffs into our decision
process theory.
We believe therefore that we must incorporate two related concepts
from game theory: the concept of a game value and the concept of an
equilibrium set of strategies. Decision process theory is not derived
from game theory so we don’t have a precise method for adding these
concepts. Our method can be illustrated best by using an example and
showing how it is treated in game theory and how we might then
incorporate these game theory ideas into our framework. The example
we have chosen is the Prisoner’s Dilemma.
The prisoner’s dilemma has been extensively investigated by game
theorists. As an example, see Rapoport [1989]. It has been scrutinized in
both theoretical and empirical contexts. The scenario is that two
prisoners are being held for a crime where it is suspected they have acted
in concert. Each is given a choice to confess or not confess with penalties
that are supposed to induce confession of their guilt. However if neither
confesses they will get off lightly. If both confess they will be penalized
but not as severely as the case in which one confesses and implicates the
other. The dilemma is that the prisoners would be collectively best off by
not confessing. However the game theory analysis below recommends
they both confess.
We start with the formulation of the prisoner’s dilemma as a game in
normal form between two players/prisoners specified by the following
payoffs to player 1.

G121 N2 C2
N1 0.1 1 . (1.10)
C1 0 0.9

The payoff values are nominal and don’t influence the conclusions
drawn. There are identical payoffs to player 2:

G212 N1 C1
N2 0.1 1 . (1.11)
C2 0 0.9
Introduction 25

Game theory supposes that there are quantitative payoffs for the matrix
G k of possibilities as illustrated above for player k  1,2 .
We describe the meaning of the elements of the matrix in detail for
player 1, noting a similar description holds for player 2. For player 1, the
rows are labeled (C1) if player 1 confesses and (N1) if player 1 does not
confess. The columns are labeled in a similar manner (N2) and (C2).
The payoffs reflect quantitatively the posed problem: if both confess
player 1 loses 0.9 units, if both don’t confess player 1 loses a much
smaller value: 0.1. If player 1 confesses and implicates player 2 who
does not confess, then player 1 loses 0.0 units. On the contrary, if player
1 does not confess and is implicated by player 2, player 1 loses 1 unit.
Though rather simple, the example illustrates several general
properties of game theory. The payoffs for the two players need not add
up to zero; when they do, Von Neumann & Morgenstern [1944] called
them zero-sum games. Zero-sum games have been analyzed in detail and
have an equilibrium value that is argued as follows.
Suppose the payoffs Eq. (1.10) for player 1 were a zero-sum game.
The most conservative strategy for player 1 would be to determine the
minimum outcome for each of its pure strategy choices. If player 1
chooses (N1), the minimum case is 1 . If player 1 chooses (C1) the
minimum is 0.9 . The maximum of these two is 0.9 . The most
conservative strategy is to choose this max-min. There may not always
be a max-min solution however.
What Von Neumann & Morgenstern [1944] showed however, is that
for a zero-sum game with any number of players, each with any number
of strategies, there is always a max-min using mixed strategies. This is
called an equilibrium strategy. It is the best one can do acting
defensively. When there is equilibrium, there is also a game value equal
to the payoff associated with the equilibrium strategy (or weighted
payoff if the strategies are mixed).
When the game is not zero-sum and there are separate payoffs for
every player, the above argument has been replaced by Nash [1951]. He
showed that again there will be a best choice now called the Nash
equilibrium. Nash says that the best strategy has the property that no
player can do better against this strategy as long as all the other players
adhere to their best strategy. There will now be multiple game values,
26 Geometry, Language and Strategy—Vol. 2

one for each player at the Nash equilibrium. By inspection, the Nash
equilibrium for the prisoner’s dilemma is that each player confesses. The
game value for each player is 0.9 .
Notwithstanding the great insight provided by the Nash equilibrium
and the importance of considering the payoffs separately for each player,
we believe insight into the dynamics comes first from looking at each
zero-sum game and studying how one computes that equilibrium
strategy. For the example chosen here, it is possible to determine that
strategy by inspection. In the general case the max-min problem is solved
numerically. A general method is to first convert the game to a fair game
(one whose game value is zero) which is equivalent in that by
construction it has the same equilibrium value. For player 1, the general
method assumes that the fair game has one additional strategy,
commonly called the hedge, whose payoff is adjusted to get the required
equilibrium:

Fab1 N2 C2 N1 C1 H
N2 0 0 0.1 0 0
C2 0 0 1.0 0.9  m 0.9
1
. (1.12)
N1 0.1 1.0 0 0 1
m 1.0
C1 0 0.9 0 0 1
m 0.9
H 0 1 0.9
m  m1.0  m 0.9
1 1 0

This fair game produces exactly the same strategic results as the original
game.
However, computing the equilibrium value for this fair game can be
done as a dynamic process. It is of great interest to us that one of the
methods of finding the equilibrium is to write a differential equation
involving these payoffs:

dV
g   eF  V . (1.13)
d

In a different form, this method was introduced by Von Neumann and is


described by Luce & Raiffa [1957].
Introduction 27

The equation says that the rate of change of the mixed strategy is
proportional to the product of the payoff matrix and the mixed strategy.
At equilibrium, the rate of change of the mixed strategy is zero so the
product is also zero. It can be shown that the product is zero is equivalent
to the max-min condition normally ascribed to zero sum games. The
advantage of the reformulation of the game to a fair game is that the
equilibrium is defined by the product vanishing. This technique can be
done for any zero-sum game and any number of players with any number
of strategies.
We believe that we now have a specific attribute to look for in a
dynamic theory. The theory needs to provide in a natural way a payoff
matrix of the type above that is associated with each player. The notions
of equilibrium, Nash equilibrium and game value then have
generalizations in the theory. The dynamics represented by Eq. (1.13) for
a payoff matrix that is antisymmetric is analogous to electrodynamics.
We present the arguments for why this is true and relevant in the next
section.

1.8 Electromagnetic Fields as Payoffs

We find that a characteristic behavior of decision processes is a tornado


or vortex-like behavior such as one sees in weather phenomena (Vol. 1).
A fundamental goal of this volume is to explore the important underlying
deterministic mechanisms, one of which is vorticity. A key to that
understanding is the similarity between the form of Maxwell’s equations
and the form we identified in the previous section for payoffs.
However to see that similarity and appreciate its import, we have to
rewrite the fundamental laws of Maxwell in the covariant form
introduced by Einstein. We shall review the physical origins of
Maxwell’s laws and then show how the equations can be rewritten in the
covariant form. We will then see the connection.
Maxwell based his work on phenomena that in part were known in
antiquity. At the core are the ideas that there exists an attribute called
charge that is associated with matter. The charge can be positive,
negative or zero: like-charges repel and opposite-charges attract without
28 Geometry, Language and Strategy—Vol. 2

any visible matter acting as the cause or intermediary. In Newtonian


language the law of attraction is an instantaneous action-at-a-distance
force inversely proportional to the square of the distance and directly
proportional to each charge. It is customary to describe the law of
attraction as a charged body interacting with a vector electric field E
defined at each point of space. The force experienced by a charged body
is then proportional to its charge and the electric field in which it finds
itself. Like gravity, what is notable about the electric field is that it
makes things move “instantaneously” without any observable
connection.
The existence of magnetism has also been known since antiquity. The
most notable application is the use of magnetic material as a compass for
navigation. Again the effect of magnetism appears to be an action-at-a-
distance force and it is customary to describe the force in terms of a
magnetic field B . There is no particular reason to suspect that the
electric and magnetic fields or the corresponding forces are related.
However, a number of experiments with electric and magnetic
phenomena indicated that this might be so. One experiment described the
motion of charged bodies in the presence of a magnetic field and noted
that when charges accelerate they generate not only an electric field but a
magnetic field. Maxwell synthesized such observations with what was
already known into a comprehensive theory that bears his name.
He provided a complete and self-consistent set of equations that
summarized the laws of how a charge density  produces an electric or
magnetic field based on its motion V and how the magnetic and electric
fields influence the motion of charged particles:
Introduction 29

 x Ex   y E y   z Ez  
 x Bx   y B y   z Bz  0
1 Bz
 x E y   y Ex  
c t
1 Bx
 y Ez   z E y  
c t
1 B y . (1.14)
 z Ex   x Ez  
c t
1 E z 1
 x B y   y Bx   Vz
c t c
1 E x 1
 y Bz   z B y   Vx
c t c
1 E y 1
 z Bx   x Bz   V y
c t c

The first equation generates the law of attraction for charges (Coulomb’s
Law); the second equation expresses the fact that there have never been
any magnetic charges observed; the simplest magnet always comes as a
dipole.
The next three equations reflect Faraday’s Law that changing
magnetic fields induce electric currents and hence an electric field. The
last three equations reproduce Ampère’s Law that moving charge
generates a magnetic field. The last three equations however are not
identical to Ampère’s Law because of the presence of the “conduction
current,” the extra gradient of the electric field with time. There is no
contradiction since the early experiments were done using steady state
electric fields and the extra term would have been zero.
Maxwell realized that without the extra term, charge would not be
conserved for time varying fields. With the extra term he got a continuity
equation similar to the equation for conservation of mass:

d
  jVj  0 . (1.15)
dt
30 Geometry, Language and Strategy—Vol. 2

We introduce the convention that for repeated indices j we sum over the
possible values x, y, z .
The conservation law agrees with the observation that charge does
not spontaneously appear or disappear: its flow in or out of a box or cell
in space results in an increase or decrease of the charge inside that cell.
So, like mass, charge is conserved within cells.
Maxwell’s equations are consistent for electric and magnetic fields
that vary in time. Moreover, the electric and magnetic fields separately
satisfy wave equations in free space whenever the charge density is zero
 0:

1  2E
  j jE  0
c2 t 2
. (1.16)
1  2B
  j jB  0
c2 t 2

This is an astounding result since it says that a wave will propagate


through space with the speed set by the constant c in the theory. This
was verified by the generation of radio waves. The waves associated with
free space are called electromagnetic radiation and include not only
radio waves but light, microwave and x-ray radiation.
The power of writing a self-consistent and complete set of equations
is demonstrated by the prediction of radiation. The presence of radiation
fundamentally changes the conceptual nature of the law of attraction
between two charges (Coulomb’s Law). This is most dramatically
demonstrated by looking at the electric field that is generated from the
motion of a moving charge q with position specified by a unit vector e r
a distance r away (Feynman, Leighton, & Sands, 1963, p. I.28.2):

q  er   e    2e  
E  2  r  r2    2r   . (1.17)
4  r t  r  x  t  x 

The first term is what one would expect if the force were instantaneous.
In Maxwell’s theory, however, the force is not instantaneous. The
effect of charge propagates with the velocity c and so is not felt at a
given point until it has propagated there. Thus the force one gets is based
Introduction 31

on the “retarded” distance of where the charged body was. The additional
correction terms to Coulomb’s law are based on doing that arithmetic
correctly. The correction terms are not measurable until the speeds are
comparable to the constant c .
Moreover, the equations make connection with a seemingly unrelated
phenomenon: light. We have learned a good deal about light since
Newton. We know light travels as a wave and we know further that
under certain circumstances it behaves like a particle moving in straight
lines. The above analysis construes light as a special case of
electromagnetic radiation. The constant that appears in the equations is
the speed of light (in a vacuum) and is approximately 186,000 miles/sec.
For such speeds, it is not surprising that correction terms were not
initially noticed. A speed of 50 miles/sec, which seems quite fast being
180,000 miles/hour is only 0.00027 the velocity of light. The correction
terms above are actually of the order of this amount squared, since the
coefficient of the corrections to first order is zero.
Maxwell’s equations have been written in a greatly simplified
notation since Maxwell. The electric and magnetic fields are now
understood to be part of a single tensor structure in space-time. We note
the simplifications that occur when the electric and magnetic fields are
written as a single antisymmetric electromagnetic field Fab where the
indices a and b can take on the three spatial coordinates and one time
coordinate (represented as x0  ct ):

Fx 0  Ex Fy 0  E y Fz 0  Ez
Fxy  Bz Fyz  Bx Fzx  By . (1.18)
Fba   Fab

Given this notation, Maxwell’s equations are:


32 Geometry, Language and Strategy—Vol. 2

 x F0 x   y F0 y   z F0 z  
 x Fyz   y Fzx   z Fxy  0
 x Fy 0   y F0 x   0 Fxy  0
 y Fz 0   z F0 y   0 Fyz  0
 z Fx 0   x F0 z   0 Fzx  0
. (1.19)
1
 0 Fz 0   x Fzx   y Fzy   Vz
c
1
 0 Fx 0   y Fxy   z Fxz   Vx
c
1
 0 Fy 0   z Fyz   x Fyx   V y
c

There are two classes of equations.


The first equation (Coulomb’s Law) and the last three equations
(Ampère’s Law as modified by Maxwell) describe how charged matter
provides the source for the electromagnetic field. The second equation
(no magnetic monopoles) and the following three (Faraday’s laws) have
the generic form:

 a Fbc   b Fca   c Fab  0 . (1.20)

An analysis of this set of equation implies that the electromagnetic field


Fab is always determined in terms of a vector valued function Aa called
the vector potential (in analogy to the potential energy):

Fab   a Ab   b Aa . (1.21)

In order to write the Maxwell’s equations more compactly (and thus


make connection with the payoffs from the previous section), we have to
take into account the fact that space coordinates have an extra minus sign
compared to the time component x0 .
Maxwell’s equations demonstrate transformation properties that are
not the same as Newtonian mechanics. Ampère’s Law, consisting of the
last three equations of Eq. (1.19), Coulomb’s Law, the first equation of
Eq. (1.19), the conservation of charge Eq. (1.15) and the wave Eq. (1.16)
Introduction 33

reflect simple geometric forms when expressed in terms of the vector


coordinate displacements

dx 0
 cdt , dx1  dx , dx 2  dy , dx 3  dz

and a symmetric “metric” tensor g ab  gba that has the values:

g 00   g11   g 22   g33  1
. (1.22)
a  b  g ab  0

We use g ab as the inverse of the tensor (considered as a matrix), which


in this case will have the same numerical values. We then distinguish
different tensors depending on whether their indices are up or down by
whether we have multiplied by this metric: Fa b  g bc Fac .
In Maxwell’s theory, light travels with a constant velocity c . This
means that the distance light travels in an interval of time is dx 0 and so
the following invariant length is zero:

ds 2  dx a g ab dxb   dx 0    dx1    dx 2    dx3  .


2 2 2 2
(1.23)

A general path in space and time whose length is given by an expression


of this type describes non-Euclidean geometries (the geometry of curved
spaces) that have been exhaustively studied by mathematicians, notable
Riemann and used by Einstein as the basis for the theory of relativity.
Using this notation we write the four equations that involve a source
(Ampère’s Law and Coulomb’s Law) in a form that shows that all four
equations are of the same form, the laws are covariant when expressed in
terms of the four dimensions of space and time:

1 1
c Fab g bc   gabV b  b F ab  V a . (1.24)
c c
We introduced for the time component of the flow, V 0  1 .
To complete the comparison of the electromagnetic fields to payoffs
Eq. (1.13) we need the behavior of charged matter in an electromagnetic
field, the Lorentz force:
34 Geometry, Language and Strategy—Vol. 2

dVx
m  eEx  e Vy Bz  Vz By 
d
dVy
m  eE y  e Vz Bx  Vx Bz  . (1.25)
d
dV
m z  eEz  e Vx By  Vy Bx 
d

We put this into the covariant form, which also adds a time component:

dV a
m  eF abV b . (1.26)
d
Based on the comparison we see that the form of the payoff (Eq. (1.13))
is the same as this (Lorentz force) form for the behavior of charged
matter in an electromagnetic field.
The comparison suggests we identify the hedge strategy with time,
the flow with the mixed strategy flow and the payoff matrix with the
electromagnetic field. There are differences however that we must still
deal with. The dimensionality above is three space-dimensions and one
time-dimension: the payoff matrix has a dimension n  1 based on the
number n of pure strategies. Distances and energy are appropriate to
decisions not physical distances. Both are a consequence of a differential
geometry. This identification is helpful and we deal with these issues in
the next chapter.

1.9 Outcomes

Our goal was to present a comprehensive and robust framework in which


to discuss decisions. By their nature decisions involve complex strategies
and agents, the latter we will identify with the game theory concept of
player. These players make choices that in fact may change over time
and space. We have been led to a framework in which the strategic
choices are active dimensions of space. We shall argue in the next
chapter that the players provide one or more inactive strategies each of
which is hidden but manifests as a payoff field associated with that
player. The geodesics of this geometry provide a dynamic behavior that
Introduction 35

replaces Eq. (1.13), which in game theory represented only a


mathematical trick to compute the equilibrium strategies.
We summarize the chapter from the standpoint of using it in a course.
By the end of the chapter, the student should be able to do the following:
 Identify real-world decision processes
 Put decision processes into normal form
 See a connection between Maxwell’s equations and Payoffs
The attainment of these outcomes will be facilitated by doing the
exercises in the next section. As a help, the detailed outcomes expected
of the student are listed below.
 From Sec. 1.4, as part of the foundational aspects of game theory, the
student should be able to identify the intensive form of a decision
process and isolate the strategic choices each player has available.
 From Sec. 1.5, the student should be able to understand the essential
concepts of Newtonian mechanics.
 From Sec. 1.6 the student should be able to appreciate the
reformulation of Lagrange that allows Newtonian mechanics to be
effectively applied to the solution of practical problems that involve
constraints. Such problems are related to the types of problems we
encounter in making decisions: many variables with many constraints
in which the essential aspects of the problem are captured by a subset
of the variables.
 From Sec. 1.7 the student should gain an appreciation of how it is
possible that decision processes as described by game theory might be
computable as a deterministic process.
 From Sec. 1.8 the student sees the mathematical connection between
payoff fields in game theory and the electromagnetic fields in physics,
extended to more dimensions.

1.10 Exercises

(1) A childhood game is TIC-TAC-TOE. This game has two players


who make an X or an O on a piece of paper on which there is a
3  3 grid. The players take turns and can make only a single mark.
The extensive form of a game is simply the moves each player
36 Geometry, Language and Strategy—Vol. 2

makes. For example player 1 picks a mark, which he must stick to in


all subsequent moves, puts his mark, say an X, in the 1-1 position.
Player 2 now must use the opposite mark, say an O, and picks a spot
say in the 1-2 position. Player 1 puts an X in the 1-3 position. The
game continues in this fashion. The game ends when either all the
spots are filled or when either player has his or her marks in a line
consisting of a row, a column or along a diagonal. If all the spots are
filled without three marks in a line, the game is tied and is called a
cat’s game. Write out each possible scenario for the two players.
(2) In Ex. 1 the extensive form of TIC-TAC-TOE leads to a very large
number of possibilities. In looking at the decision tree identify the
paths that lead to a win and those that don’t. How many paths lead
to a win? We call each decision path for a player the pure strategy
and the enumeration of the pure strategies of each player the
intensive form of the game. Is there a strategy (strategies) that will
ensure a win or cat’s game every time?
(3) A company has two product lines: it makes razors and it makes
blades. Each product line employs a vast number of people and
requires substantial resources. Discuss the extent to which the
choice between making razors, blades or both is a decision process
in extensive form.
(4) From Eq. (1.1) show the three Newtonian laws: a body at rest
remains at rest, a body in motion remains in motion and for every
action there is an equal and opposite reaction.
(5) Use Eq. (1.1) and the fundamental inverse square law of gravity to
show that the motion of a planet around the Sun is an ellipse.
(6) Using the principle of least action, determine how long it takes for a
body of mass 150 lb. to reach the ground starting from a height 10
yards assuming the body falls freely (ignore air resistance).
(7) Using the principle of least action, determine how long it takes for a
body of mass 150 lb. to reach the ground starting from a height 10
yards assuming the body is constrained to ride a track without
resistance whose equation of height h  f  d  is known as a
function of horizontal distance d .
Chapter 2

Dynamics of Decision Processes

Our choice to avoid a probabilistic or Bayesian approach has a cost. We


must separate out the constraints from the active variables but must still
describe the constraint contributions to the energy (Sec. 1.6). Thus we
need to know more about the dynamics than a purely probabilistic
approach would require. We shall see that there are two sources of such
dynamics. One source that we deal with in this chapter is from hidden
dimensions. The other we deal with in Chap. 3 is from inertial behaviors.
These form the geometry of change. There is an appropriate geometric
space-time with an appropriate set of dimensions, metric and units of
measure. In this chapter we provide the language and set of distinctions
for this discussion.
Our goal is to show that hidden dimensions are required and are
already indicated by the existence of a payoff field in game theory. A
discussion of hidden dimensions is significantly simplified by using the
right mathematical language. That language is differential geometry. It
provides what is for us a very convenient way to approach a discussion
of strategic changes over time that may occur in a variety of dimensions,
whose exact number depends on the number of strategies involved for a
particular application.
A side benefit of such a general mathematical approach is that it
suggests new distinctions about familiar phenomena. In this chapter we
introduce some of the modern terminology and distinctions associated
with differential geometry. This language may not be familiar to those
whose expertise is decision making, to those familiar with economic
thoughts or those familiar with game theory. It may also be unfamiliar to

37
38 Geometry, Language and Strategy—Vol. 2

those with a training in more classical physics or engineering. For this


reason we go over the ideas in some detail.
As a possible motivation for readers to spend the effort, consider
some of the ideas in the media that we might like to understand better.
Based on our work here, we distinguish between decisions made based
on a sense of entitlement from those based on a sense of engagement
with others. In later chapters, we will show that they follow different
differential geometry mechanisms and both contribute to all decision
processes. We might equate entitlement to decisions made by our gut and
engagement with decisions made by our reflective sense. To understand
such distinctions we need to delve more deeply into the theory. We also
consider the distinctions around a code of conduct as distinctions that
make a strategic choice different from a personal view.
To meet the challenge of using the language of differential geometry
for economic processes, we follow what we believe is a sound
pedagogical approach in our selection of ideas from differential
geometry, electromagnetic theory and gauge theory including the theory
of gravitation. The ideas are just those that are necessary for application
to our area of study. We have also selected the key ideas of game theory
and value theory. We believe it is possible for a student who masters
these ideas to have sufficient knowledge to understand and apply the
resultant theory to practical problems. Indeed, we hope that students will
have sufficient understanding to extend these ideas.
We start with an analysis of units of measurement (Sec. 2.1),
followed by a definition of geometric forms (Sec. 2.2) starting with 0-
forms. In Secs. 2.3-2.5 we examine Maxwell’s equations in arbitrary
space dimensions, with the eye to generalizing these equations to
decision process theory. In Sec. 2.6 we examine what it means to have a
space that is locally flat and the implications on change and uncertainty.
Holonomic coordinate bases are introduced in Sec. 2.7 and extended to
general frames in Sec. 2.8. A theory that describes change requires the
concept of rates, which must be defined in a general space; this is dealt
with in Sec. 2.9. For reference, Sec. 2.10 summarizes general geometric
structures such as scalars, vectors and tensors. We end with the creation
of the kinematic component of decision process theory using hidden
dimensions, Sec. 2.11.
Dynamics of Decision Processes 39

2.1 Time, Effort and Utility

Before embarking on a more complete justification of decision process


theory (Sec. 2.6), we anticipate some of the results and address the issues
raised in the last chapter. We suggest units for a theory of decisions and
provide a generalization of electromagnetic theory to n space
dimensions and 1 time dimension.
We will address the question of units first. For decision processes, we
believe the appropriate units are time, effort and utility. They correspond
to time, distance and energy in the physical theory. One objection to our
approach might be that we live in a 3+1 dimensional world. How can it
be that there are more dimensions? We answer that we are describing the
behavior of multiple agents, each of which occupies space at any
moment of time. Because of this many-body nature, we must allow for
many dimensions. In a decision process they manifest as different
strategic possibilities. We must have many dimensions. Recall the car
example (Sec. 1.6) and the discussion of constraints. There in fact might
be many more, but we are pulling out just those that are unconstrained
and are of interest.
We use units of time T  that are big enough to sample enough
decisions for an ongoing set of decisions to provide an accurate measure.
This is not unlike picking units of distance in physics that allow for
repeated measurements so that one can get an accurate measurement of
distance. The idea is that the intervals can’t be too big because there is no
accuracy. They also can’t be too small because the decision process may
not be complete. These concepts are closely related to the principle of
least action (Sec. 3.6) and the concept of a minimal or quantum of action
(Sec. 3.7). In these sections we will sharpen the domain of validity of the
principle of least action and the concept of time being “big enough”.
Making a decision involves the act of making a choice and making
the relevant effort in (distance) units  L  to execute that choice. In a
business, we might measure the effort in staff-years. We could also
measure the effort in staff-seconds if the decision process is sufficiently
simple. The effort made per unit time is then the speed, which is given by
the units of effort and time,  L T  . For much of what follows, we pick
units so that the maximum possible speed c is unity: in other words it is
40 Geometry, Language and Strategy—Vol. 2

the maximum or ideal effort a member of the staff of the decision


process might make. All other speeds will by definition be less than this
ideal.
Distinct from the effort we make in choosing, there is the value or
utility of that choice. We discuss utility theory in more detail in Sec.
12.2. For now, it is sufficient to note that we measure utility as a value
U  that is a convertible value such as dollars, pounds, Euros or yen. We
are aware of the arguments in the literature that the value is only relative
to the decision agent. We support this view, which provides input on the
choice of geometry. Based on the geometry, in Sec. 13.5, we make the
case that all utilities are convertible, though the mechanism may be
complex. That being the case, we choose units for payoffs that are related
to utility.
We assume that decision process theory geometry will yield
something like Eq. (1.21), in which case we assume that there is a payoff
potential Aa , which we take to be unit-less. The corresponding payoff
field Fab has units of inverse effort,  L1  . To convert effort to utility,
we require a productivity factor 1  . We measure payoffs 1 Fab as the
utility per unit strategic surface, U Ln 1  where the strategic surface is
a cross-section of the strategic volume of the n strategic choices of the
decision process. Consequently, the productivity factor 1  has units
U Ln2  , which is utility per unit strategic boundary and will be
consistent with our usage in Eq. (3.31) and subsequent usages such as
Eq. (4.14).

2.2 Geometric Forms

We restate the basic equations from Sec. 1.8 for the electromagnetic
fields in the language of relativity and geometry by going from the
differential formulation to an integral formulation, [Warner, 1971]. We
use the language of forms from differential geometry to represent the
integral quantities of interest.
We start with a 0-form, which is a scalar quantity  . It has the
property that its differential, integrated along any closed path is zero:
Dynamics of Decision Processes 41

 d  0 . (2.1)

The differential of a scalar depends on the gradient components1:

d   a dx a . (2.2)

It makes sense to think of the differentials dx a as a basis (in the sense of


a linear vector space) and the components along this basis  a  as
defining a new quantity called a 1-form.
More generally, any set of components Aa define a 1-form, though in
general the integral of this 1-form around a closed path is not zero:

 A dx  0.
a
a (2.3)

The 1-form A represents a differentially short path segment of a curve in


space:

A  Aa dx a . (2.4)

Line integrals are constructed from the integration of such segments


along a curve. They represent the circulation of Aa along this path.
The differential dA of a 1-form defines a 2-form and represents the
movement of the entire segment, which spans an area. In three-
dimensional vector calculus, a differential area is conveniently written in
terms of the cross product of two differential vectors, one along the
segment and the other along the transport path. In higher dimensions the
cross product is generalized to the antisymmetric wedge product of two
1-forms, a 2-form, defined in terms of the measure for area:

A  B  Aa dx a  Bbdx b  1 2  Aa Bb  Ba Ab  dx a  dx b . (2.5)

In general, the integral of a 2-form around a closed surface is not zero;


however, in the special case that the 2-form is the differential of a 1-
form, the closed surface integral is zero. In the general case, we think of
the 2-form as the flow through the surface.

1
We use the summation convention: Cf. Eq. (2.22).
42 Geometry, Language and Strategy—Vol. 2

We use n-forms to re-write the basic electromagnetic field equations


in integral form in any dimension of space time. We propose to identify
the payoff field with the electromagnetic field Fab , using the units set
out in Sec. 2.1. The key properties of the electromagnetic field equations
depend on whether certain closed integrals are zero or non-zero. This is
the concept that will be generalized.

2.3 Source-Free Field Equations

To begin, we define the field 2-form F in terms of the field intensity


Fab and the element of surface dx a  dx b formed as the wedge product
of the corresponding components. The field intensity represents a flow
through this surface element: this is the flux (field intensity per unit area)
that flows through this infinitesimal 2-surface:
1
F Fab dxa  dxb . (2.6)
2!
These forms are ideal mathematical constructs to describe the physical
notion of the flow of flux. The units of the field 2-form is effort  L  (e.g.
staff-years).
The integral or sum of this 2-form over a closed surface is then given
by Stokes’ Law for the differential form of the flux integrated over the
enclosed 3-volume
1
dF   c Fab dx c  dx a  dxb :
2!

 F   dF  0 .
K K
(2.7)

We make use of two general properties of n-forms. First, Stokes’ Law for
any n-form G expresses the fact that the integral of the differential of
G over any subspace K of an n-dimensional space is equal to the
integral of G around the closed boundary K of that subspace:

K
 G   dG
K
. (2.8)
Dynamics of Decision Processes 43

Gauss’ law is an example of such a law, stating the integral of the


electric field over a closed surface is the gradient of that electric field
integrated over that volume.
Second, the differential of any n-form that is constructed from the
differential of an n  1 -form is zero:
dd  0 . (2.9)

The electromagnetic flux is the differential of the electromagnetic


potential, F  dA . This carries over into any number of dimensions and
so works for the payoff field in particular.
Considering time and space components separately, there are two
types of equations that result from Eq. (2.7). The first occurs if the
bounding 2-surfaces contain only spatial components.

 F
K
mn dx m  dx n  0 . (2.10)

This is the analog of no magnetic monopoles

 BdS  0 .
K

The net flux through the closed boundary is zero. The implication is that
the magnetic flux, defined as the spatial components Fmn of the field
intensity, cannot stop or start inside a volume, but flows continuously.
The differential form of this is

 l Fmn   m Fnl   n Flm  0 .

The general form involves both space and time components:

a Fbc b Fca c Fab  0 . (2.11)

In particular, the flux doesn’t get created or destroyed inside a surface in


which one of the dimensions is time. We may think of the time
components in a different way; express the differential form in terms of
the electric fields Em  Fm0 and the magnetic field components Fmn :
44 Geometry, Language and Strategy—Vol. 2

0 Fmn  m Fn0  n F0m  0  m En  n Em  0 Fmn . (2.12)

This is the generalization of Faraday’s Law, Eq. (1.20), showing that an


electromotive force (left-hand side) is generated whenever there are
magnetic flux lines that change in time.
The electromotive force is determined by the one-dimensional
integral around a closed curve of the magnetic flux that goes through the
surface bounded by this curve (Faraday’s law):

 1
 F
K
m0 dx m  
t K 2!
Fmn dx m  dx n . (2.13)

The left-hand side is the circulation of the electric flux around a one-
dimensional boundary, and is defined as the electric potential.
This equation is a special case of Eq. (2.7) where one of the
dimensions is time and the initial and final times are very close. In this
case the two-dimensional surface on the left originates from t times a
one-dimensional integral of the boundary, and the three-dimensional
volume on the right originates from the difference of two-dimensional
volumes over a short interval of time. The above result is obtained by
taking the time difference and dividing by the differential time. We think
of the result as the time rate of change of the spatial components of the
payoff flux flowing through a two-dimensional surface determining the
electric potential.
In the general case of a payoff field, the magnetic flux (calling the
space components of the field intensity Fmn the magnetic field) flows
like a continuous media without sources or sinks (lines are not created or
destroyed). Changes in the magnetic flux with time generates the
electromotive force (unit-less) described by the left-hand side of Eq.
(2.13). The magnetic flux lines cause charged particles to move in
patterns identical to the paths predicted by elementary game theory,
identifying the magnetic field components with the payoff matrices of
the players.
In the steady state case, no electromotive forces are generated,
analogous to the situation of steady state electric circuits: The above
equation reduces to Kirchhoff’s Law of electrostatics. The result of
Dynamics of Decision Processes 45

changing flux with time creates electrical forces that induce currents
(motion of charges). Inertial energy is converted to electromagnetic
energy and vice versa.

2.4 Sourced Field Equations

The remaining set of equations of generalized Maxwell involves the


sources of the electromagnetic fields. The sources are current 1-forms j
constructed from the circulation of charges (utility flow ja ) along a path
in space-time. For any flow at each point, there is a dual description: the
flow along a path can be described by the flow through the orthogonal or
dual surface. The form of the current flowing through the dual surface is
in terms of the element of volume, an n-form:

1
*j   ab f j a dx b    dx f . (2.14)
n!

The dimension of space time is n  1 and we write the current in terms of


a totally antisymmetric tensor  ab f with n  1 -indices and a vector j a
representing the current density flowing through the volume (the dual
surface). Typically the current density is related to the charge density
 ch and velocity flow V a by j a   chV a . We measure current density
in units of U Ln  .
This process of constructing a new tensor using the totally
antisymmetric tensor is quite general and the resulting tensor is called the
dual tensor and the operation of creating the dual is indicated by an
asterisk. Corresponding to the 2-form payoff flux flowing through a 2-
surface, there is a corresponding (n – 1)-form payoff flux *F describing
the flow along the dual space orthogonal to the 2-surface. The remaining
Maxwell’s equations are expressed in terms of this dual 2-form:

1
*F   adef F ef dx a    dx d . (2.15)
2  n  1!

The scaled payoff *F  has dimensions of utility and represents the


flow of utility flow along an (n – 1)-dimensional strategic surface.
46 Geometry, Language and Strategy—Vol. 2

Utility can be created and destroyed since the current acts as the source
or sink.
This leads to generalizations of Coulomb and Ampère’s law that
relate the differential of the scaled payoff to the current and are
expressed here as a single relation:

*j  d * F  . (2.16)

Through Stokes’ theorem Eq. (2.8), the integral form is:

 *F    d * F    *j .
K K K
(2.17)

The right-hand side describes the utility. The left-hand side describes the
effort to produce that payoff in terms of the dual payoff *F . The inverse
of the productivity factor  is a measure of the effort required to
produce one unit of utility.
Utility is created or destroyed through the flow of currents. The units
of the current are U Ln  . As before, these equations represent different
physical situations depending on whether the components of the
integrand are purely spatial or not. If we suppose that on the right, the
volume is purely spatial, then we find that the integral is the charge over
the surrounding volume,

1
 n! 
K
0mm j 0dx m    dx m .

In this case, the left-hand side represents surfaces which are purely
space like, so one and only one of the components can be timelike. The
left-hand side is therefore the net electric field density that is normal
to the (n – 1)-dimensional strategic surface bounding the n dimensional
strategic volume,

 F Sm   Qencl .
0m

K

This is the integral form of Coulomb’s Law


Dynamics of Decision Processes 47

 m F 0 m   ch ,

Eq. (1.19), where the surface integral

1
Sm   0 pqm dx p    dx q
 n  1!
is the dual of a 1-form in the n-dimensional strategic subspace.
The decomposition of a payoff field into its spatial and time
components is frame dependent. In each frame, the space coordinates
determine a volume; the spatial payoff component (flux) that flows
through any two-dimensional strategic surface of that volume is
magnetic. The space-time flux that flows through an (n – 1)-dimensional
strategic surface is electric. The net number of flux lines that emanate
from this spatial volume equals the enclosed utility. This nomenclature
however is unnecessarily tied to the physics analogy, though at times it is
comforting and provides a useful mnemonic for looking up standard
results. We introduce new terminology to emphasize we are not tied to
that analogy.

2.5 Dynamic Payoff Equations

We have said that the total utility (“charge”) is determined by the net
flow of the electric components of the payoff field through the strategic
surface. This supports the idea that utility is the source of value as used
in game theory: utility determines the electric field and value as used in
game theory becomes the electric field in decision process theory. We
therefore suggest the electric decomposition of the payoff fields be
valuation fields. Further, we suggest negotiation fields be the magnetic
fields. We suggest this since there are always two distinct negotiation
strategies for which the payoff represents the payout. Each pair of
strategies defines a negotiation plane. Each payout represents a flow of
negotiation flux lines through this negotiation plane. In general, since
the negotiation field is an antisymmetric matrix, at each point we can
always find a frame of reference in which the matrix is block diagonal
reflecting a disjoint union of negotiation planes. If we think of the
48 Geometry, Language and Strategy—Vol. 2

negotiation flux flowing along the negotiation plane normal, we have a


generalization of magnetic field in any dimension.
As part of our definition of equilibrium solutions, we treat two
dependent concepts from game theory as independent. Though
independent, these two concepts are tied together dynamically. First, the
time dependence of the negotiation flux through a two-dimensional area
is the source of the valuation field, Eq. (2.13). Second, we show below
that the connection works the other way: the time dependence of the
valuation flux through the dual (n – 1)-dimensional surface is the source
of the negotiation field, Eq. (2.19).
This set of equations is determined from Eq. (2.17) if one of the
coordinates of the “volume” is along the time direction. As before, it is
helpful to start with the differential form of the equations:

 k F jk   0 F j 0   j j
. (2.18)
 k F jk   j j   0 F 0 j

We expect two terms from Eq. (2.17).


The first is an (n – 2)-dimensional surface times the interval t along
which we integrate the magnetic components F mm  . The second is an
(n – 1)-dimensional surface times the incremental difference (in time) of
the electric field along the normal to the surface. This is the analog of
Ampere’s law:

1 F mm  F 0m
 2!    
t K 
m
S mm j S m Sm . (2.19)
K K

The right-hand side we call the enclosed utility current, including the
contribution of the displacement current that is induced from valuation
fields that change in time, flowing through the strategic surface.
The boundary can be thought of as the dual of the 2-form created
from the wedge product of the surface 1-form and a second intersecting
surface 1-form that defines the boundary. The dual of this 2-form is

1
S mm   0 pqmmdx p    dx q ,
 n  2 !
Dynamics of Decision Processes 49

which is an (n – 2)-form and in general is not a 1-form (except in three


dimensions). It does generalize a line integral: the 2-form is the wedge
product of two 1-forms and thus represents a subspace normal to the two
1-forms. Its dual is thus an integral element in that subspace.
The left-hand side is thus the circulation of the scaled negotiation
field F mm  along the n  2 strategic boundary Smm that surrounds the
strategic surface. What passes through the 2-surface circulates in the
dual (n – 2)-surface. The negotiation field is between the direction m
that is normal to the surface and the direction m  that is normal to the
intersecting surface defining the boundary. This law states that the utility
currents that flow through the surface generate, or are the source for, the
negotiation fields that flow along the boundary. Each term has units of
U T  .
We thus have a complete set of generalized Maxwell’s equations in
(n + 1)-dimensional space-time. It is relevant to add that these Maxwell’s
equations require the current to be conserved,
d *j 0,

which follows from Eqs. (2.16) and (2.9). It is equivalent to

a ja  0 ,

in differential form and to the more intuitive integral form:

Q
 *j   d * j  0   j Sm  
m
. (2.20)
K  K K
t

This states that the rate at which the enclosed charge (utility) changes in
time is equal to the net current leaving the n-dimensional strategic
(space) volume K through the enclosing n  1 strategic surface K .
Charge is not created or destroyed, but flows like a fluid through space.
We have anticipated our adoption of this generalization of the
electromagnetic field as the payoff field. We have provided the units that
make sense for this generalization. Implicit in our approach is the notion
that decision strategies change continuously over time. In the next
section we introduce the basis for this anticipation.
50 Geometry, Language and Strategy—Vol. 2

2.6 Continuous Change versus Uncertainty

In many approaches to decisions (Cf. Sec. 12.1), the key aspect is the
uncertainty of the decision and its independence from past events. The
approach is to use Bayesian or inverse probability to estimate future
behaviors. It is a stochastic approach. Yet in modern businesses, the
opposite holds: businesses practice continuous process improvement as
the basis for improving profits and quality of their products.
Incrementally improving past behaviors is the key to future successes.
We believe as well that continuous change is a necessary attribute of
decisions and can be applied to mixtures of strategies viewed as
frequencies (Sec. 1.4). Decision processes can be put into a normal form,
albeit locally. For sufficiently short periods of time and sufficiently small
strategic variations, the foundational aspects of game theory provide a
framework for decision process theory. This is a statement of “flatness”.
To be locally flat however is not to say that that is true globally.
To discuss strategic change, we distinguish between changes in
choices that occur at a constant rate and those that don’t. The latter
changes are characterized by acceleration, which is the rate at which
strategic rate changes occur. In many cases we are unaware of such
accelerations. We will often refer to the physical analogy of our
perception that as we sit comfortably at home at rest, we are unaware that
we are spinning around the North-South axis once every 24 hours and
rocketing around the sun every 365 ¼ days. Yet those of us that pay
attention to weather are aware of the induced acceleration effects
attributed to Coriolis, because they contribute to the formation of
cyclonic weather conditions [Crawford, 1978]. A discussion of dynamic
changes involves not only the ability to order our preferences, but the
ability to measure the magnitude of change of those preferences and
whether the changes occur at a constant or accelerating rate. As with
weather, we must learn to identify when such acceleration effects are
present. No acceleration is yet another type of “flatness”.
To address these issues we apply what we feel are common sense
notions of space (applied to strategic choices) and time to the domain of
decision making. We assume that choices that are near to each other
strategically or near to each other in time, lead to actions that are
Dynamics of Decision Processes 51

correspondingly near. This requires a measure or metric between


decision events: we need the ability to determine what it means for
choices to be near or far apart. We need a notion of continuity. We must
be able to assign the measure to the rate of changes and thereby
distinguish constant from acceleration effects. We then have a basis for a
dynamic theory.
Be warned however that in general, common sense notions are not
automatically common to all or even mutually consistent. A famous and
relevant example is revealed by Einstein’s analysis of the Newtonian
notions of space and time. Somewhat simplified, Newton’s view was that
gravity effects propagate instantaneously to all parts of the universe;
Einstein, however, maintained that gravity propagates its effects as a
signal that goes at the speed of light. This is consistent with Maxwell’s
view for light. Because in a vacuum, the speed of light is very large, the
two views are ordinarily indistinguishable. In today’s world, however,
we produce devices and look at phenomena where the effects are easily
distinguishable and we must adopt Einstein’s view that has yet to be
common. This highlights the value of supplementing our common sense
with the requirement of a self-consistent framework within which we can
formulate our ideas and models.
Einstein’s view is relevant to us because it provides a self-consistent
and common sense view of acceleration. With this view, in Sec. 2.9, we
show that two new notions, torsion free and measure independence,
lead to a determination of the acceleration effects. Acceleration is
proportional to the forces that determine the dynamics and is thus an
important attribute of dynamics in any dynamic theory.
Our starting point for decision process theory is that decisions are
events that generate action and they occur at a moment in time and a
point in space. The foundation of our framework is both topological and
geometric. We need the concept of order to indicate preferences and
measure to determine nearness. We articulate our ideas using relevant
information on topology and differential geometry taken from the
literature as noted below. We require physical space time to be locally
isomorphic to flat space, which corresponds to our experience, that
curvatures are not ordinarily noticed. This physical assumption
presupposes that at each point there is an inertial frame of reference, i.e.
52 Geometry, Language and Strategy—Vol. 2

a frame in which there is no perceived acceleration. This generalizes the


common perception that locally the world around us is flat and not
moving. This assumes there is an independent set of strategic directions
that represents the set of possible choices.
Globally, time and space need not be flat. To explore this we start
with the observation that the dimensions of time and space are
fundamentally connected by Einstein’s notion of cause and effect. A
cause generates a signal that propagates outwardly from that event and
this signal influences future effects. It also generates a network
relationship to neighboring strategic events. Though we consider a space
of applied dimensions, we maintain this property that signals propagate
with a finite velocity. We choose units (Sec. 2.1) so that the fundamental
speed of propagation is unity; distances are measured in units of time.
This is analogous to astronomers measuring distances in light-years, the
distance light travels in one year. These signals are physically as real as
the events that produce them. As such they carry energy and have
momentum. They are extended structures in space and time and exhibit
stresses analogous to physical media such as fluids.
Since we assume that space is locally flat, at every point there is an
inertial frame of reference in which a signal generated will propagate
outward as a spherical wave. In this frame of reference after any small
interval of time dt , the wave resides on a sphere around that point with
strategic coordinates dx a , where a runs through the possible strategic
coordinate directions. Since the wave is a spherical wave, the relation
between time and space on the sphere is:

ds 2  dt 2   dx a dx a  0 . (2.21)
a

We will have occasion throughout this book to indicate sums as we have


indicated above that include using the mathematical “sigma” sign for
summation and will involve time components. We find this notation
cumbersome.
With a little practice the following shorthand for quadratic forms is
easier to read and understand:

ds 2  g ab dx a dx b . (2.22)
Dynamics of Decision Processes 53

There are three things to keep in mind: first, if there are repeated indices
with one upper and the other lower, add a summation “sigma” for that
index. Second, allow the indices a, b to represent not only the
allowed strategies but also time. Third, specify all the tensor quantities
so that the desired expression is replicated.
In this case we replicate Eq. (2.21) if the non-zero components are
gtt  1 and the strategy components are diagonal and gaa  1 . The
geometry of this space is not Euclidean but the curved Minkowski space.
For our purposes in this book we call spaces with a Minkowski metric
flat.

2.7 Holonomic Coordinate Bases

To achieve our end goal of formulating decision process theory, we need


to understand the mechanics of describing a system from different
perspectives or frames. In differential geometry, these frames are
distinguished by different types of coordinate systems. In this and the
next section we consider such coordinate systems, starting with those
that are called holonomic.
Consider a point in the strategic space that corresponds to making a
strategic choice at some instance of time. It is a point in an  n  1 -
dimensional space. We write the space-time point x a as the  n  1 -tuple

x a  x 0 x1  x n  ,

where the time component is x 0 . The components x a are scalar


functions and provide a holonomic coordinate basis to describe the
space-time point. By holonomic, we mean there are scalar functions x a
associated with each coordinate. Surfaces on which each of these scalar
functions is constant provide a coordinate basis with which to describe
positions at each point in time.
The ordinary rules of calculus can be used where the order of partial
derivatives yields the same result:

 a  b   b a .
54 Geometry, Language and Strategy—Vol. 2

The consequence is that changes in scalar quantities between two


neighboring points can be computed by traversing along first one
coordinate surface a differential distance and then along the second. It
does not matter which surface is chosen as first. Not every basis is a
coordinate basis. We will see shortly that for more general bases, the
order does matter.
The existence of a coordinate basis is an important part of our
theoretical framework. A familiar example occurs when describing the
points on a sphere. One can carry around an actual sphere or map a
sphere onto a piece of paper, a two-dimensional projection or chart. For a
sufficiently small section of the sphere, the chart provides an accurate
flat representation of the sphere where latitude and longitude might be
chosen as the coordinate variables. In the general case, however, the
constant value of the scalar field x a describes a hyper-surface on which
all points have the same coordinate value. In the special case of the
sphere this hyper-surface is a circle on the sphere. The example of the
sphere demonstrates that the property of being locally flat does not mean
the space is globally flat. When the space is not flat, you need more than
one chart to navigate the surface.
A new frame of reference is defined by an independent set of n  1
functions

y a  x 0 , x1 ,, x n 

that depend on the old coordinates. The new frame of reference also
provides a holonomic coordinate basis.
These frame considerations allow us to implement the common sense
notion that the specification of pure strategies may not be unique: many
equivalent choices are possible. We thus avoid falling into the trap that
every player adopts the same set of values, since they may be measuring
values using different coordinate bases. We create a decision process
theory in which any frame of reference is an acceptable choice. We
assume covariance, that the fundamental dynamics will be expressed by
a mathematical form that does not depend on which choice we make.
A natural consequence of our dynamic theory is the idea that the
strategies change continuously over time (due to a principle of least
Dynamics of Decision Processes 55

action). The study of such changes is closely related to how quantities in


decision process theory transform: every continuous change can be
represented as a transformation. Thus dynamics suggests we investigate
the most general frame transformations that might occur.
We start our investigation by considering the spherical propagation
signal. The outgoing signal can be viewed in many other frames. If we
relax the condition that we are in a locally flat frame, then we maintain
the compact form Eq. (2.22) with ds  0 , but the values of the tensor gab
become arbitrary. We say that the “distance” ds is a scalar field. Since
the changes of the tensor gab may vary from one point to the next, we
call such a tensor a tensor field to emphasize the space-time dependency.
We must be able to compute the values of this tensor in an arbitrary
frame. We see that the tensor describes the cause and effect relationships
since it prescribes the behavior of the outgoing signal.
We next describe in more detail the transformation properties
between frames and how they lead to an understanding of the cause-
effect signals and how this relates to the tensor gab .

2.8 General Coordinate Bases

We recall the rules from calculus that differential changes of variables,


which are important for specifying cause and effect in Eq. (2.22), can be
determined from the partial derivatives and the chain rule:

y a  x 0 , x1 ,, x n 
dy a  dx b   b y a dx b . (2.23)
x b

Note our use of the summation convention to simplify the writing and
our introduction of a more compact notation to indicate the partial
derivatives. Though the expression can be a laborious computation, the
essential conceptual point is that the new differentials are linearly related
to the old differentials.
We expand the concept of a frame transformation by allowing any
linear transformation of the differentials, including those that are non-
holonomic. The resultant expression E we call a 1-form using the name
from the theory of differential geometry (Sec. 2.1), which, when these
56 Geometry, Language and Strategy—Vol. 2

infinitesimal quantities are summed over a path, provides the basis for
defining a line integral:

Ea  E a b dx b . (2.24)

This expression is a general coordinate basis and represents the most


general linear frame transformation at a point that might occur in our
dynamic theory of decisions.
We note that for the special case that the 1-form consists of
coordinates that are perfect differentials, we have

c E ab  b E a c ,

which follows from the equality of the mixed derivatives in Eq. (2.23),

 c  b y a   b c y a .

A non-holonomic frame is one in which

c E ab  b E a c

for at least one pair of coordinate values. The usual rules of calculus
don’t hold. Changes to fields now depend on the path taken.
We assumed that at each point there is a frame in which the cause and
effect signal propagates outwardly as a spherical wave. In a general
frame E a we assert that this implies that the propagation is determined
by the vanishing of the quadratic form ds 2  g ab Ea Eb constructed from
the frames. The coefficients are determined by the transformation.
Cause and effect must apply to all physical phenomena. We expect
that regularities of physical phenomena will involve the rates of change
of variables, such as a scalar field x a , a vector field X  , or an order 2
tensor field gab . There will be higher order tensor fields to consider as
well. We need rules that apply to any tensor field and its rates of change.
In the frame in which space is locally flat, rates of change are determined
by ordinary calculus, by the derivatives. Transformation to rotating
frames requires that the effects of the acceleration be included in the
definition of the rates of change. We again recall that the Coriolis force
due to living on a spinning sphere (earth) is an example of a rotating
Dynamics of Decision Processes 57

frame effect. These considerations are thus really common sense


additions that must be part of the decision process theory. This becomes
especially relevant when we establish that payoffs are essentially
rotations.

2.9 Parallel Translations

The basic rules for including rotating frame effects for any tensor field
follow from how we treat the rate of change of a vector field. The rate of
change of a vector field along a path parametrized by a scalar  is the
difference of the field at two different points divided by the differential
change d along the path. In dealing with rotating frames we must be
concerned with Coriolis-type forces: we must be clear about how we
compare vectors at two different points. In physical applications, we
make the comparison by parallel translating the initial vector along the
curve to the comparison point. This removes the effect of the rotating
frame, exposing only the actual change in the vector field.
As a result, the rate of change of the vector field will be a differential
change of its values plus a contribution due to the frame rotation, which
linearly mixes the vector components:

DX a  dX a  ωa b X b . (2.25)

This defines the covariant derivative. The linear mix or orientation


a
potential ω b is a 1-form (remember that the 1-form is determined by
the frame properties ω a b   a bc dx c ).
The differential change is determined by the ordinary partial
derivative 1-form

dX a   b X a dx b .

A useful variant of this form is to define the covariant generalization of


the partial derivative as

DX a  X a ;bdx b .

Using this expression for the covariant derivative we obtain:


58 Geometry, Language and Strategy—Vol. 2

X a ;b   b X a   a cb X c . (2.26)

Rules that are expressed in terms of the covariant derivative will be


covariant in that they will have the same form in any frame of reference.
This is an important and required aspect of any physical theoretical
foundation including ours.
We transform from a holonomic coordinate basis to a general
coordinate basis using Eq. (2.24). In the general coordinate basis, the
covariant derivative has the same form Eq. (2.25). Therefore the
orientation potential 1-form ωa b will now be expressed in terms of the
new coordinates with

ωa b   a bc Ec .

The partial derivative  a transforms to the frame derivative

 a  Ea b  b ,

which depends on the inverse E c b of the frame transformation, Eq.


(2.24):

dX a   b X a Eb . (2.27)

The second term of Eq. (2.25) is zero if globally there is no frame


rotation.
In the general case, the properties of this 1-form are completely
determined by the following two assumptions.
First, for any general coordinate basis components E a , we assume
that the orientation of each frame is unchanged under coordinate
displacement. We express this assumption compactly as saying frames
are torsion free:

DEa  dEa  ωa b  Eb  0 . (2.28)

In a space that is locally flat, all frames will be torsion free. The 1-form
has properties of a vector field so we expect the two terms for the
differential of a vector 1-form. The 1-form represents a differentially
short path segment of a curve in space. The line integrals are constructed
Dynamics of Decision Processes 59

from the integration of such segments along a curve. The differential


represents the coordinate displacement of the entire segment and so
spans an area. In three-dimensional vector calculus, the area is the cross
product of two differential vectors, one along the segment and the other
along the displacement path.
In higher dimensions the area is determined by the antisymmetric
wedge product of two 1-forms, called a 2-form, defined in terms of the
measure for area:

A  B  Aa dx a  Bbdx b  1 2  Aa Bb  Ba Ab  dx a dx b . (2.29)

The study of forms is therefore the study of measures of length, area,


volume, etc. Based on these forms, the first term of Eq. (2.28) is
determined by the structure constants

C a bc  C a cb

defined in terms of the partial derivatives of the frame transformation


components:

dΕ a  1
2  E
b
a
c   c E a b  dx b  dx c
. (2.30)
dΕ a  1
2 C a bc Eb  Ec

The second term of Eq. (2.28) states that coordinate displacement of a


short segment has an additional contribution for a rotating frame. Any
general frame under coordinate displacement is unchanged as expressed
by Eq. (2.28). For rotating frames this provides one relationship between
the linear transformation ωa b 1-forms and the structure constants C a bc .
We thus also call a general coordinate basis a fixed coordinate basis.
Second, we make the key assumption of measure independence: the
covariant derivative of Eq. (2.22) is zero. An outgoing spherical signal
wave must be a frame independent statement. This implies that the
covariant derivative of the tensor gab is zero in any frame. The covariant
derivative of this tensor transforms in a dual manner to the symmetric
product of two vectors. Based on this consideration, the covariant
derivative can be shown to be:
60 Geometry, Language and Strategy—Vol. 2

c c
Dgab  dgab  ω a gcb  ω b gac . (2.31)

We require

Dg ab  0 .

These considerations lead to the important result that provides a


relationship between the ordinary (frame) derivatives of the tensor gab
and the orientation potential 1-form ωα β , which when combined with the
first assumption provides the explicit relationship of orientation potential
components and the structure constant coefficients C a bc in Eq. (2.30):

abc  1 2  Cabc  Cbca  Ccba   1 2   a gbc  b gca   c g ab  . (2.32)

In other words these various fields are not independent.


This simplifies for a coordinate basis:

abc  1 2   a gbc   b gca   c gab  . (2.33)

It simplifies because the differential of the coordinate frame is zero


d  dx a   0 , which implies that the structure coefficients Cabc  0 are
also zero. The orientation potential in this basis, also called the
connection, is symmetric in the last two indices,

acb  abc .

Another special case is the orthonormal coordinate basis in which by


definition, the metric is everywhere equal to the Minkowski metric and
so the frame derivatives are zero:

abc  1 2  Cabc  Cbca  Ccba  . (2.34)

The orientation potential in this basis is antisymmetric in the first two


indices

bac  abc ,

which follows from the antisymmetry of the structure constants.


Dynamics of Decision Processes 61

2.10 Geometric Structures

We have stated that in decision process theory, the rules for decisions
should be covariant. The basis for this comes from utility theory in game
theory, in which preferences are determined only up to linear
transformations (Sec. 12.2). Thus for any given event, we express the
theory in such a way that the rules are independent of linear
transformations. Theories that are invariant under transformations have
characteristic geometric structures, such as scalars, vectors and tensors,
which we will define in more detail below. We start the discussion by
laying out in detail what this covariance implies.
We identify a geometric structure as a set of related functions in the
theory that maintain their form under transformation. The simplest
example is a single function that transforms into itself. We define a field
 to be a scalar field if     , which indicates no change under the
transformation other than expressing the field in terms of new variables.
The next more complicated structure is a collection of functions in 1-
1 correspondence to the number of space time dimensions. The simplest
example is that of functions that define a frame. For each direction b , the
frame transformation components E a b form the components of a
contravariant vector2 in the new basis: the collection of such vectors is
E a . (One can equally well look at each direction a , and the frame
transformation is a set of covariant vectors in the old basis; similar
statements can be made about the inverse frame transformation Ea b .)
We consider a frame transformation λE , which is a product between
the matrix λ , which is a frame transformation in its own right, and the
frame components E that transform a set of frames into a new set
(denoted by the prime):

E a   a b E b . (2.35)

As with the scalar field, the new frame components are also expressed in
terms of the transformed coordinate values.

2
It is defined in the next paragraph.
62 Geometry, Language and Strategy—Vol. 2

We generalize this and define X a to be a vector field if the


components transform to the matrix product X   λX , and are expressed
in the new variables. We will have occasion to distinguish between two
types of vector fields. Following the literature we call this a
contravariant vector. The 1-form ω  a E a has components a that are
dual to contravariant vectors in that they transform according to the
inverse matrix λ 1 . These are components of a covariant vector.
A tensor field is the generic name for geometric structures that
include scalar fields, vector fields and more general fields in which there
is some number of contravariant components and some number of
covariant components that separately transform as the corresponding
contravariant or covariant vectors respectively. We can now be more
precise about our requirement that decision process theory be covariant.
We consider only laws that are expressed in terms of such geometric
structures or tensors.
In any theory that allows frame rotations, it will not be generally true
that geometric structures at one point carry over to the same geometric
structure at some other point. Yet that is the type of theory we require.
The general class of such theories is called a gauge theory in physical
theories, for example [Hawking & Ellis, 1973] and a fiber bundle by
mathematicians, [Steenrod, 1951]. We require the following notions.
We require that the transformation matrix is a continuous function of
position and we refer to it as a gauge transformation. Decision process
theory is a member of the class of gauge theories that have been well-
studied. The orientation potential 1-forms ωa b form a matrix ω of
1-forms. Though these potentials are not components of a tensor, they do
have a simple transformation rule:

ω  C  λ  x   λ  y  ω  C  . (2.36)

In words, a transformation to a new frame followed by a parallel


translation along a curve C is the same as a parallel translation along
the curve C followed by a transformation to the new frame. This
provides the connection between the geometric structure between one
point and any other point. We show this in more detail.
Dynamics of Decision Processes 63

When the parallel translation is differentially small, from y  dy to


dy , then this transformation rule, which also holds for each frame
transformation, is:

ω  y   λ  y  ω  y  λ 1  y   λ  y  dλ 1  y  . (2.37)

The first term is the transformation of a mixed tensor. The second term
indicates that for rotating frames, the orientation potential is not a mixed
tensor.
It is also true that the ordinary derivative of a contravariant vector
field is not a tensor since the transformation of the differential of a
contravariant vector field is:

dX   λdX  dλλ 1 X  . (2.38)

Putting this together with Eq. (2.37), we see that the covariant derivative
Eq. (2.25), formed from the parallel translation of the contravariant
vector field, does transform like a vector field.
The transformation components of Eq. (2.37) that don’t transform as
a vector field cancel the corresponding components of Eq. (2.25) of the
differential of the vector field:

DX   λdX  dλλ 1 X   ω X   λdX  λωX  λDX . (2.39)

In particular this holds for the torsion free coordinate basis vectors:

DE  λDE . (2.40)

Frames are geometric structures defined over the whole space; they
transform as vector fields.
We thus have a proof that local gauge invariance for a space that
is locally flat (Minkowski) implies that every frame is torsion free. This
provides a justification for Eq. (2.28). Because the derivative D that
results from parallel translation preserves the tensor field
transformation, the name covariant derivative is appropriate. It is a
differential operator that preserves the geometric structures of the theory.
Although the orientation potentials are not tensors or geometric
structures that are preserved, they are associated with such structures and
64 Geometry, Language and Strategy—Vol. 2

are physically real fields. Preserved structures are by definition, gauge


invariant. It is proved in the literature that the flux that goes through a
closed line integral formed from the orientation potential 1-form matrix
is equal to a gauge invariant 2-form:

R  dω  ω  ω . (2.41)

This orientation flux field carries energy and momentum through the
2-surface, is a tensor field and provides the physical field that propagates
the cause to the effect. In physical theories it is called the curvature
tensor.
If we can relate this physical orientation flux field to the forces, we
will have specified decision process theory as a dynamical theory. We
make a start at this in the next section where we explain how hidden
dimensions lead to the creation of decision process theory and the payoff
fields suggested in Eq. (1.26) are identified as frame rotation effects.

2.11 Decision Process Theory

The challenge is bridging the view we have been developing. On the one
hand, that view approaches decisions in terms of the strategies or choices
available to a player, considered as applied dimensions of a many-body
system. On the other hand, that view sees the dynamic content suggested
from game theory as resulting from payoffs for each player.
We find similarities with a challenge posed in physics to unify the
applied forces viewed in space and time with internal symmetries such as
electromagnetic fields. This is by no means obvious. Suffice to say at this
introductory point that a solution by Kaluza [1921] and Klein [1956] was
proposed almost a century ago to consider the electromagnetic fields as
an internal symmetry or hidden symmetry represented by an additional
dimension to be added to ordinary space and time. This only works if the
additional dimension plays no active role in the dynamics in a sense that
we make precise below. In some ways it is like a constraint as opposed to
one of the applied dimensions.
We shall see that the analogous mechanism for decision process
theory is to add a player-specific (internal) strategy for each player,
Dynamics of Decision Processes 65

which we call inactive from the usual perspective of strategies. The usual
strategies we shall call active. This approach, we believe, resolves the
conceptual problems and provides the necessary bridge. We use this as
the framework for decision process theory.
It is thus our view that decisions are carried out jointly by one or
more subjects who act based on their personal knowledge of past actions.
There is some attribute or behavior of each player or agent that is
persistent; that does not change with time. The notion of persistence is
tied with our notion above that the player-specific strategy is inactive. In
contrast to previous authors, we view that the players’ active behaviors
or strategic choices need not be persistent, rational or necessarily
converge to some equilibrium flow. Their choices in general are variable
precisely because they learn from past outcomes and are affected by
network relationships.
Following this idea, we add inactive dimensions that we associate
with the players in the decision, to the active strategy dimensions
associated with the pure strategies. In a coordinate basis, the most
general form of the line element Eq. (2.22) with hidden dimensions  j
is:

ds 2   jk  d j  Aaj dx a  d k  Abk dxb   gab dx a dxb . (2.42)

We indicate active strategies by the indices a, b, and the inactive or
hidden strategies by the indices  j, k , . We define hidden to mean that
the inactive metric components  jk , payoff potentials Aaj and active
metric components gab , are independent of the inactive dimensions  j .
For the purpose of this discussion time is active. The hidden or inactive
dimensions play a restricted dynamic role.
In differential geometry, when there is a frame in which the metric
elements have the form Eq. (2.42) and are independent of the dimension
 j , then there is an isometry associated with that dimension: the metric
does not depend on that dimension. As part of decision process theory
we postulate that there is at least one isometry for each player in a
decision. To the extent that players may exhibit schizophrenic
tendencies, there may be more than one isometry for each player.
66 Geometry, Language and Strategy—Vol. 2

We view players as being independent and so the isometries lead to


metric elements that are simultaneously independent of all inactive
dimensions. For the mathematically oriented reader, each isometry
implies the existence of a vector field. These vector fields when
considered as operators form a Lie algebra. The associated group for
these players is a commuting group: two successive transformations
yield the same result independent of order.
The existence of these isometries leads naturally to the notion of
payoff fields, which we now illustrate. We observe that in the expression
for the line element Eq. (2.42), in what we call the normal-form
coordinate basis, the natural choice for a basis is a coordinate or exact
basis for the active directions, U a  dx a by which we mean dU a  0 , but
not for the inactive dimensions

U j  d  j  Aaj dx a ,

by which we mean dU j  0 :

Ua  dxa
. (2.43)
U j  d j  Aaj dxa

We show that the payoff potentials A j a lead to payoff fields F j ab ,


Eq. (2.47) below.
We first note that these frames remain unchanged if to the inactive
dimension we add an arbitrary function  j of the active variables

 j   j   j  x

and to the payoff potential subtract a gradient of this function,

Aaj  Aaj   a  j .

Such a transformation is a gauge transformation associated with player


j . It is a subset of the general class of gauge transformations Eq. (2.36).
As such we expect laws to be expressed in terms of tensors that are
invariant under such a transformation. By construction the inactive and
active metric elements are independent of these transformations.
Dynamics of Decision Processes 67

To identify the payoff potential with the payoff matrix that occurs in
the flow Eq. (1.13) suggested by game theory we compute the behavior
of geodesic curves defined in terms of the (vector) flows:

dx a
Va 
d
d j
Vj
d . (2.44)
a
DV
 V ;bV  V ;kV  0
a b a k


DV j
 V j ;aV a  V j ;kV k  0


A geodesic curve is characterized in geometric terms as a curve that is


the shortest distance between two points. For example on the surface of
the sphere, the geodesics are great-circles. Though simple in form, in the
normal-form coordinate basis there will be orientation potentials in
addition to the derivatives of the flow.
In order to relate these orientation potentials to the metric
components in Eq. (2.42) we use the concepts described in the previous
section. We use the inverse of the transformations in Eq. (2.43) as well as
the frame partial derivatives:

 U j U bj    kj Abj 
U  U     ka a 
 
U k U b   0  ba 
 U kj U bk    jk  Abk 
U  U    a
1 

  U j U ba   0  ba 
. (2.45)
  
 j  U j     j
 a  U a     a  Aak  k

On the space of functions that are independent of the inactive


coordinates, we see that  j gives zero and  a   a . The line element
Eq. (2.42) can be written in the normal-form coordinate basis:
68 Geometry, Language and Strategy—Vol. 2

ds 2   jk U j U k  g ab U a U b   jk U j U k  g ab U a U b . (2.46)

In this basis, the inactive and active metrics are orthogonal.


We obtain the orientation potentials using Eq. (2.32) by first
computing the structure constants3 C   from examining the 2-forms
constructed from the differentials of the frames, where the matrix F j is
shorthand for the payoff field F j ab   a Abj   b Aaj defined in terms of the
payoff potential:

dU a  0
dU j  F j  1 2 C j ab U a  U b
C a   0 . (2.47)
C j kl  C j ka  0
C j ab  Fabj

In other words the active components are exact and the inactive
components are orthogonal to the active components, but internally are
not orthonormal. So the active orientation potentials are determined by
the active metric. We still need to verify that the payoff field that appears
here is also the one associated with the flow equation. To do this we need
the orientation potentials that appear in Eq. (2.44).
The orientation potentials that have one or more inactive components
are determined algebraically from Eq. (2.32):

   1 2  C  C  C   1 2  Δ    Δ     Δ   


 jkl  0
 ajk  1 2   Δa jk  Δ j ka  Δ k  aj    1 2  a jk
 jak  1 2  Δ j ak  Δ a kj  Δ k  ja   1 2  a kj

3
Greek indices cover both active and inactive dimensions.
Dynamics of Decision Processes 69

 jka  1 2   Δ j ka  Δ k  aj  Δ a jk   1 2  a jk
 jab  1 2  C jab  Cabj  Cbja   1 2  jk Fabk
 ajb  1 2  Cajb  C jba  Cbaj    1 2  jk Fabk . (2.48)
 abj  1 2  Cabj  Cbja  C jab    1 2  jk Fabk
 abc  1 2   a g bc   b g ca   c g ab 

This result is insightful. The orientation potentials with three inactive


coordinates are zero. The orientation potentials with three active
dimensions depend entirely on the active metric as if the inactive metric
and payoff potentials were absent. The remaining orientation potentials
are determined either by the gradient of the inactive metric or the payoff
field for each player.
The payoff fields generate vorticity. We use the components of the
orientation potential in the expression for the geodesics Eq. (2.26) for the
active flow:

DV a dxb d j
 V a ;b  V a; j  V a;bV b  V a ; jV j  0
 d d
dV a
  a cbV cV b   a jbV jV b   a bjV bV j   a kjV kV j  0 . (2.49)
d
dV b
g ab   abcV bV c  Vk Fabk V b  1 2 V jVk  a jk
d

The third equation is a statement of decision process theory. We obtain


the result that the game theory form Eq. (1.13) is obtained when both
active and inactive metric tensors are constants. From Eq. (1.26), we
identify the inactive flow Vk roughly with  ek m , with a different
conserved charge for each player.
Note that the game theory form is a special case, not part of the
theoretical foundation. The theoretical foundation gives us insight into
the mechanisms of decision behaviors. In particular, it has been shown in
Vol. 1 that the decision flow above describes a vortex behavior. In this
volume we use our approach to understand better the forces that generate
that behavior from individual choices.
70 Geometry, Language and Strategy—Vol. 2

We see that the payoff fields have two distinct contributions based on
the decomposition of the payoff field into its valuation and negotiation
fields. There is one force (acceleration) along the valuation field
direction and is proportional to the utility density. There is a second force
due to the negotiation field. Unless the flow is in the two-dimensional
plane defined by the negotiation strategies, this force is zero. The force
is magnetic: it is orthogonal to the flow and to the normal to the plane.
Negotiation leads to change when one of the strategies is along the
direction you want to go and the other is an independent alternative with
a non-zero payment.
We expect additional contributions both from the orientation
potentials in the active space abc as well as from the scalar gradients of
the inactive metric. We see that the inactive space though hidden
contributes to the dynamics, albeit in a restricted way. As an example
from the physical domain, we note that the rotation of the earth is also
hidden because we think we are at rest; the Coriolis force is the evidence
from the orientation potentials that we are spinning.
The inactive flow appears as a “charge” density4 for player k , which
we can call the interest flow for player k . We now extend game theory
by exploring the properties of the player interest flow. We demonstrate
that its value stays the same (is conserved) along a geodesic. An attribute
of a spinning body is the conservation of angular momentum. There is an
analogy to decision theory. The geodesics along the hidden directions
can be computed from Eq. (2.26):

DV j dV j
   j cbV cV b   j kbV kV b   j
V cV k  0
 d
ck

dV j
d
  jl lk V k  0 . (2.50)
d d
dVk
0
d

In other words the interest flow V j is conserved or persistent along a


geodesic. There are no dynamic changes along the hidden direction since

4
To be more precise, the inactive flow is the ratio of the charge to the mass.
Dynamics of Decision Processes 71

its “momentum” is constant. It is a general rule that for each isometry, in


addition to a payoff matrix, there is a conservation law and a conserved
charge. In this case we can say that the player’s interest flow is
persistent. Based on these equations, we can also say that there will be
two associated forces: a Coriolis force and a centripetal force.

2.12 Outcomes

We have completed our first pass: an in-depth description of decision


process theory. It is kinematic in content since we have not addressed
inertial forces, which will be covered in the next chapter. We have
created a decision process theory that contains significant aspects of the
theory of games as a special case, but the framework is by no means
derived from the theory of games. Our goal was to present a
comprehensive and robust framework in which to discuss decisions. By
their nature decisions involve complex strategies and agents, which we
have identified with the game theory name player. These players make
choices that in fact may change over time and space. We have been led
to a framework in which the strategic choices are active dimensions of
space, the players provide one or more inactive strategies each of which
is hidden but manifests as a payoff field associated with that player. The
geodesics of this geometry provide a behavior that replaces Eq. (1.13),
which in game theory represented not dynamics but a mechanistic device
to identify the equilibrium strategies.
We believe decision process theory is a realistic and self-consistent
extension to the discussion of decisions from game theory. We reaffirm
our belief that decisions are acts in this real world and can be understood
using the ideas and language we have learned from a study of that real
world. Though the subject matter involves human and society
interactions, we believe a deterministic and empirical method applies.
We assert that the best way to study decisions is to have a robust and
self-consistent decision process theory within which we frame our
arguments. We wish to avoid what we see as a common practice of using
a piecemeal approach to interesting problems in which one tailors a
model to fit the facts.
Having come this far however, we believe we must extend decision
process theory further to be able to include other interesting aspects of
human decision making. There is no question but that there are forces
72 Geometry, Language and Strategy—Vol. 2

that we have not dealt with, which we call inertial forces. They are
constraints and play a necessary role in the realm of decision making and
are so far absent in our theory. In the next chapter we explore such
inertial forces using as a learning example, the behavior of fluids and
suggesting how those effects may be included in the theory of decisions.
We here summarize the chapter from the standpoint of using this
volume in a course. By the end of the chapter, the student should be able
to do the following using this framework:
 Understand global behaviors of the theory
 For specialized cases be able to apply global theory to realistic
decision processes
 Be able to identify and use the underlying connections of physical
theory to gain insights into decision processes
 Be able to classify general decision processes in terms of a standard
taxonomy.
The attainment of these outcomes will be facilitated by doing the
exercises in the next section. As a help, the detailed outcomes expected
of the student are listed below.
 Understand units that are used in decision process theory—Sec. 2.1.
 Understand the concepts of geometric forms—Sec. 2.2.
 From Sec. 2.3 understand the source-free Maxwell equations as
expressed in the language of forms.
 From Sec. 2.4 understand the Maxwell equations that arise from
sources as expressed in integral form.
 Be able to translate these equations into the language of game
theory—Sec. 2.5.
 The student should learn the essential point that aspects of decisions
are deterministic, generate actions and are influenced by actions—
Sec. 2.6.
 From Secs. 2.7-2.8, be able to distinguish between transformations
that are holonomic and those that are not. Furthermore, actions are
related by cause and effect, which propagate at a finite speed.
 The geometric properties of these effects are determined by the
common sense properties that transformations are torsion free and
measure independent, and these properties determine the orientation
potential in terms of the metric and structure constants—Sec. 2.9.
Dynamics of Decision Processes 73

 From Sec. 2.10, the student will have learned that the concept of
covariance is that the rules of the theory are expressed in terms of
geometric structures that maintain their property over time and space.
The key such properties are scalar fields, vector fields and tensor
fields. A consequence of the theory is that cause and effect are carried
by the orientation flux field, called the curvature tensor in differential
geometry. It measures the degree to which global space-time is not
flat.
 From Sec. 2.11, the student will have learned that hidden dimensions
provide the mechanism to create decision process theory. Like the
Coriolis acceleration, payoffs reflect accelerations associated with
hidden dimensions and provide a game theory mechanism for
strategic behavior in decision process theory. They also provide a
persistent conserved quantity analogous to charge conservation in
physics.

2.13 Exercises

(1) In Secs. 2.3-2.5 we have equations similar to electromagnetic theory


in n dimensions. A) To better understand this relationship, show
that the fact that parallel displacements of a coordinate differential is
zero, Ddx a  0 , means that we can parallel transport our clocks and
measuring devices (for strategic values or preferences) locally
without change. B) Show that for each coordinate u , the
contravariant vector  ua is associated with the surface orthogonal to
the direction u , in terms of the 1-form Su   ua Sa . Furthermore, its
integral curve dx a du   ua implies that x u  u . Show that this
implies that the normal to the surface is along the integral curve, the
distance along the curve is the coordinate and such distances can be
used locally to compare values and times.
(2) Derive the form of the orientation potentials, Eq. (2.48).
(3) Derive the active geodesic flow (without sources), Eq. (2.49).
(4) Derive the inactive geodesic flow (without sources), Eq. (2.50).
(5) The Coriolis acceleration can be expressed as saying that air flow in
the Northern hemisphere will move from a high to a low pressure
74 Geometry, Language and Strategy—Vol. 2

zone with a bend to the right [Crawford, 1978]. Show that Coriolis
forces arise because of the spin of the earth and can be expressed as
a non-diagonal metric term gt   tt A , [Ryder, 2009]. The vector
potential A gives a characteristic payoff z  F and a velocity
( v ) force term ω  v . Assume points on a sphere are measured in
latitude  and longitude  . Show that this follows from the
conservation law Eq. (2.50) with time as the inactive coordinate. If
there is azimuthal symmetry, show that this symmetry produces a
centripetal acceleration  r   .
(6) Show that the earth’s rotation produces forces that represent three
dimensional cyclonic flows that rise (increase the distance r from
the center of the earth), and move for example from the equator
towards the pole, fall and then travel back to the equator. Such
flows divide the earth into atmospheric cells that play a critical role
in global weather formations [Crawford, 1978]. They generate
frame orientation effects that would be unexplained if we were
unaware of the rotation.
(7) To obtain a geometric visualization of Eq. (2.19), imagine a small
area of space in which a thin and short hyper-cylinder of current is
along the z  axis transverse to the cylinder plane with radius r .
The surface of the cylinder is a n  2 dimensional sphere. In this
region, show that the payoff flux components F rz are the only
contributions (approximately) to Eq. (2.19) and that they are
multiplied by the hyper-length S rz . Discuss the meaning of these
flux components and the components of Frz in Eqs. (2.10)
and (2.13).
Chapter 3

Inertial Behaviors

In the last chapter we arrived at a kinematic formulation of decision


process theory as a self-consistent consequence of decisions viewed as
events in a space of strategic choices in which some of those choices are
hidden and associated with players or agents. In a population of 10, 100
or 1,000,000 players, it is possible to imagine that the theory might apply
even though actually carrying out the calculation might be horrendously
complicated and not practical. We generally focus on fewer players or
agents and treat the rest of the players or agents as part of the
background, part of the constraints. The background may also have
environmental and physical effects that might influence decisions. This is
the situation envisioned in our discussion of applied and constrained
forces, Sec. 1.6.
Our view is that the dynamics of systems is governed by all of the
forces. We don’t need the same detail depending on whether the forces
are applied or constrained, as long as we know the energy and
momentum for each. This is dealt with by separating out the dimensions
that are active from those that are inactive. These dimensions correspond
to the active and inactive strategic choices available to each player in
making a decision. The dynamics is determined by the kinematics of the
last chapter and by additional contributions that are not kinematic in
nature.
The non-kinematic part of the background is still described by the
active decision behaviors. Each decision is an act that occupies a point in
strategic space-time. Strategic occupation has particular attributes that
are commonly identified: an occupied space is identifiable and may
attract behavior towards it or around it. Some of these behaviors are

75
76 Geometry, Language and Strategy—Vol. 2

included in the kinematics of the last chapter. Other behaviors are based
on the constraints not included. Newtonian mechanics views all
background forces as being generated by characteristic masses and their
accelerations. In more general theories such as Maxwell and Einstein
theories, the background forces are characterized by their energy
densities and energy flows. In this way we take into account background
forces. We call these the inertial attributes and incorporate them into our
theory.
Looking purely at the kinematic or non-inertial aspects, we provided
in the last chapter a motivation for incorporating the payoff field of each
player into decision process theory. We borrow the term “inertial” from
the physical sciences: an inertial force reflects the resistance of a body at
rest to move or a body in uniform motion to change direction. In
Newtonian physics it is the product of the mass and the acceleration. By
extension, an inertial frame is one in which these attributes of
Newtonian physics hold. By contrast, a non-inertial frame is one in
which these attributes don’t hold. Two common examples are bodies in a
gravitational field and the Coriolis Effect seen by an observer on a
spinning globe. A slightly less obvious example is an observer on the
inside of the spinning globe feeling the effect of the centripetal
acceleration. We argue that vorticity or spin also plays a role in decision
making and so we anticipate the importance of non-inertial frames.
We say gravitation is non-inertial because the acceleration of an
object is independent of its mass: it reflects the equality of inertial and
gravitational mass. In this language, the effects of the payoff field are
non-inertial and are due to looking at the problem in a non-inertial
frame: these are non-inertial forces. In decision process theory, a system
subject only to such non-inertial forces follows a geodesic, Eqs. (2.49)
and (2.50). It is the generalization of Lagrange’s Eq. (1.9) with only the
kinetic term L  T . In this chapter, we expand decision process theory to
include inertial forces.
Our reason for expanding the theory is the belief that in its current
form, the theory though already an extension of game theory, does not
yet describe important effects that are clearly part of realistic decision
processes. In the business world we are aware of inertial forces that
reflect how hard it is to move some organizations to become more
Inertial Behaviors 77

productive, less wasteful, or towards other desirable goals. We are also


aware of applied forces on organizations that reflect quick learners and
adopters of best practices because they observe competitors attempting to
do the same or similar jobs. Any study of dynamics of necessity deals
with two such contradictory mechanisms: one mechanism staying the
course and the other making radical changes. We have borrowed the
terms characterizing these behaviors from the study of physical systems
and say that there are inertial forces that provide the resistance to change
and applied forces that induce systems to change. We refer to both as
inertial behaviors. You may think of the inertial forces as arising from
effects of family, community or nation; these forces are societal forces
that are not captured by a consideration solely of the individual
interactions.
Are these non-inertial forces important? We believe the answer is yes.
We can’t ignore friction in mechanics or viscosity in fluids if we want to
describe the most interesting effects. We can’t ignore resistance in
electrical circuits and build useful applications. At a deep level, even
mass is a background that one should not ignore, even though unified
theories start with massless elementary particles. This chapter is a study
of such inertial effects in decision making.
In Secs. 3.1-3.2 we determine the general form for inertial behaviors.
We find that for decision process theory, the general form can be
expressed in terms of a tensor that describes the energy and momentum
of the system, Secs. 3.3-3.4. To obtain this result we first recall
Lagrange’s Eq. (1.9) and note that it follows from a principle of least
action. For inertial media (such as fluids) that carry mass, this principle,
Sec. 3.5, will give us the mechanism for formulating a general principle.
To get this principle, we include forces that reflect the resistance to
change of any decision process as well as forces that reflect learning or
applied forces. These effects are not included in game theory or in our
treatment from the last chapter for non-inertial effects due to active and
inactive strategies, Secs. 3.6-3.7. The effects we are looking for are
attributes of (massive) continuous media. We determine the energy
momentum tensor for fluids, an example of (massive) continuous media
and then revisit the electromagnetic fields and show that such fields are
(massless) continuous media.
78 Geometry, Language and Strategy—Vol. 2

We then provide a general characterization of inertial behaviors in


terms of the energy and momentum fields and show the principle of least
action for such behaviors. We provide a working model for the energy
momentum tensor that we use throughout this book. In Secs. 3.8-3.9, we
conclude with advanced topics on some potential limits of our theory
based on the limits of the principle of least action in physical theories.
We also elaborate on the notion of uncertainty versus frequency.

3.1 Inertial Media

The characteristics of decision processes that we see missing are the


resistance to change and the influence to change by neighboring
processes. This inertial cohesion is distinct from those influences to
change that depend on payoffs and exist without neighboring processes.
We call such coupled neighboring processes an inertial media. Financial
markets provide an example of such mutual interactions. In this section
we investigate the form of such interactions. We borrow from the
physical sciences the analysis of continuous inertial media. As in the
physical sciences, forces generate acceleration. We find a form of the
mutual interactions that can be expressed using the Lagrangian
formulation Eq. (1.9), which has a geometric interpretation using the
principle of least action.
The geometric interpretation arises as follows. Along any path in the
conceptual space, the action is defined as the path dependent integration
of the Lagrangian times the incremental interval of time:

S  Ldt .
path
(3.1)

There is one path that is special, the one in which the action is an
extremum. To find this path imposes a condition on the Lagrangian,
which leads identically to Eq. (1.9) and so we are justified in stating that
all mechanical systems obey the following principle of least action:

S   Ldt  0 .
path
(3.2)
Inertial Behaviors 79

The geodesic paths described in the previous chapter are examples of


such an extremum. For each of these examples, there is a well-defined
action and the minimization of the action is the same as finding the
shortest path (geodesic). The principle of least action provides an
insightful way to view complex systems.
To understand the dynamic behavior of inertial media, we note its
similarity to matter. Matter comes in a variety of forms from gases and
fluids to solids such as glasses, metals and crystals. These are continuous
systems whose parts mutually interact in a way that is similar to how we
see inertial media interacting. We approach the study of matter by
partitioning it conceptually into small cells, an approach ascribed
[Tolman, 1987] to Laue. We compute the action of each cell and add
these together:

S    Lcell dt . (3.3)
cells

The smaller the cells, the better will be the approximation. The
Lagrangian Lcell for each cell is determined by the kinetic energy Tcell
minus the potential energy Vcell . It is convenient though not mandatory to
consider the cells to be cubes and use continuous variables to describe
their essential behaviors.
We do this, starting with the kinetic energy, which is determined by
the product of the mass density, the square of the velocity of the cell and
the volume of the cell:

Tcell  12   x, y, z, t  rcell rcell dxdydz . (3.4)

We introduce spatial coordinates that represent the center of the cell, a


displacement vector r  x, y, z, t  that represents the relative motion of the
fluid (and hence is the dynamic variable for the Lagrangian) at a moment
in time t and the average density of the cell at that point and time as
  x, y, z, t  . If the cell is sufficiently small, the average density is
effectively constant over the whole cell at the given moment. Similarly
the velocity of the relative displacement at that point is a vector function
of the coordinates, rcell  V  x, y, z, t  .
80 Geometry, Language and Strategy—Vol. 2

The cell is being compressed and stretched by the surrounding forces


as well as turned and rotated. For the inertial media of decision
processes, the compressions and rotations in strategy space reflect the
possible influences on one system due to all the other neighbors. Because
the cell is a cube1, one can obtain the equations of motion from a
consideration of the forces on each of the opposing faces. For inertial
media, this is generalized to a cube in more dimensions.
In three dimensions, there are three forces on the face orthogonal to
the  x direction. One of the components is along this direction and the
other two forces are transverse along the other coordinate axes. The
changes in the forces from one face to its opposite are small and
proportional to the area, so we use the characteristic of continuous
matter that there is a relative stress defined as the force per unit area
along that face. For our first case, we name the components of stress
 
t xx t yx tzx and look to obtain equations that relate the stresses to
the rate of change of the momentum.
Along the x direction, given the distance between faces is dx , the
components of stress along the x direction t xx dx x txx depends on the
gradient of the stress; the net stress is therefore (minus) the gradient
 dx  x t xx . In a similar fashion the other two stress components on this
face are dx xt yx and  dx x t zx . To get the total force along the x axis
we add the component from this face,  dx  x t xx , with the face orthogonal
to y resulting in dy y txy and the face orthogonal to z resulting in
 dz  z t xz :

f x dxdydz  dydz  dx x t xx   dxdz  dy x t xy   dxdy  dz x t xz  . (3.5)

This gives the total force along the x axis due to the forces being exerted
from all six faces.
There are similar expressions for the other two components of force,
all of which can be compactly summarized as follows, where there is a
sum over repeated indices that represent the three coordinate axes:

1
Space can always be partitioned into cubes. What is important is the partitioning, not
the exact shape. If other shapes are considered, it may be necessary to consider more than
one such shape to “fill” space.
Inertial Behaviors 81

f j  k t jk . (3.6)

In Newtonian mechanics, the inertial force is the product of the mass


and acceleration and is therefore the time rate of change of the
momentum of the cell, which can be given in terms of the momentum
density g j as g j dxdydz  Vj dxdydz . The inertial force is balanced by
the net force caused by the stresses f j dxdydz :

d
dt
 g j dxdydz   f j dxdydz  k t jk dxdydz . (3.7)

The resultant expression requires that we include the rate of change of


volume and then simplify it to provide a general description of the
behavior of matter in terms of the momentum density g j :

g j
  k Vk g j    k t jk . (3.8)
t
For a continuous media, these are Euler’s equations which represent the
conservation of momentum: the rate of change of the flow times the
matter density is the transverse gradient of the stress.

3.2 Inertial Cohesion

Note that the generalization to inertial media is accomplished by


extending the sum so that it goes over the total number of spatial
strategic directions: the form of the equations remains the same. We now
have the first result in terms of the inertial cohesion represented by the
stresses t jk : changes in the stresses generate changes in the momentum
of the system. Since the momentum of the system depends on its inertia,
we are justified in our assertion that the applied effects are inertial in
character.
The second attribute of fluids is the conservation of mass, based on
the observed fact that the net matter flowing into a cell causes the
amount of matter inside that cell to increase by the same amount:
82 Geometry, Language and Strategy—Vol. 2

d 
  jVj   Vj  j    jVj  0 . (3.9)
dt t
This result reflects one attribute of the resistance to change of inertial
systems. It generalizes to inertial media by extending the summation to
all the spatial directions.
The momentum density and matter density are thus determined by
two coupled equations:


 jgj  0
t
. (3.10)
g j
  k Vk g j    k t jk
t
These two equations generalized to inertial media represent the major
result for this section, though not yet written in the most illuminating
form. We will write these results using the principle of least action. To
incorporate these results into decision process theory, we will further
rewrite them in a fashion consistent with the terminology from the last
chapter.
To achieve these goals, we put the equations into a form that
combines both the space and time aspects together in a covariant form.
We introduce the absolute stress:

p jk  t jk  g jVk  t jk  V jVk . (3.11)

For a frame in which the media is moving, the stress tensor t jk will in
general not be symmetric; however the absolute stress will be symmetric
(Tolman, 1987, p. 65ff).
We next rewrite Eq. (3.10) in terms of these absolute stress
components:

g j
  k p jk  0
t . (3.12)

 jgj  0
t
Inertial Behaviors 83

The second equation shows that the increase in matter is due to the
amount of momentum that flows into the cell; the first equation shows
that the increase in momentum is due to the amount of stress that flows
into the cell.
As an aside, a good idealization for many common materials is that of
a perfect fluid, one in which the absolute stress is determined by a single
quantity typically called the pressure: p jk  p jk , where the “delta”
notation indicates that the value is unity whenever the indices are the
same and zero otherwise. The above equations then become:

V j
   k  V jVk  p jk 
t . (3.13)
d
  jV j  0
dt
At low velocities, the first equation is Euler’s equation that the rate of
change of momentum is minus the gradient of the pressure. When there
are no applied forces, a system in constant motion remains in motion; a
system at rest remains at rest. These are the expected effects for inertial
forces.
For decision process theory if we adopt the concept that inertial
effects are characterized by a density  that is a continuous function of
position and time, we expect to get equations of this type as well. The
conservation law (second equation) in Eq. (3.13) expresses the
conservation law for inertia, namely the net flow of inertia into a cell
determines the rate at which the amount of inertia inside that cell
increases.
Returning to the general case, we represent these results with a matrix
that adds to space a new dimension proportional to time: x0  ct . The
proportionality constant has units of a velocity so that the new dimension
has units of distance. We define the components of a tensor, a field with
two indices, Tab as follows:
84 Geometry, Language and Strategy—Vol. 2

T jk  p jk
T0 j  cg j
. (3.14)
T00   c 2
Tab  Tba

The absolute stress, the momentum density and matter density combine
to form a symmetric energy-momentum stress tensor. The conservation
laws of momentum and matter in this notation are:

 0T j 0   k T jk  0
. (3.15)
 0T00   jT0 j  0

The component T00 has the units of energy per unit volume, an energy
density, since the proportionality constant is a velocity and mass time
velocity squared has units of energy. Similarly the component T0 j has
units of energy density since it is a momentum density times a velocity.
We now have a form for the inertial media expressed in the notation
from the previous chapter:

g bc  bTac  0 . (3.16)

For the general case we therefore characterize inertial cohesion using a


conserved energy momentum tensor. Many systems including non-
inertial systems have such conserved tensors. Before examining the
important example of payoff fields that also have this form, we reflect on
our result.
Newtonian physics starts with the notion that the inertial effects result
because bodies have mass. In Newtonian physics, bodies with mass are
idealized as rigid bodies. More significantly, Newtonian physics requires
rigid bodies in order to define the measurement process: lengths are
measured using rigid rulers. Nevertheless, within that Newtonian
framework it is possible to describe continuous bodies. They are the
opposite of rigid bodies, showing that as one part moves, the effect is
communicated to neighboring parts through the local continuity of
momentum and matter. It is clear that such effects do not move
instantaneously, but at a speed characteristic of the medium. By contrast,
Inertial Behaviors 85

a rigid body communicates its motion instantaneously to all other parts.


Our best observations to date indicate that such instantaneous
communication does not happen. Thus, matter as we know it, reflects the
properties of a continuous medium or field. For this reason we consider
inertial effects as being associated with continuous media rather than
rigid bodies.
Einstein noted that general relativity is the physical theory that
properly describes continuous media, though for consistency it must
depart from Newtonian mechanics. In the theory of relativity, all
phenomena are described by an energy momentum tensor, reflecting
continuous media. In the next section, we give an exemplar phenomena
of continuous media—electromagnetic phenomena—that is not
associated with mass or inertia. In decision process theory, payoff fields
are a generalization of such phenomena.

3.3 Unifying Inertial Media with Payoff Fields

We will write a consistent principle of least action for inertial media (the
generalization of matter) and for payoff fields (the generalization of
electromagnetic fields). We saw in the last section that inertial media
communicate effects with a finite velocity, not instantaneously. This is
also a striking aspect of Maxwell’s equations and their payoff field
generalizations. Information is transmitted from one part of space to
another via the electric and magnetic fields at the speed of light as
indicated by the wave equations Eq. (1.16). We expect that this will be a
general consequence of the principle of least action for any type of
continuous media.
The action for the electromagnetic field and its interaction with
matter is:

S    1 4 Fab F ab  V a Aa  dxdydzdt . (3.17)

The extremum for this action yields Maxwell’s Eq. (1.24) that includes
the interaction of the matter with the electromagnetic field. The major
difference between matter and radiation is that the electromagnetic
86 Geometry, Language and Strategy—Vol. 2

radiation is massless: there is no frame in which light is at rest. Signals


move more slowly through matter than they do through radiation. The
electromagnetic field behaves like a continuous medium, with its own
energy momentum tensor that follows from the above action.
The energy momentum field for the electromagnetic field that results
is:

Tab  Fac F cb  14 gab Fcd F cd . (3.18)

In the absence of matter, the equations of motion Eq. (1.24) are


summarized succinctly by the requirement that the energy-momentum
tensor is conserved, an equation identical to Eq. (3.16):

g bcbTac  0 . (3.19)

This tells us that in the absence of matter, the electromagnetic


phenomena move in a way to conserve momentum and energy2. In this
case, the energy is that of the field and has no connection to any rest
mass. The energy component T00 is 1 2 E  E  1 2 B  B , the sum of the
electric field squared and the magnetic field squared3. It is positive and
independent of matter. The continuity equation describes the continuity
flow of energy. Similarly one can compute the momentum for an
electromagnetic wave, which is formed from the product of the electric
and magnetic fields. The momentum points in a direction orthogonal to
both. Thus momentum travels as a wave orthogonal to the electric and
magnetic field vibrations and does so as if the vibrations were in a
continuous medium.
Since light is a particular type of electromagnetic wave, the study of
optics provides a good example of the character of the waves under

2
The energy-momentum tensor is also conserved in the presence of matter. We would
write the total energy-momentum tensor as the sum of the electromagnetic field and the
matter field. Each have an energy-momentum that is conserved separately.
3
An electrical engineer understands that energy is really stored in capacitors as the
electrical field and in inductors as the magnetic field. These fields are observed to be real
and often quite powerful.
Inertial Behaviors 87

specialized circumstances in which the wavelengths are small compared


to the size of the “lens”.
In this case the wave phenomena appear to be more particle-like, with
light traveling along paths of least time (which is a special case of least
action). If light goes through water it travels slower than going through
air and this leads to the well-known bending of light at the boundary. The
often quoted example is that of a person on the beach wishing to save
someone; since they can run faster than they can swim (normally); it is
faster to run towards a point that has less swimming distance, than
directly at the person drowning. This optimizes their time to the
drowning person.
Of course it does raise the question of why light travels at a different
speed in matter than in a vacuum. The deeper answer would again
require us to have the action for light and matter, where we would then
see that light interacts with charged particles (the electrons) and gets
absorbed by atoms and re-emitted. The effective interaction is
characterized by saying the light moves slower, though the deeper
explanation would allow one to compute the amount by which light
slows down depending on the properties of matter. The slowing of light
is described empirically by assigning to matter an index of refraction that
specifies the velocity of light in that medium.
We are now in a position to write a consistent principle of least action
for inertial media (the generalization of matter) and for payoff fields (the
generalization of electromagnetic fields). The simplest unification is a
single conserved energy momentum tensor that contains both fields and
is derivable from the principle of least action. We give below a more
sophisticated unification in terms of the action, Eq. (3.24) and the
conservation law Eq. (3.26). The more sophisticated unification requires
a change in perspective from the way we are used to thinking about
measurements (geometry) and continuity (topology). We assert that this
thinking is relevant and necessary for decision process theory. To
construct a dynamic theory of decision processes, these are precisely the
concepts we must articulate clearly. We outline the change of perspective
that was needed to describe the physical world and indicate the
generalizations needed for our theory.
88 Geometry, Language and Strategy—Vol. 2

We recall that the Newtonian view is based on our common sense that
certain forces, such as gravity and charge are instantaneous. We showed
that this view, however, is not consistent with Maxwell’s Eqs. (1.20) and
(1.24), which show that the Coulomb force Eq. (1.17) propagates with
the velocity of light. The Newtonian view also maintains the common
sense view of the existence of rigid bodies, which have the same defect
of propagating effects instantly from one part of the body to another.
However, we have suggested that real bodies are continuous and
communicate from one part of the body to another at a finite speed. We
were thus led to describe inertial effects by means of inertial media.
Einstein unified these viewpoints into a theory of relativity in which no
instantaneous transmission of forces exists and in which there are no
rigid bodies. We use his new common sense view as basis for our general
decision process theory.

3.4 Distance

To appreciate the philosophical change in perspective that Einstein


requires of us, we discuss geometry and the measurement of distance.
Measurements are made using some type of ruler. They presuppose an
understanding of the space in which the measurements take place. This
space already has some basic topology that distinguishes near from far.
At this level there is no distinction between a coffee cup and a donut:
both are spaces with a hole in them. We distinguish such objects when
we provide a measure of distance between neighboring points. Euclidean
geometry in three dimensions is based on the distance between two
points being given by

 x A  xB    y A  yB    z A  z B  .
2 2 2
s  (3.20)

The coordinates reflect the three dimensional nature of space—the x,


y and z axes of space which are characterized as being orthogonal.
Euclidean geometry further characterizes space as flat, so that the above
formula holds whatever the distance between the two points. The ruler is
assumed to be a rigid body whose existence is not part of the dynamics:
Inertial Behaviors 89

it measures the same at every point in space. Thus it is enough to know


the coordinates of the two points in order to determine their distance.
In Newtonian physics, if the two points represent successive positions
of a body that is moving, then the distance depends on the time
difference of the two points. If the two points are extremely close
together in time then Newtonian physics observes that the distance is
proportional to time; the proportionality constant is the speed and the
velocity is a vector in the direction of motion of the body at that point in
time, whose magnitude is the speed. This is a non-trivial statement about
how measurements occur in time. Many first year students in physics and
engineering in their physics’ lab course take a series of photographs of
rapidly moving objects as a way to “see” the speed. They observe that
the speed as defined as the ratio of distance interval over time interval is
a number independent of the time interval, once the time interval is made
small enough. The ideas of distance, velocity and the rate of change of
velocity (called acceleration) are sufficient distinctions with which to
describe the physical laws in Newtonian physics.
We suggest that for dynamics of any system, there are in fact two
attributes: the topology and the metric. We don’t give up the notion of
measuring, only the notion, following Einstein, that the measuring sticks
are rigid and outside of the dynamic equations. Rigidity is a good
approximation for points that are neighbors, but not for points that are
well separated. It is plausible therefore that we should return briefly to
the notion of the underlying topology and inquire into the nature of what
we mean by local.
An example may be helpful. We are well aware that the surface of the
earth is a (oblate) sphere. From a topological point of view, there are two
dimensions to the surface of the earth, whatever its shape: this is a
topological property. Moreover, the sphere has no “holes”, another
topological property. The sphere has very interesting local properties
however. At any point, there is a continuous, differentiable and one-to-
one mapping of the sphere onto a plane. In normal parlance we call such
a mapping a chart and use such charts for navigation purposes. We
navigate with rigid rulers on the charts, despite the fact that we are
plotting points on a sphere. The differentiable map, or diffeomorphism,
is an exact statement only at the point and a small region around it. We
90 Geometry, Language and Strategy—Vol. 2

know that the charts lose accuracy away from that point, so when we
travel we need a number of charts in order to navigate around the earth.
The charts have to match on their common areas. It is possible to
describe the sphere in this fashion as a series of two-dimensional charts.
The collection of all the charts will then be our atlas. The structure
defined in this way is termed a differential geometry by mathematicians
and used by physicists such as Hawking & Ellis [1973]. It is also studied
in a branch of mathematics called fiber bundles [Steenrod, 1951].
It provides a precise language with which to describe all the attributes of
such structures.
The differential geometry makes only the assumptions at the point
that reflect our direct experience. At the given point we know that any
rotation of the two dimensions yields an equivalent set of charts. In order
to navigate between points we need to know the symmetry structure at
each point and we need to know how to glue the charts together, which is
tantamount to our knowing how a path appears on each chart as it moves
over the surface of the earth so that the charts provide a faithful
representation of the physical geometry. To apply this to any system, we
have to identify the dimensions in which we live, our topological space,
identify the local symmetries that are present for all dynamics and
identify the charts that are needed to describe motion. For decision
process theory, we generalize Einstein’s approach to identify the
topology as consisting of strategic space and time.
Einstein takes the implied form of the metric from Maxwell’s
equations. Consider an electromagnetic wave that satisfies the wave
Eq. (1.16). The waves are real numbers, but are usefully thought of as a
superposition of complex numbers:

Ea  cak eit ik r


. (3.21)
Ba  d ak eit ik r

Electrical engineers use these phasors to simplify calculation and take


real parts at the end. The electric and magnetic fields for radiation are
computed as the real or imaginary parts of this superposition. Based on
the wave equation, the relationship between the frequency  and the
wave number vector k is:
Inertial Behaviors 91

 2  c 2k  k  0 . (3.22)

This expresses the fact that the wave travels with a constant speed c
called the speed of light. Einstein identifies a 4-vector as one in which
the time component is the frequency  and the space components are
the wave numbers ck . With this identification the length of the 4-vector
is determined by the metric we suggested earlier, Eq. (1.23).
Though the Newtonian view and Maxwellian views are different, it is
possible that both are correct in their domain of application. However
there are phenomena that require both theories as part of the explanation.
The speed of light from a moving object should increase or decrease its
speed according to the Newtonian view; the energy (frequency) of the
light should increase, but the speed should stay the same according to the
Maxwellian view. Michelson and Morley demonstrated that the speed in
fact stays the same, showing that the Maxwellian view is more accurate.
The effect they observed was very small, so that no known deviations to
Newtonian mechanics would have been observed. Einstein’s method was
to provide a theory of mechanics consistent with the Maxwellian view.
He took as the observational fact the form of the metric as used in
Maxwell’s equation. Based on what was understood about
electromagnetic phenomena, he postulated that the descriptions of all
physical phenomena would have the same form in any frame of reference
as long as one understands that measurements are done using this metric,
called a Lorentz metric in physics after the physicist who first articulated
its importance:

s  c 2  tb  ta    xb  xa    yb  ya    zb  za  .
2 2 2 2
(3.23)

This is the answer to the question: what is the topological space? It is


a 4-dimensional manifold, diffeomorphic at each point to a pseudo-
Euclidean (Minkowski) space with the metric Eq. (3.23).
Transformations that leave this form invariant also leave the form of the
electromagnetic action Eq. (3.17) invariant when expressed in terms of
the metric. Einstein conjectured that in general, at each point the action
would be invariant under these transformations, thus supporting the idea
that space is a 4-dimensional continuum of values. This conjecture
92 Geometry, Language and Strategy—Vol. 2

requires changes to the Newtonian laws and in fact to the form of


Maxwell’s laws in areas where the space is not flat.
The use of a Lorentz metric to measure distance changes subtly,
though profoundly, the concept of time, which is related to the fact that
energy and momentum densities move with a finite velocity. Newton
saw time as a parameter that had the same value at all parts of the
universe. Einstein said there was no experimental justification for that
assumption; when we use time we in fact always observe only the local
time. If we gave someone a watch and told them to tell us what time it
was at some distant location, they would have to send us a signal
(telephone, radio wave, etc.) that would take a finite amount of time to
reach us and we would know at our local time what they believe is their
local time. We still would not know their time simultaneously with a
measurement of our local time. In other words, we only know the time
relative to our chart; a distant person will use a different chart and we
will have to learn how to match up the charts.
Second, Einstein asserted that time is a quality just like space and the
two form the conceptual space-time for physical events. Space and time
are different from the topological positions; they have a physical reality.
In particular, measurement of distances between two events is now a
quadratic form that involves time. It creates a space that is non-Euclidean
because the measurement of neighboring points is not the sum of squares
of the distances.
The third change of Einstein was to formulate the laws of motion in a
way consistent with this new way to measure distance and consistent
with a new set of mechanical laws based on the same principle of form
that Maxwell’s equations satisfied. Maxwell’s equations (1.24) have the
property that the form of the equations is the same (covariant) in every
frame of reference, irrespective of the relative motion of the two frames.
A special case of this is that the speed of light is always a constant in
every frame.
Einstein’s conceptual space provides an ordering and a topology
(concept of continuity) as well as fields that describe matter, radiation
(such as the electromagnetic radiation) and what is new, a physical field
for measuring distances (at least in a holonomic coordinate basis) called
the metric g ab that carries energy and momentum. It determines the
Inertial Behaviors 93

a b
distance between points gab dx dx as the product of the metric and the
product of two coordinate distances, summed over all such product
possibilities. The continuous metric (media) replaces the rigid measuring
rod. Two observers moving relative to each other will use different
metrics to describe the same physical process. At any point in space and
any instant of time, it is always possible however to find a frame of
reference in which the metric is flat, Eq. (3.23); however at some other
point and some other point in time, this observer will in general not see
the same flat metric.

3.5 Principle of Least Action

We conclude from the previous section the foundational results that unify
inertial fields with the payoff processes of game theory. Since rigid
bodies do not exist in the Einstein view, we say that matter (inertia) is
described in terms of fields. Such inertial fields are determined by the
Lagrange principle of least action in terms of the Lagrangian density,
where we integrate over space and time, analogous to the examples we
have given earlier for the Maxwell fields:

S  L g dxdydzdt . (3.24)

This form for the Lagrangian can be used to specify a form for matter
that is continuous and consistent with Maxwell’s equations and our new
common sense. For decision process theory, this suggests that we take
seriously not only the principle that flow is influenced by the payoff
fields Eq. (2.49) but that we apply the generalized version of Maxwell’s
Eq. (1.24) to decision processes: decision flows are the sources for the
payoff fields. We say more about the exact form of the generalization in
the next section. Here we clarify the relationship between the inertial
fields and the action.
We require the form for the Lagrangian in Eq. (3.24) to be covariant
in terms of the metric. First we assume that the Lagrangian density L is
a scalar, i.e. it has the same value in every frame of reference. The action
is a scalar, so we require the unit of volume to be a scalar. So for
94 Geometry, Language and Strategy—Vol. 2

example if each coordinate were replaced by itself multiplied by the


constant  , the volume dxdydzdt would be increased by  4 . Thus the
volume is not a scalar quantity. An invariant unit of volume is obtained
using the fact that the new metric, will have elements that are   2 times
the old elements; in the special case that the metric is diagonal, the
square root of the product of the diagonal elements, which we call g ,
is   4 . Thus the product g dxdydzdt is a scalar function and has the
same form in every frame. We call g the determinant of the metric
viewed as a matrix and show that the above argument holds for all
transformations and all possible forms for the metric.
The action for the electromagnetic fields Eq. (3.17) can also be put in
a form valid in any frame by correcting the volume element. The
Lagrangian density is already invariant. Moreover, for inertia media,
whose equations of motion follow from the principle of least action
using the Lagrange form above, we can uniquely define the energy
momentum stress tensor by considering the variation of the action  S
with respect to changes in the metric  g ab :

 S   T ab gab . (3.25)

The energy momentum tensor so defined will be a symmetric field,


expressible in terms of the inertia media fields and will satisfy the
conservation/continuity Eq. (3.15) that is now written covariantly, Eq.
(2.26), in terms of the metric:

g bcTab;c  0 . (3.26)

Thus we recover and generalize the description given earlier for the fluid
example of continuous matter, we obtain a description that is consistent
with the observations of Maxwell for charged continuous matter and we
provide a covariant description whose equations are generated by the
principle of least action. We thus accomplish our goal of incorporating
inertial resistance and inertial cohesive forces into decision process
theory through the least action principle Eq. (3.25). We specify possible
models by specifying possible inertial media contributions to the energy
momentum tensors Tab .
Inertial Behaviors 95

Following Einstein then we take the general decision process theory


from the last chapter and add to it the possibility that there are inertial
media (matter) fields that provide inertial forces that can be specified by
an energy momentum tensor Tab . However, the measurement part due to
the metric is new. We have added non-inertial orientation potential fields
that are closely related to the metric, Eq. (2.32), which provide for
measurements. With all of these changes, decision process theory will be
internally consistent. We need to look more closely at how the
orientation potentials generate fields Eq. (2.41), the orientation flux field
and how they carry energy and momentum. We have already seen one
aspect of that by demonstrating that the payoff fields are specific
components of these orientation potentials, Eq. (2.48). Einstein proposed
and we adopt the same general rule that we must have dynamic equations
that unify not only the inertial media and payoff fields, but also the
orientation flux fields. We do that in the next section.

3.6 Orientation Flux Fields

We have proposed a foundation for decision process theory in the last


section, with an implied set of equations that follow from the principle of
least action. In this general sense we follow a strategy consistent with
that followed in the physical sciences. Two new additions to game theory
are: the idea that the payoff fields are generated by the decision flow; and
the existence of inertial fields that influence the dynamics. The first
addition implies that there is a relationship between some components of
the metric fields as determined by the payoff potentials Aa and the
sources. We find in this section that this is a consequence of a more
general relationship proposed by Einstein for physical properties: the
metric is determined by the inertial forces through the principle of least
action applied to the metric fields. We make that argument below with
the generalization to decision process theory in mind.
As part of decision process theory, we have identified a metric that
provides a measure for decision events. Such a measure however is not
outside of the theory but an integral part of the theory. Einstein observed
that measurements are physical processes and we make the assumption
96 Geometry, Language and Strategy—Vol. 2

that for decision processes, measurements are a type of decision process.


Since measurements are constrained by the orientation potentials that we
described in Sec. 2.11, these quantities must also be determined by a
principle of least action. These considerations lead us to the equations
for decision process theory, Eq. (3.31), which follows from the principle
of least action.
The potentials for electromagnetism carry energy and momentum,
implying no instantaneous action at a distance. The requirement for the
orientation potentials must be the same. The analogous field Eq. (2.41) is
the generalization of Stokes’ theorem for a non-commuting gauge group:

R  dω  ω  ω . (3.27)

Stokes’ theorem relates the total field through a bounded area to the sum
of differential quantities around the boundary. For a general group, in
which some elements don’t commute, the total field through an area
bounded by a closed loop equals the covariant differential of the
potential integrated around the loop. At a deep mathematical level, the
orientation flux field above is this covariant differential.
For the electromagnetic field the second term of Eq. (3.27) is zero.
This is not the case for the group of frames, the symmetry group
associated with the orientation potentials. Ordinarily, the orientation flux
field is called the curvature tensor of the space and is written as R . The
notation Eq. (3.27) is compact: the curvature is a matrix of 2-forms that
measures the flow or change of the orientation of a frame through a 2-
surface into the dual of a second 2-surface:

R ab  Rabcd Ec  Ed . (3.28)

The components, the tensor R a bcd are antisymmetric in the last two
indices. The tensor is also antisymmetric in the first two indices. By Eq.
(3.27), the curvature tensor components are determined by the
orientation potentials.
Since we have an equation of action for the electromagnetic field in
terms of the field strength F , we expect a similar principle to hold for
the curvature. Of the many forms that have been found and are
Inertial Behaviors 97

equivalent, the following is one that depends directly on the curvature 2-


forms, and physically represents the flow of the orientation flux field:

S   R ab  * Ea  Eb  . (3.29)

The curvature 2-form is multiplied by a 2-form constructed from the


“dual” or “  ” operator of Ea  E , defined for a general n-dimensional
b

space as the (n – 2)-form:

* Ea  Eb    abcd Ec  Ed . (3.30)

In four dimensions this is constructed from the tensor  abcd that is


antisymmetric in every pair of indices and has the value +1 for  0 xyz . It
can be used to evaluate a determinant and so in fact is used to define the
volume element that is written above as the square root of the
determinant of the metric g times a 4-form. The Lagrangian density
is thus the wedge product of two 2-forms and will be an invariant under
all frame transformations. It is also a scalar under gauge transformations,
so the resultant theory is a gauge invariant theory. We require the
generalization of this to any number of dimensions.
The evidence for the principle of least action using Eq. (3.29) is the
Newtonian law of gravity. In Einstein’s view, gravity is not an
instantaneous force, but one that propagates with the speed of light from
one point of space to another, preserving continuity of the energy and
momentum stress densities. He showed that the various attributes of
Newtonian theory were consequences of this theory, as well as small but
observable changes to the Newtonian view. One consequence is that
gravity is not a static scalar field as it is described by the orientation
potentials whose components change dynamically with time.
The result of the principle of least action with Eq. (3.29) extended to
include inertial media fields is shown [Gockeler & Schucker, 1987] to
result in the following expression:

Rab  1 2 gab R  Tab . (3.31)

In evaluating the least action, the full curvature tensor does not appear,
only two of its contracted forms, Rab  Rdabc g cd and R  g ab Rab . The
98 Geometry, Language and Strategy—Vol. 2

right-hand side is the product of the energy-momentum tensor Eq. (3.25)


of the inertial media fields and a universal constant  . The equations
are gauge invariant and frame covariant. A subset of these decision
process theory equations provides the generalization of Maxwell’s
equations as shown in Sec. 4.11, Ex. 10.
Energy and momentum are conserved using covariant derivatives for
the inertial media field and for the orientation flux field separately:

g bcTab;c  0 . (3.32)

Compare this with our general expressions for the stress tensor for
matter, Eq. (3.19). Here we have the covariance principle that in
describing rates of change, ordinary derivatives are replaced by covariant
derivatives Eq. (2.26).
We have the last major piece of decision process theory from the
consequence of Eq. (3.31). The unified applications of the principle of
least action to all of the fields that participate in the decision process
bring together the orientation flux fields and the inertial media fields.
The inertial media fields reflect both inertial and application forces. All
of these fields carry energy and momentum. All of these fields are real.
All of these fields obey the conservation laws of energy and momentum
Eq. (3.32). Our last result for this chapter will be to investigate useful
models for the inertial media field, which we do in the next section. First
however we summarize a few results for reference later in the book.
We consider some of the consequences of the general form of the
Einstein’s Eq. (3.31). In general, the space will not be flat relative to a
given set of orientation potentials. Energy and momentum create the
source for the orientation potentials, which themselves create a flux that
carries energy and momentum. So although we argue that the space is
locally flat, the existence of non-trivial fields changes how individual
charts are pasted together to describe a more complex global structure.
We have argued that this is no different from looking at the earth in
terms of flat charts locally, yet nevertheless the space is curved. The
curvature is a consequence of how the charts must be glued together; it is
a consequence of the atlas. We measure effects of local curvature when
we look at second order rates of change. For example, we envision this
Inertial Behaviors 99

a
process by looking at a small square with sides dx a and dy . We take a
vector V a and displace it first along the direction dx a and then along
dy a ; alternatively we go first along dy a and then along dx a . The
difference of the two calculations can be evaluated in terms of the
covariant derivatives:

V a
;bc  V a;cb  dxb dy c  Ra dcbV d dxb dy c . (3.33)

We make the reasonable statement that the difference between the two
a
directions will be a set of numbers R bcd that provide the proportionality
constants multiplying the area of the square and proportional to the size
of the vector being translated. This is again a statement of Stokes’
theorem and is a restatement of the arguments that led to Eq. (2.41), so
that the numbers R a bcd are the components of the orientation flux tensor.
The theorem from geometry is that a necessary and sufficient condition
for the space to be flat is that the curvature components vanish
everywhere.
For detailed calculations, we start with the components of the
curvature tensor given in terms of the orientation potentials, Eq. (3.27):

R a bcd   c a bd   d  a bc   a ce e bd   a de e bc . (3.34)

The reduced tensor is:

Rab  g cd Rcabd . (3.35)

It is formed by contracting two of the indices with the metric and so it is


by construction a tensor that is second order in the derivatives of the
metric. This matrix and the scalar R  g Rab formed by further
ab

contracting the tensor with the metric are the quantities that determine
the equations of motion from the action Eq. (3.29). The curvature tensor
R ab can be evaluated in terms of the orientation potentials. For the
special case of a holonomic coordinate basis, the orientation potentials
are determined by the metric using Eq. (2.33) and the reduced curvature
tensor is:
100 Geometry, Language and Strategy—Vol. 2

Rab   a b ln g  cab
c
 ab
c
c ln g  acd bdc . (3.36)

These components are second order derivatives of the metric and


generate a “wave nature” of the solution when sources are absent and are
related to the sources that are present by the principle of least action
Eq. (3.31).

3.7 Organizational Dynamics—Energy Momentum Stress


Tensor Tab

The decision process theory includes the active forces due to competition
and cooperation of players as well as constraint forces described by the
energy momentum tensor Tab . We will provisionally call the constraint
forces the inertial forces. They lead to inertia and mass. We have
suggested in Vol. 1 that such forces lead to organizational dynamics
with such forces behaving like a fluid in physical systems.
In the study of weather, the atmosphere is modeled as a fluid with
flow, energy density and pressure. One resultant behavior key to
atmospheric physics is that wind flows from high pressure to low
pressure. It means that energy flows away from the concentrations of
energy. In a purely fair system, we would expect the flow to be purely
isotropic: no particular direction is favored. When applied to
organizational dynamics, it is the condition that all parts of strategic
space are equally populated in the absence of the active competitive and
cooperative forces. Any concentration of decision making in one area
should dissipate. This is a principle of homogeneity. All the choices are
equal and if they are not equal they tend to revert to that average.
The organizing sources that we have in mind are many. Decisions
are not made in a vacuum but in a context. They are made within a
family, a community, a society, not to mention in a physical
environment. They must include both competition and cooperation. Any
combination of these might form the relevant system. Even if it were
possible to include all of these effects, most of the details generated
would not be of interest. We want to extract only the effects of the
constraints on the variables of interest. The effects are captured for the
Inertial Behaviors 101

organizational system by the organizing source, the energy-momentum


stress tensor Tab . We start by thinking of organizing sources as inertial
sources arising from the constraints, Sec. 1.6. We will need to develop
explicit models for these sources. It will turn out that the theory will
provide the form of the sources due to competition and cooperation.
Assume we have only the inertial sources. These organizing sources
generate the orientation flux fields, Eq. (3.31). The flux fields satisfy the
principle of least action that leads to an inter-relationship of inertial and
orientation flux fields, Eq. (3.31). This is a covariant expression, so is the
same in every frame. It relies on theorems from differential geometry
that the orientation flux tensor (curvature tensor) is in fact a tensor field.
The reduced tensors are also tensor fields. Specific models then depend
on studying possible forms for the energy momentum tensor Tab .
We start with the constraint forces that we derived for a general fluid,
Eq. (3.14):

Tab  VaVb  pab . (3.37)

We consider this expression in the co-moving coordinate basis that


moves with the fluid and introduce a unit vector field V a to describe the
flow of the fluid:

g abV aV b  1 . (3.38)

In the co-moving basis this determines the time component of the vector
field in terms of g00 . From our previous work, the stress components
p 00 and p 0b are zero. The time component of the energy momentum
tensor is the energy density T00  V0V0 .
If the basis is not only co-moving but Minkowski, the energy density
is equal to the time component of the stress tensor. In this same basis,
there are no momentum components and the space components of the
energy momentum tensor are given by p ab . Therefore we call the latter
the stress tensor components and in a general frame impose the
conditions:

pab  pba
. (3.39)
pabV b  0
102 Geometry, Language and Strategy—Vol. 2

We have defined a fairly general model for the inertial constraint forces.
The energy momentum is a symmetric tensor, which will in general have
real eigenvalues at each point, one of which will have a positive length
 ; the remaining will have negative values. The eigenvector
corresponding to the positive eigenvalue defines the vector field V a .
We can then look at the difference Tab  VaVb that we define as
 pab . It has no positive eigenvalue, but only negative eigenvalues that
are projected onto the space orthogonal to the flow by the projection
operator:

hab  g ab  VaVb . (3.40)

We recover the general form Eq. (3.37). We have a general form for the
inertial sources to be included in decision process theory. We define
pressure as the average stress, with n space components:

p  1 n h ab pab . (3.41)

We remain consistent with our unified principle of least action Eq. (3.31)
as long as our energy momentum tensor components are conserved
according to Eq. (3.32). We provide the result of the conservation laws:

DVb
h ab
 p ab 

 h a c hb d p cd ;b
. (3.42)
d
 V a ;a  Va ;b p ab  0
d

We have the inertial and orientation flux fields expressed in Eq. (3.42).
The right-hand side represents the inertial fields. The left-hand side
includes the orientation flux fields through the geometry, along with the
flow of energy. The gradient of the stresses provides a force; the second
equation is the matter conservation law for the inertial matter density  .
Let us return now to the ideas of organizational dynamics. In the
simplest case, we expect the inertial stress tensor to be isotropic and the
form to correspond to the elastic perfect fluid Eq. (3.14) written in
covariant form:
Inertial Behaviors 103

Tab     p VaVb  pgab . (3.43)

In this case, the stress tensor elements are equal to the pressure:

p ab  phab . (3.44)

The conservation laws simplify:

DV a
  p  h ab p;b
 . (3.45)
d
 V a;a    p   0
d

The first equation is Euler’s equation and the second is the conservation
of inertial media, defining the inertial media density  in terms of the
energy density and pressure:

d d
 . (3.46)
 p

This equation provides a differential equation for the inertial media


density in terms of the energy density. At various points in our study, we
return to these ideas.

3.8 Hamiltonian Formulation4

Our theoretical foundation relies heavily on the physical principle of


least action. We need to operate in a domain where that principle makes
sense. To explore this issue, we study the corresponding issue in physics.
As part of classical physics [Goldstein, 1959], Hamilton noted in 1834
that surfaces of constant action (Hamilton’s principle function) propagate
as if they were a wave. Since the Newtonian view articulated in
Hamilton’s language is the principle of least action, a surface of constant
action is a mathematical construct. Taking this construct, he argued that

4
This and the next section are advanced. Though their results set the context for our
theoretical foundation, they are not needed for engineering calculations of decisions.
104 Geometry, Language and Strategy—Vol. 2

it is possible to view classical mechanics as the geometric optics


approximation to a wave equation. The equation he obtained is very
close to Schrödinger’s equation of quantum mechanics. Moreover
Schrödinger’s equation describes a continuous system; it provides yet
another continuous system in which the energy momentum stress tensor
is conserved.
We start with some general background and summary of Hamilton’s
result. Using a slightly different but equivalent variation in which the
energy is conserved, the classical motion is determined by Hamilton’s
principle of least action written in terms of the kinetic energy T :

  Tdt  0 . (3.47)

The integral provides the historical definition of action5. Since the kinetic
energy is always a quadratic form in terms of the generalized
coordinates, T   m jk q j qk , this variation principle can be written in
terms of a “path length” ds 2   m jk dq j dqk in the geometry of the
active coordinates:

  2mT ds  0 . (3.48)

Hamilton’s surprising observation is that this form of least action looks


like the principle of least time in geometric optics.
Geometric optics is the field of study in which it is assumed that light
travels in straight lines like particles whose speed is determined by the
index of refraction n :

  nds  0 . (3.49)

The index of refraction is the inverse of the speed of light through a


media, so geometric optics is also formulated as stating light goes from
one point to the other in the least time. This simple idea explains why

5
Goldstein [1959, p. 231] notes that this principle of least action is associated with the
name of Maupertuis, though the objective formulation of the principle is due to Lagrange
and Euler.
Inertial Behaviors 105

light bends when it enters water, because the index of refraction of water
is different from that of air. It provides the theoretical foundations for the
construction of optical equipment. Hamilton’s principle of least action
has the same mathematical form as geometric optics with an index of
refraction 2mT determined by the kinetic energy. If there are no
external forces (other than possible constraint forces), then Eq. (3.48)
shows that the path will be a geodesic (extremum) in the space of
generalized coordinates. Equation (3.47) shows that the path will be the
one that takes the least time. In order to relate these ideas to action we
have to go deeper into the equations of motion.
Both Hamilton and Lagrange hoped to write equations of motion that
contained only the essential ingredients required for the solution. The
constraints were one thing that could be eliminated. The other thing to
identify was the set of variables that were constant in time: these are
associated with symmetries of the problem. Even though it might be true
that motion along a surface depends on the position, the force might not
depend on that position. The earth is a sphere to a high degree of
approximation. The gravitational force does not depend on the position
on the surface, just on the height. Hamilton’s approach to this problem
was to consider a function that he could identify with the total energy of
the system, and that depends only on the momentum and position. So for
example if there were only a single independent coordinate, Hamilton
introduced the “Hamiltonian” that is related to the Lagrangian (Sec. 1.6):

H  p, q, t   pq  L  q, q, t  . (3.50)

In conservative systems, the Hamiltonian H  p, q, t  is conserved and


equals the total energy of the system. The first order equations of motion
expressed in terms of the Hamiltonian determine the rate of change of the
momentum and the position respectively:

H H
p   , q  . (3.51)
q p

The Hamiltonian formulation makes clear the distinction of symmetry by


noting that if the Hamiltonian is translation invariant along q , then the
106 Geometry, Language and Strategy—Vol. 2

conjugate momentum p does not change in time. Here by translation


invariance we mean that if the value q is changed by a constant amount
to q  a , then the Hamiltonian is unchanged. Another way to express
this is to say that the momentum p is conserved. There is a similar
statement in the Lagrange formulation, since in that case the equations of
motion are:

L L
p  , p . (3.52)
q q

If the Hamiltonian is independent of the coordinate, then the Lagrangian


is also independent, because of Eq. (3.50).
The Hamiltonian formulation is symmetric with respect to coordinate
and momentum since if the Hamiltonian is independent of the
momentum, then the rate of change of the coordinate is zero. Momentum
and coordinate are truly conjugate variables. We show how this applies
to the observation about geometric optics. Hamilton hoped to reduce the
solution of every problem to one in which the Hamiltonian was
independent of both the momentum and the coordinate. The problem of
solving mechanics would then be reduced to transforming variables so
that the only unknowns would be the initial conditions. Of course there
would be some difficult (partial differential) equations to solve to get to
this point. Nevertheless the approach would yield insight into the
equations, and from the equations insight into the physical phenomena
they describe.
Hamilton defined (contact) transformations that generate new and
equivalent Hamiltonian functions that have no coordinate or momentum
dependence. He defined a contact transformation as a function S  q, P, t 
that transforms the original coordinate q  q  Q, P, t  and momentum
p  p  Q, P, t  to a new coordinate Q and momentum P :

S S S
p , Q , K H  . (3.53)
q P t
Inertial Behaviors 107

The new Hamiltonian K corresponding to the new coordinate and


momentum will satisfy the same type of relationships with the coordinate
and momentum as did the old relative to their Hamiltonian:

K K
P   , Q  . (3.54)
Q P

Setting the transformed Hamiltonian to zero (depends on no variables)


determines an equation for the contact transformation called the
Hamilton-Jacobi equation, and is a Hamiltonian in which all the
momenta and coordinates are conserved:

 S  S
K  Q, P, t   H  , q, t    0. (3.55)
 q  t

This contact transformation is the action S :

dS
 L. (3.56)
dt
The equation that determines the transformed Hamiltonian is a partial
differential equation for the action.
Ordinarily, we look only at the paths of least action, and not on the
value that action takes on other paths. What Hamilton discovered was
that if you looked at the action as a function determined from the above
partial differential equation, you would see that it describes a
(mathematical) wave front determined by surfaces of constant action. To
see this, the equation is separated into an energy term and a term defined
by Hamilton’s “characteristic” function W to determine the wave
velocity u of the front:

S  W  Et . (3.57)

In terms of the characteristic function, Eq. (3.55) is:

 W 
H  qk , E. (3.58)
 qk 
108 Geometry, Language and Strategy—Vol. 2

A surface of constant action is one in which changes in W in time must


be proportional to the energy E , or changes along the surface at the
same time must be proportional to the gradient  kW and be zero. If
there are changes in both time and distance, the changes in W must be
proportional to the path length ds orthogonal to the surface and
proportional to the spatial gradient  kW . These considerations
determine the wave front velocity u :

dW  Edt   kW ds
ds E E . (3.59)
u  
dt  kW 2mT

Hamilton’s variation principle Eq. (3.48) can be written in terms of the


wave velocity showing explicitly that the path taken is that given by the
least time:

ds
  0. (3.60)
u
The result is general, holding in the presence of external forces.
This deep dive into physics illustrates that the principle of least action
may be an approximation. The concept of geometric optics, which
Hamilton showed describes mechanics, is known to be an approximation.
We know that if the wavelength of light is large compared to the optical
device the wave approach must be taken into account to get an accurate
description. Hamilton noted the same thing; if Eq. (3.55) had a small
term on the right-hand side, then the classical solutions would deviate
from Newtonian mechanics, and the paths would no longer be paths of
least action in terms of the classical variables. The result is related to
decision process theory for decisions in that we must also view the
principle of least action as approximate. We alluded to this problem in
Sec. 2.1 in our discussion of units of time. We argued that we must
consider times large compared to the characteristics of the process in
order to insure that a decision was in fact made.
We suggest that the way to set units of time is to identify a quantum
of action that describes that decision process. Knowing the scale set by
Inertial Behaviors 109

the quantum of action would set the units of measuring time and would
augment the natural units defined in Sec. 2.1. In the next section we
indicate how quantum effects are introduced in physics, which suggests
how one might extend our theoretical foundation in the region in which
the principle of least action breaks down.

3.9 Quantum of Action6

To understand why least action breaks down, we need to understand


where Newtonian physics fails to describe experimental phenomena. We
also need to see that the deterministic principle still holds, as well as
understand the virtue of frequency versus probability. The first
indications of a breakdown were that Newtonian theory deviates from
experiment at atomic levels. At atomic levels, a good approximation to
observations was provided by Schrödinger, who provided a deterministic
continuum model for electrons as fields—a complex number   x, y, z, t 
called the wave function describes the electron field:

 2
i  k  k  V . (3.61)
t 2m
The most important aspect of this equation is Planck’s constant,  ,
which has dimensions of action and sets the scale distinguishing classical
behavior from quantum behavior. For decision process we suggest
introducing a similar constant.
The equation has other attributes worth noting. The potential energy
V specifies the forces acting on the electron. Typically it has the same
form as the Newtonian potential for a given type of force: Coulomb
attraction will be a potential whose strength is inversely proportional to
the distance; thus the gradient will give the inverse square law. The
electron is described by a continuous field called the wave function that
is a complex function of space and time and has a behavior that follows
our expectations for a continuous matter distribution. Recall that a

6
This and the previous section are advanced. Though their results set the context for our
theoretical foundation, they are not needed for engineering calculations of decisions.
110 Geometry, Language and Strategy—Vol. 2

complex number z  a  ib is expressed in terms of i  1 , whose


square is negative one, or equivalently z  re i in polar coordinates of
modulus r and phase  .
The position of the electron at any given time is provided by the
absolute value squared of the wave function. In the case described here,
the phase of the wave function is not observed. Some people think of this
as the probability that the electron will be at a particular position.
However, the square of the wave function is not a probability. It would
be more accurate to say that it represents the frequency with which you
would measure something at that point. The distinction is important since
the wave function here obeys a deterministic equation; it is really a wave
not a “particle”.
The predictions of this theory differ significantly from Newtonian
physics only when the action is small compared to Planck’s constant. We
demonstrate this by writing the complex number explicitly in terms of
the phase S and the modulus  that we identify with the matter
distribution of the electron:

 iS 
   exp   .
1
2
(3.62)


We show that the phase is the classical action in the limit that the
parameter  is small. We do this by separating out the real and
imaginary contributions of the resultant equation. The real part provides
the flow of matter and represents its conservation:


k  vk    0. (3.63)
t
This is precisely the same form we obtained for continuous matter
distributions Eq. (3.9). Even in the quantum domain, there is a concept of
a continuum distribution that obeys the laws of continuity. In making this
comparison, the velocity of the electron is defined in terms of the
momentum p k  m v k   k S . The second equation is determined by the
imaginary part:
Inertial Behaviors 111

S 1 2  12
  k S  k S   V   k k  2 .
1
(3.64)
t 2m 2m
In the limit that Planck’s constant goes to zero, this is the Hamilton
Jacobi’s Eq. (3.55) and demonstrates that the phase S is the classical
action. It also demonstrates that the paths will be those of least action,
whenever the left hand size is significantly greater than the right hand
size, which is proportional to  2 .
The action described by S corresponds to the classical “rigid” body
we might call an electron. The action of the wave however can be
associated with the electron viewed as a continuous field defined over
space, in the same way we have defined the electromagnetic field. We
can determine Schrödinger’s equation through a principle of least action
from the field perspective (Goldstein, 1959):

2   
S     k k  V  
8 m 
2
* *

2
 *  *  dxdydzdt . (3.65)

The wave function fields  and  are assumed here to be independent,


*

so the variation equation of least action yield the Schrödinger equation


and its complex conjugate. In particular, such equations will yield an
energy-momentum stress tensor created from the electron field, and so
energy and momentum will flow continuously from one part of space to
another. The modern view is to replace the above equation for a non-
relativistic field with that for a relativistic field. The current view of
physics is that matter consists of quanta of fields as opposed to rigid
bodies or particles.
The path integral approach [Feynman & Hibbs, 1965] provides an
alternate picture of how the transition is made from the quantum domain
to the classical domain in which the principle of least action is valid. The
path integral approach has been applied to both non-relativistic and
relativistic phenomena and so generalizes the Schrödinger approach. The
path integral approach provides a unified description of quantum
electrodynamics, a fairly complete description that unifies the
observational phenomena of quantum mechanics, special relativity and
classical mechanics. It provides a conceptual framework for relating the
112 Geometry, Language and Strategy—Vol. 2

classical action to the phase. We presume that this would be a fruitful


approach to use with decision process theory.
Feynman at the outset takes the view that rates are determined by the
absolute value of the amplitude (wave function) and that the amplitude is
constructed from the action in exactly the same way as the amplitude for
light assuming photons are particle-fields: Feynman thus goes back to
Newton’s original idea that light is composed of particles. Each path
contributes a complex number whose phase is the integral of the
Lagrangian times the time over the path. The wave function is the sum of
these complex numbers.
Since this description is abstract, we provide a simple example
[Feynman & Hibbs, 1965] of such an integral for the case of a free
particle that we can then compare to Schrödinger’s equation for a rigid
body (characterized by a mass m constrained so that the motion is
described by a single active variable q ) constrained to move in one
dimension, starting with an expression for the action S [ q ] :
tb

S  q    1 2 mq 2 dt . (3.66)
ta

Feynman’s prescription for computing the amplitude for such a


constrained particle is to consider all possible paths between the two
initial points in coordinate and time qa ta  and qb tb  and for each
path compute the Newtonian action:

i 
K  qbtb ; qa ta    exp  S  q  . (3.67)
all paths   

The expression depends only on the initial and final positions and times.
The sum can be carried out by considering a finite number of equally
spaced time intervals with the interval of time determined from the time
difference and the number of intervals t   tb  ta   N  1 , and then
taking the limit N   :
 12
 2 i  tb  ta    im  qb  qa 2 
K  qb tb ; qa ta     exp  . (3.68)
 m   2   tb  ta  
Inertial Behaviors 113

By construction this is for times tb later than ta : tb  t a . Because of


causality, the amplitude for propagating backward in time is zero. The
result obtained in this way is a solution to the Schrödinger equation.
 12
 2 i  t  ta    im  q  qa 2 
  q, t   K  qt ; qa ta     exp  . (3.69)
 m   2  t  ta  

Although we start with an idea of rigid bodies, we end with a continuum


description or field that describes the electron.
What is helpful about the path integral formulation is that the
evaluation of the path simplifies in the limit that the action is very much
larger than the Planck constant. The only paths that contribute are the
paths near the values where the action is maximal or minimal: the paths
of least action. The path integral approach is also insightful when there
are two or more classical paths, such as the above problem where the
motion is on a ring, Cf. the review in Vol. 1.
The important thing we learn from this analysis is that there are good
reasons to use the principle of least action and equally good reasons to
take that principle as an approximation. As an approximation, we expect
any theory to contain a parameter such as  with dimensions of action.
For decision theory, the dimensions would be UT  . What distinguishes
our approach from physics is that each decision process has a theory
of its own and so has its own quantum of action set by its own  .
We implicitly acknowledge this by using different units of time
characterizing any given type of decision process. Some processes are
mere moments, some minutes, days, months or years. What we require in
order to apply decision process theory is that we work in the geometric
optics limit where actions are much bigger than  .

3.10 Outcomes

We acknowledge that we are not the first to suggest using mechanical


models for decisions. We have gained insight from Von Neumann &
Morgenstern [1944], though there were others before them. The principle
of the greatest good for the greatest number was the basis for Bentham
114 Geometry, Language and Strategy—Vol. 2

utilitarianism [Bentham, 1829]. Certainly a principle of greatest


happiness resonates with the physical idea of least action: in both cases
one imagines an extremum. A physical theory along these lines was
proposed by7 Edgeworth [1881] equating the greatest happiness idea to a
principle of maximum energy. These approaches lack mechanisms to
account for unjust behaviors such as the institution of slavery. They are
not set up to allow offsetting effects and so are not complete. It is
required of many students in engineering to take a course in ethics in
which such issues are discussed in great detail, see for example Tavani
[2011].
The elements missing from economics theories based on principles
termed egoistic by their authors are off-setting mechanisms that enforce
justice. This is not however, an appeal for a theoretical proof of justice;
only a call that a theoretical foundation must allow for such ideas. It is
our claim that we extend the more modern discussion of utility by Von
Neumann & Morgenstern [1944] in a way that allows for such new ideas.
There are plenty of mechanisms we can think of that we might want to
explore. We return to these issues in Chap. 11.
In general terms, in this chapter we have updated the decision process
theory so that it contains not only non-inertial forces but inertial forces.
We have used the principle of least action to provide a self-consistent
view of these effects. Our goal to present a comprehensive and robust
framework forced us to include non-inertial forces effects including
rotating frames. Decisions involve complex strategies that are executed
by agents or as they are called in game theory, players. These players
make choices that in fact may change over time. We have been led to a
decision process theory in which the strategic choices are represented by
active dimensions of space, the players provide one or more inactive
strategies each of which is hidden but manifests as a payoff field
associated with that player. The principle of least action Eq. (3.31)
replaces Eq. (1.13), which is a mechanistic device to identify the
equilibrium strategies. By providing different working models for the
energy momentum tensor Eq. (3.37), we provide a common and
scientific framework to expand our understanding of the decision making

7
I am indebted to Mr. K. Kane for bringing this essay to my attention.
Inertial Behaviors 115

process. In the next chapter we will focus on the consequences of


players, their isometries and persistent behaviors in this theory.
As in the previous chapter, the attainment of the outcomes in this
chapter will be facilitated by doing the exercises in the following section.
We list here more detailed outcomes expected by section.
 From Secs. 3.1-3.2, the student should understand the need for inertial
effects in decision processes and gain an understanding of the form
such effects have in a physical based theory.
 From Secs. 3.3-3.4, the student will know how to incorporate both
inertial media effects as well as non-inertial effects in a unified
manner. The prime example of non-inertial effects is that due to the
payoff fields of game theory.
 From Sec. 3.5 the student learns how the unification of payoffs and
inertia is accomplished with the principle of least action.
 From Sec. 3.6, the student learns the final unification that leads to a
quantitative decision process theory, Eq. (3.31). It is motivated by the
need to include orientation flux fields in the theory in a unified
manner. Measuring distances or time in decision processes is
dependent on the frame of reference. The changes are captured in the
orientation flux fields and related metric fields. Such changes are
necessary consequences of the common sense notions that values and
time have local meaning, whereas comparing values and time at
different points of space and time require some continuous
mechanism; that mechanism carries both energy and momentum.
 From Sec. 3.7, the student should be able to use a working model for
the energy momentum tensor of the organizational system media. The
student should understand the dynamic equations that result from the
conservation laws.
 From Secs. 3.8 and 3.9, the student should have an appreciation that
the principle of least action need not be an exact principle. It may fail
when the value of the action becomes too small. This is an area for
future research, both in terms of observation and theoretical
development. It has proven to be an area of great interest in the
physical sciences.
116 Geometry, Language and Strategy—Vol. 2

3.11 Exercises

(1) Give examples of inertial effects from the physical world.


(2) Give examples of inertial effects in economics.
(3) Show that the principle of least action Eq. (3.2) leads to the
mechanics equations of Lagrange, Eq. (1.9).
(4) Show that the application of the principle of least action to Eq.
(3.17) yields Maxwell’s Eq. (1.24).
(5) Show that the energy and the momentum components of the
electromagnetic field are given by Eq. (3.18).
(6) Show that the conservation of the energy momentum tensor for the
electromagnetic fields, Eq. (3.19) also leads to Maxwell’s Eq.
(1.24). What needs to be added to the energy momentum tensor to
include the matter contributions?
(7) With the help of [Gockeler & Schucker, 1987], show that the
principle of least action yields Eq. (3.31).
(8) Derive the conservation laws Eq. (3.45) from T ;b  0 .
ab

(9) In a coordinate basis Eq. (2.33), use Eq. (3.34) to demonstrate the
following symmetry relationships for the curvature tensor
components, which by covariance must hold in any basis:

Rabcd   Rbacd   Rabdc  Rcdab . (3.70)

(10) In a coordinate basis Eq. (2.33), use Eq. (3.34) to demonstrate the
following symmetry relationship for the curvature tensor
components, which by covariance must hold in any basis:

R d abc  R d bca  R d cab  0 . (3.71)

(11) The symmetry relationships can be derived without recourse to the


coordinate basis, but from more general principles. Using only the
first two (anti) symmetry relations in Eq. (3.70),

Rabcd  Rbacd   Rabdc


Inertial Behaviors 117

and the cyclic symmetry condition Eq. (3.71), which follows from
the second Bianchi identity [Gockeler & Schucker, 1987] in a
general coordinate system, derive the symmetry relation:

Rabcd  Rcdab . (3.72)

(12) Using the full set of symmetries Eq. (3.70), show that we don’t get
the cyclic symmetry Eq. (3.71) but the cyclic tensor below is totally
antisymmetric in all indices:

X abcd  Rabcd  Racdb  Radbc . (3.73)

(13) Use the results of the previous exercises to show that the total
number of independent curvature components in a space-time of D
dimensions is

1
2  1
2 D  D  1    1 2 D  D  1  1

because of the symmetries Eq. (3.70), less

D!
4! D  4 !

because of the remaining constraint of setting each independent


tensor Eq. (3.73) to zero to satisfy the cyclic symmetry constraint
Eq. (3.71). Show that the resultant number of independent curvature
components is
1
12 D 2  D 2  1 .

(14) Using the average pressure Eq. (3.41) demonstrate the following
form for the contracted curvature tensor, where n is the number of
spatial dimensions:

  p p
1
 Rab  VaVb      p     pab  hab p  hab . (3.74)
 n 1  n 1
118 Geometry, Language and Strategy—Vol. 2

(15) Derive the conservation laws Eq. (3.45) from the form of the energy
momentum tensor Eq. (3.43).
(16) Show that the conservation of the energy momentum tensor Eq.
(3.32) is a direct consequence of the conservation of the Einstein
tensor G ab  Rab  1 2 Rg ab , which follows from the definitions of the
curvature tensor and its contractions:

g bcGab;c  0 . (3.75)

(17) In Sec. 6.3, we will provide arguments for adopting the conductivity
model, in which there is a charge current proportional to the electric
field in a frame in which the medium is at rest. This can be modeled
in a covariant way by adding a viscosity contribution to the energy
momentum tensor. With the following form, show that one obtains
the conductivity model, where the viscosity  is the conductivity:

T  VV  ph   


DV
V ;       V
 . (3.76)
      
  h          1 n  h

In addition, using Eq. (3.42), show that the longitudinal and


transverse conservation laws are respectively (Cf. Eqs. (A.9) and
(A.10), Vol. 1):

V  ;     p       ,
DV
   p  1
2          

 p; h   n2n2 ; h   ;    2 h V ; h
1  0. (3.77)

(The second equation corrects an error in Vol. 1.) Compute the


components of these equations for the active and inactive strategies.
Chapter 4

Persistent Behaviors

In Chaps. 1-3, we developed a consistent theory of decisions built from


lessons learned from game theory and the physical world. It is a
mathematical and engineering language for a decision process theory.
Our proposed theory is Eq. (3.31), extended to include time and both
active and inactive strategies:

R  1 2 g  R  T . (4.1)

It is not derived from game theory, physics, biology or chemistry. The


theory is based on geometry in the space-time of strategic choices, not
physical space-time. It is a hypothesis as well as a framework that must
be tested by observations.
The concepts created in the last three chapters have consequences that
are not applications of existing physical theories, though we may use
their mathematics to gain insight on how to better our understanding. We
are applying established scientific principles to a new physical domain,
which may well generate new insights in this domain as well as to
existing domains.
The rest of the book is devoted to exploring the computational
consequences of this theory. In the process, we extend Vol. 1. In this
chapter, we focus on the persistent behavior that is distinctive to
decision processes, which leads to the notion of players or agents. For
the mathematically oriented reader, we note that persistent behavior as
used here corresponds to isometry transformations that leave the value of
the metric-fields invariant at every point. Each isometry requires the
existence of a vector field with special properties. The collection of all
isometries forms a mathematical Lie algebra, which characterizes the

119
120 Geometry, Language and Strategy—Vol. 2

internal symmetry of the theory. Because of this, a player’s persistent


behavior defines a well-defined and persistent mathematical structure.
Persistent behaviors arise generally. They are associated with any
isometry. In Vol. 1 we considered the possibility of the isometry of time.
This isometry led us to introduce the centrally co-moving hypothesis that
asserts that the active1 vector field components K a  V a corresponding
to the time isometry are proportional to the energy flow; the inactive
strategy components are zero K j  0 . We consider that isometry here as
well as additional isometries that we identify with persistence of player
strategies that represent social agreements or codes of conduct.
Though restrictive, these additional isometries still allow us to find
new behaviors; in particular we hope to find behaviors that correspond
not only to the resonant behaviors of Vol. 1, but behaviors that reflect
transient and steady state behaviors that oscillate with any frequency
depending on the inertial forces. We numerically compute complete
solutions to Eq. (4.1). In Vol. 1 we were able to provide only partial
solutions.
We explore model consequences of the theory by examining the
theory in several frames of reference or bases, Sec. 4.1. We look at the
theory in Sec. 4.2 in the normal-form coordinate basis. In this basis the
persistent behaviors of players are characterized by the property that all
metric and orientation potentials are independent of the set of internal
player dimensions. The consequences of persistency in this basis
manifest in the field equations, which are examined in Sec. 4.3 in a
specific gauge, the harmonic gauge. We look at the covariant expression
of persistency and define isometry in Secs. 4.4-4.5.
We gain additional insight into the consequences of persistency by
looking at the behaviors in a co-moving coordinate basis, Sec. 4.6,
which gives us a potentially simpler numerical approach to the field
equations that we follow in later chapters. Our initial focus will be on an
orthonormal co-moving coordinate basis, Sec. 4.7. A focus in later
chapters will be on a holonomic co-moving coordinate basis; in this
chapter we define what we mean by holonomic. In Sec. 4.8 we introduce
the player fixed frame model that provides a class of models, Sec. 4.9

1
By active we mean both time and the active strategies.
Persistent Behaviors 121

that highlights the vorticity aspects of the theory and may be solved using
known current numerical techniques. It satisfies the centrally co-moving
hypothesis.
We use such numerical results in later chapters to illustrate how
decision process theory provides a common ground for discussion of
economic issues. The model allows us to extend, using an electrical
engineering analogy, behaviors that are like DC circuits, to behaviors
that are like AC circuits.

4.1 Coordinate and Non-Coordinate Bases and Potentials

In the physical world, we take the ability to measure distances and


intervals of time for granted. This assumption is fundamental to our
approach to the physical sciences. We assume that we can determine a
physical location and a time for each element of a physical event.
Because of that, we can assign a distance between physical events. For
these and other reasons, we feel confident that we can describe the
dynamic behaviors of physical processes. We can describe the events as
they evolve in time.
In this book we assert that we have the same ability with decision
processes. Though we have not done that here, we suggest that this
assumption is one that needs to be established empirically; we have
looked at some data and indeed do observe that it appears to be
provisionally true.
We adopt the continuity viewpoint from the physical sciences that for
each decision event, at each moment of time, there is a local region in
which each strategic choice has a value that varies only slightly from the
event in question and that furthermore, the values change only slightly at
nearby times. We make this mathematically precise by asserting that
there always exists a coordinate basis at each point of strategic space
time: there is a coordinate value (for time and each strategic choice) that
has the same value not only at that point but on an extended surface that
contains that point. A complete set of independent surfaces through the
point then provides a coordinate basis.
122 Geometry, Language and Strategy—Vol. 2

The coordinate value at a point corresponds to a surface through that


point. A complete set of coordinates of this type are called holonomic
[Cf. Sec. 2.7]2: the coordinates can each be derived from a potential and
the gradients along each coordinate direction, the coordinate vectors   ,
mutually commute,        . Such holonomic coordinates
implement our belief of how distances are measured between decision
process events in time. Thus we use coordinates here in the same way
electrical engineers use the word potential fields.
Prior to Einstein, our common sense notion was that we could get by
with a single and global coordinate basis: a universal view from which to
view interactions. This common sense notion we can think of as
Newtonian: it corresponds to a frame of reference in which space-time is
essentially flat. Einstein suggested that this view is too simplistic.
Consider a well-known object, the sphere. A map maker takes a patch of
a sphere and maps it onto a sheet of paper, the chart. This mapping
creates the coordinates (local latitude and longitude) that are locally
holonomic. However, as is well known, the sphere can’t be mapped onto
a single chart; at least one point will be left out. More than one chart is
needed because of the left-out points. Our common sense view must be
extended to describe even such simple and well-known surfaces.
Given the need for an expanded common sense about the geometry of
spaces and the fact that we know little about the geometry of the space-
time for strategic choices, it seems prudent to allow the possibility for
multiple charts. Our use of coordinate bases will therefore be local, not
global. Multiple charts will in general be needed to describe even rather
simple structures and thus we allow for multiple charts in decision
process theory.
The addition of multiple charts makes more complex the discussion
of which spaces are the same or different. We use the language from the

2
The idea in physics is that simple systems are integrable. More complex systems are
those with constraints, but if the constraint equations are each integrable, the remaining
variables will also be integrable. For decisions we envision that the active and inactive
variables are those that remain after all the constraints have been imposed. The
assumption is that the remaining variables will be integrable and correspond to variables
that are exact as defined in the following paragraphs.
Persistent Behaviors 123

literature, e.g. [Hawking & Ellis, 1973], to say that the spaces are
manifolds, which means we include in the definition all of the local
properties and charts, as well as a prescription for how the charts relate to
each other when they overlap. We then say that two manifolds are the
same if there is a mapping that takes one into the other and vice versa in
such a way that the manifold properties are the same: they have the same
topology and geometric structure. Such mappings are called
diffeomorphisms. Our linear transformations may or may not be
diffeomorphisms. Thus different coordinate bases may not represent
equivalent manifolds. Nevertheless it is useful to use general linear
transformations to explore different models. Often, different models have
a common embedded manifold that can be identified by a suitable linear
transformation. We find that to be the case with the co-moving
orthonormal coordinate basis defined below, Sec. 4.6.
For coordinates that are locally holonomic, we expect certain
mathematical properties. Each constant surface that corresponds to a
holonomic coordinate has the attributes of a potential field. Potential
fields generate a unique vector at each point on the surface that is normal
to the surface. In general, parallel surfaces generate the same vector
fields. To gain more insight in potential fields, you may recall potentials
 whose gradient  a  determines the properties of the vector field that
is normal to the surface of constant potential. Surfaces of constant
potential are used in engineering and physics. For example the potential
surfaces within a capacitor of arbitrary shape determine the electric field.
The gravitational potential in Newtonian mechanics defines surfaces that
map the gravitational forces. The potentials provide a convenient method
to discuss physical effects associated with vector fields. The vector fields
are typically those that more directly determine dynamics, not the
potentials. Moreover, not every vector field that determines physical
behaviors can be written as the gradient of a potential.
Although locally holonomic coordinate bases provide an intuitive
meaning for measurements, there are also advantages of looking at
global coordinate systems that are not holonomic. They may better
illuminate curvature effects. As a simple example, recall that we live on a
rotating sphere whose holonomic coordinate system would be fixed.
However, a rotating coordinate system has advantages because it
124 Geometry, Language and Strategy—Vol. 2

corresponds to the world we actually see. In many ways the rotation of


the earth is hidden from us. We must then capture the effects of our
hidden rotation as two separate effects: Coriolis Effect and centripetal
acceleration.
Similarly, we suggest that the concept of persistency that underlies
the existence of players and agents is described on the one hand by the
locally holonomic hidden dimensions of Sec. 2.11 and on the other hand
is described by a fixed non-holonomic frame. We illuminate some of
these persistent attributes in the normal-form coordinate basis, which
we introduced in Sec. 2.11. Using the frame independent principle of
least action, which provides decision process theory Eq. (4.1), dynamics
can be described by the ways in which these persistent attributes interact.
To include frames that don’t form a holonomic coordinate basis, we
refine our mathematical distinctions. We start with the basic idea that in
a holonomic coordinate basis, each vector field can be derived from a
potential field. We recall that a vector field U when thought of as a 1-
form, Eq. (2.24) discussed in Sec. 2.8, has a potential only if it is exact,
dU  0 . In a holonomic frame, this reduces to the requirement that
 U    U , which is the necessary and sufficient condition for there to
exist a potential  such that U     . A 1-form is exact when we can
write it in terms of this potential: U  d  .
We write these as covariant requirements. Starting from a holonomic
set of frames dx , we construct a new set of (potentially non-holonomic)
frames E  E dx  using the gauge transformation E  . An exact 1-
form d U  0 , will be transformed to coordinates:

U  E       . (4.2)

The requirement  U   U   U ;  U  ; in the old frame, takes the


covariant form in the new frame:

U ;  U  ; . (4.3)

Though the components are determined by the potential using a


differential operator, these differential operators in general don’t
commute,        .
Persistent Behaviors 125

A second important distinction for holonomic coordinate systems is


the commutativity of the differential operators. In differential geometry,
any vector field X  can be associated with a differential operator
 X  X    . The commutation between two such operators is defined to
be a new covariant operator, called the Lie product of the vectors:

 X  Y   Y  X   X  Y  ;  Y  X  ;      X, Y    X ,Y . (4.4)


The covariant requirement that a frame be holonomic is that the 1-forms


are all exact and the Lie product of any pair of frame vectors is zero.
Conversely, for frames that are not holonomic we expect at least one of
these properties to be absent.
The advantage of the covariant expressions is that we can investigate
specific attributes of a holonomic frame that are present or missing in
other bases, such as the normal-form coordinate basis. So for example
using the definition of exactness, Eq. (4.2), with Eq. (2.48), which
provides the orientation potentials used in the covariant derivatives, the
assumption that the transformations are functions only of the active
strategies leads to the determination that the active coordinates remain
exact (as they are not transformed) but the (transformed) inactive
coordinates need not be exact. Further, it is not difficult to show (see
Exs. 1-3 at the end of the chapter) that the Lie products of the
transformed inactive strategies are zero and that the Lie products of the
transformed active strategies are effectively zero.

4.2 Normal-Form Coordinate Basis—Hidden Symmetries

You might think that the most convenient frame of reference would
always be a coordinate or holonomic basis. Such bases lead to
differential equations that have been well-studied and provide necessary
information about existence and properties of solutions. However, such
local bases don’t necessarily provide the best global view, especially if
there are symmetries. For example, in decision process theory,
persistency determines the property of the distance measure, Eq. (2.42):
126 Geometry, Language and Strategy—Vol. 2

ds 2   jk  d j  Aaj dx a  d k  Abk dxb   g ab dx a dx b . (4.5)

The symmetry is that the metric elements are independent of the


inactive coordinate vectors  j . We say that these coordinates are
hidden. If the expression is expanded, the most general form of the
distance measure is recovered without changing this “hidden” property.
Writing the measure of distance as Eq. (4.5) emphasizes that many
equivalent choices of variables or gauges that transform only the hidden
coordinates will lead to the same end result. Distances are made up of
two orthogonal contributions: the first an inactive or internal contribution
and the second an active contribution. All results in decision process
theory, such as the above measure of distance, are independent of the
choice of gauge. In particular, a subset of these gauge transformations
that leaves the inactive variables hidden must leave all results of the
theory unchanged. We define these transformations in more detail.
The distance measure is expressed in terms of locally holonomic
 
coordinates x    j x a that represent the inactive and active
strategies respectively (with time treated here as active). These
coordinates are not in general orthogonal. The inactive coordinates  j
 
are hidden in the sense that the tensors  jk A j a g ab are independent
of these strategies. We can verify that the following gauge
transformation  j   j   j , A j a  A j a   a  j leaves the inactive
strategies hidden. For example, the distance Eq. (4.5) does not change
under these transformations as long as the inactive and active metric
elements  jk gab don’t change (are gauge invariant). Calculations done
in the coordinate basis typically involves a significant amount of extra
algebra because many of the intermediate results are gauge dependent.
For example, these intermediate results may depend explicitly on the
vector potential Aaj as opposed to the gauge independent payoff field
F j ab , the active metric g ab or the inactive metric  jk .
Gauge independence for decision process theory follows because the
principle of least action is frame independent. This covers frames that
are locally holonomic or frames that are not. Therefore without losing
any generality, by picking a non-holonomic frame, it may be possible to
reduce significantly the algebra necessary to obtain the desired gauge
Persistent Behaviors 127

independent results. We illustrate this by looking at our theoretical


framework in the normal-form coordinate basis using the gauge
transformation formalism from Sec. 2.8 and Eq. (2.43) in Sec. 2.11:

U a  dx a
. (4.6)
U j  d j  Aaj dx a

This basis, which is suggested by the invariant distance Eq. (4.5)


provides a natural gauge invariant and orthogonal split into active
coordinates, which are holonomic (on the active subspace only, see Ex.
(3)) and inactive coordinates that are gauge invariant but not holonomic.
We adopt the following conventions. Unless specifically noted, we
label the active exact dimensions, which are time and the active
strategies, by the indices at the beginning of the alphabet a, b, c, . We
label the inactive strategies, which are not exact in this basis, by the
indices in the alphabet i, j, k , . We use the following notations for
the determinant of the inactive metric components  jk and active metric
components g ab :

  det  jk
. (4.7)
g  det g ab

Noting that these two sets are orthogonal to each other, the determinant
of the full metric for the space is the product  g , which can be either
positive or negative. We indicate the absolute value of a quantity in the
usual way, for example  is the absolute value or magnitude of the
inactive determinant.
In the normal-form coordinate basis, the payoff matrix is determined
by the differential of Eq. (4.6) that is manifestly covariant:

dU j  F j  1 2 F j ab U a  Ub
. (4.8)
F j ab  Abj;a  Aaj;b

In the normal-form coordinate basis, this expression is gauge invariant:

F j ab   a Abj  b Aaj . (4.9)


128 Geometry, Language and Strategy—Vol. 2

The expression Eq. (4.8) is derived from Eq. (2.30).


We thus have a basis in which the calculations of gauge invariant
results are obtained directly, albeit in a frame that is non-holonomic. The
degree to which the basis is non-holonomic is specified by the payoff
field. As a notational aside, we make frequent use of raising and
lowering the indices using the metric, so for example we can write:

Fj ab   jk g ac g bd Fcdk . (4.10)

This highlights that there are no non-zero metric components that mix
active and inactive strategies in the normal-form coordinate basis. As a
further physical aside, we note that the Coriolis Effect and centripetal
acceleration on the earth in which the inactive coordinate is time,
respectively would be the “payoff” tensor for the Coriolis Effect and the
gradient of the inactive metric for the centripetal acceleration, identified
as gravitation.

4.3 Normal-Form Coordinate Basis—Harmonic Gauge

Because the principle of least action remains unchanged under local


linear transformations, specifically gauge transformations that are
diffeomorphisms, the resultant field equations have a corresponding
number of degrees of freedom that are larger than the physically
distinguishable number: many choices of frames solve the same set of
equations. By going to the normal-form coordinate basis, we have
reduced, though not entirely eliminated those degrees of freedom. What
are physically meaningful are the set of coordinate vectors and the
gauge-choice that removes all unfixed degrees of freedom. This freedom
of choice is not without precedent in physics: in electrical engineering
for example, there are many choices of potentials Eq. (1.21) that lead to
the same electric and magnetic fields. In decision process theory, we are
also free to make such gauge choices, which lead to identical outcomes
for decision processes.
What is unusual about this is that it goes against the common
Newtonian notion that there is a unique global space and time not subject
to arbitrary mathematical gauge choices. The Newtonian view is
Persistent Behaviors 129

appropriate to flat space not curved space. For this reason, we go against
this view and allow gauge arbitrariness in the choice of pure strategies
and time, which allows for dynamics that reflect a topological and
geometric structure that will be constrained by its success in matching
observational data. The gauge freedom is analogous to the freedom of
using “charts” for navigation: each chart treats the earth as being flat or
Euclidean. Because they describe an object that is round, they don’t
exactly match up. The mapping that matches them is the gauge
transformation. The gauge freedom is that at any point, we are free but
not required to consider the space (our chart) to be flat. Though there is
gauge freedom, it is not a freedom that can be removed as an attribute of
the theory: topological objects with curvature such as the sphere require
multiple charts. We assert that decision processes generate curvature and
require theories with this gauge property.
To solve the local field equations, we can pick any gauge. The
solution will be valid in some region around our initial conditions. Our
choice is the harmonic gauge described in the literature, [Wald, 1984].
We adapt an argument from that literature to apply harmonic
coordinates to the normal-form coordinate basis. We start with the field
Eq. (4.1) for decision process theory that is the consequence of the
principle of least action. We write these equations in the normal-form
coordinate basis. The argument assumes we have a set of coordinates,
x a , which are scalar functions of y a in another basis in which the metric
g ab  y  is known in terms of these coordinates y a .
We define the active coordinates x a in the normal-form coordinate
basis using the following differential equation, where the Greek indices
 ,  span both the active and inactive dimensions of the normal-form
coordinates:

g  x a ;  0 . (4.11)

This differential equation can be written in any frame, so in particular


can be written in the normal-form coordinate basis with metric g ab . The
differential equations can be expanded using Eq. (2.48) and written in
terms of the determinants Eq. (4.7) and the active partial derivatives:
130 Geometry, Language and Strategy—Vol. 2

1
g
b  
g g bc  c x a  0 . (4.12)

These differential equations and suitable initial conditions have a unique


solution, which provides a solution for the normal-form coordinates.
In the normal-form coordinate basis, the differential equation can
also be written in a form that puts constraints on the number of
independent metric potentials g ab :

1
g
b  
g g bc  c x a  0 
1
g
b  
g g ba  0 . (4.13)

The coordinates x a that satisfy this condition are the harmonic


coordinates and Eq. (4.13) is the harmonic gauge condition. Without
excess degrees of freedom, these conditions, the initial conditions and the
field equations completely specify the solution to the field equations,
which we turn to next.

4.4 Normal-Form Coordinate Basis—Field Equations

In this section we develop the field equations with sufficient detail so


that the interested student can verify the results. The important steps
needed to obtain the results are specified as exercises at the end of this
chapter, so only the approach, attributes and interpretation will be given
in this section.
The field equations in the normal-form coordinate basis are obtained
directly from the curvature tensor, Ex. 6, Eq. (4.77), Ex. 7, Eq. (4.78) and
Ex. 8, Eq. (4.79). Because we operate in the normal-form coordinate
basis, the gauge invariant results, relative to the hidden dimensions, are
obtained directly. Given the full curvature tensor components, the fields
Eq. (4.1) are obtained by contracting the curvature tensor with the metric.
The number of equations depends on the space dimension n , which is
the total number of active and inactive dimensions. The resultant
equations are of three basic types, depending on whether the indices are
all active, all inactive or mixed. Technically the field equations are
Persistent Behaviors 131

restricted to fields that are functions only of the active variables, so on


this subspace the active variables x a are holonomic and ordinary rules of
calculus apply. With the choice of harmonic gauge these equations along
with suitable boundary conditions have unique solutions. We now
discuss the equations that result for these three cases.
We start with the field equations for the payoff tensor, which result
from the full set of field equations with mixed indices:

1
2 g bd b  g ac jk F k cd   Tj a . (4.14)

This is the covariant form in harmonic coordinates, and generalizes


Maxwell’s Eq. (1.24). There is one such equation for each player or
agent in the decision process. There is a source current for each player
Tj a , which is part of the organizational system, Sec. 3.7. There are
contributing sources as well from strategy and time behaviors of the
active and inactive metrics: such behaviors are frame rotation effects that
may be consequences of the particular contextual frame. These effects
show up more clearly if we solve the equations using vector potentials.
In this case the vector potentials and their gauge dependence provide a
useful way to simplify the equations and interpret the results.
We use the potential form Eq. (4.8) along with the following gauge
conditions for the potentials, which are a type of harmonic gauge applied
to the inactive space, Vol. 1:

g ab  b Aaj  0 . (4.15)

Using these gauge choices, we write the generalized Maxwell’s


equations for the vector potentials for each player:
1
2 g bd  b d Aaj  1 2  a g bd  b Adj
 1 2  b g ae g ec g bd F j cd  1 2 g bd  jl  b lk F k ad   Taj . (4.16)

The similarities and differences to the physical equation for


electrodynamics can be expanded upon. The similarity is that for each
agent, we have a wave equation, so we expect signals to propagate
independent of any matter medium with a common velocity. Each wave
132 Geometry, Language and Strategy—Vol. 2

equation generalizes the wave equation for electromagnetic fields Eq.


(1.16).
The substantial differences are as follows. There are multiple wave
equations: there is one equation for each player. Each player feels the
conserved inertial currents T j a , which are similar for each player in
simple models, such as the perfect fluid model, Eq. (3.37) and therefore
the vector potential solutions for each player will be forced to be similar.
As a special case, this may generate the Nash equilibrium in game
theory, which postulates that equilibrium exists when all players
subscribe to the same rules. More generally, we see how these same rules
enforce common behaviors.
However we note that the idea of current may be more general than
what is suggested by the simple models. As in physical models, it might
be that the currents are proportional to the applied “electric” field. Or, as
we shall discover in our numerical computations, there may be other
model choices that are equally compelling. We thus leave open the exact
form of these inertial currents.
Another set of differences is the existence of “currents” that result not
from the organizational system, but purely from frame rotation effects,
gradients of the active or inactive metric. We see a contribution
corresponding to each. There is mixing between players, a phenomenon
that is special to this theory of decisions. The last term of Eq. (4.16) on
the left-hand side mixes the roles of the players. In order for this to be
possible, the scalar fields must not be constant. The scalar fields mix
internal and active geometry effects.
The field equations that correspond to the active space depend on the
persistent attributes of the players:

Rab  1 2 gab R   Tab   Tabpayoff   Tabinactive . (4.17)

We see three distinct contributions to the “bar” or active curvature


components on the left-hand side of the equation that are computed in the
active subspace. In other words, the reduced tensor Rab is given directly
by Eq. (3.36), which includes partial derivatives through the second
order in the active metric. Its form will be simplified further by the
Persistent Behaviors 133

choice of the harmonic gauge. Though the resultant equations are


difficult to solve, we can say a few things based on general principles.
The first contribution on the right-hand side of Eq. (4.17) is from the
organizational field source  Tab . As in physical theories, for a time
isometry (defined in the next section) and very weak couplings  , the
metric is well approximated by the Minkowski metric with the exception
of the time component of the metric g00  1   . The deviation from
unity is a static (time independent) field that is determined by the matter
density:

2    . (4.18)

This is Poisson’s equation and reflects that matter attracts other matter.
In the dimensions of 3+1 space-time, this is an inverse square attraction.
In a strategic space-time with n spatial dimensions, this force law is
generalized.
More generally, we can see that if time is an isometry as defined in
the next section, there will be two effects corresponding to the Coriolis
Effect and centripetal acceleration. In general relativity the “centripetal
acceleration” is determined by the time component of the metric and the
corresponding “Coriolis Effect” is gravitomagnetism [Ryder, 2009, p.
180ff] and can be measured with gyroscopes and clocks by means of
satellites. It is interesting that there is a connection between payoffs,
magnetism and Coriolis. The source of the connection is the underlying
symmetry or hidden variable. To make the connection complete for
electromagnetism, we use [Kaluza, 1921] and [Klein, 1956] for the
Kaluza-Klein theory and the analogy to their proposed hidden fifth
dimension.
It is worth emphasizing that these effects do not arise from a physics
analogy but from the mathematical structure of the proposed decision
process theory and general principles of differential geometry. They all
share the same underlying mathematics. We find it significant that the
gradients of the orientation potentials provide a basis for an attraction
that exists independently of the payoffs. In whatever frame we work, an
important effect is that of acceleration. This is the symptom of the force
134 Geometry, Language and Strategy—Vol. 2

of attraction. It is also the cause, at the global level, of the non-existence


of a global frame that is simultaneously holonomic.
The matter density is not the only contribution to the energy and
momentum of the active space, Eq. (4.17). There is also the effect from
the payoff fields, part of the active forces:

2 Tabpayoff   jk Facj Fbdk g cd  1 4 g ab Fk cd Fcdk . (4.19)

To physicists and electrical engineers, this is clearly an analogous form


of the energy momentum fields of Maxwell, Eq. (3.18) applied to the
case of multiple payoff fields corresponding to the players involved in a
decision. As pointed out by Einstein, light has inertia. What we say in
decision process theory is that payoffs have inertia.
Again, these arise not from analogy but from the decision process
theory. These are specific consequences of the existence of persistent
behaviors. Einstein was the first to suggest that such radiation fields
carry real inertial effects just like matter. He predicted that light would
bend as it passes close to a star as a consequence of this inertial effect.
Here also there will be real inertial effects that follow from the presence
of the payoff fields. Qualitatively, it says that transactions with large
payoff matrix elements carry more energy and will have more impact on
dynamics than transactions with small payoff matrix elements. This is
different from game theory in that two games can have proportional
payoff matrices implying the same strategic consequence, even though
one game might have small matrix values and the other large.
The last contribution to the active space field Eq. (4.17) comes from
the inactive metric, reflecting a set of active forces from cooperative
payoffs:

 Tabinactive  1
2  jk  jk ;ab  1 4  a jk  b jk  1 2 g ab jk g cd  jk ;cd
 1 8 g ab  g cd  mn jk mn ;c jk ;d  3g cd  jk
;c  jk ;d  . (4.20)

Though cooperation is part of the hidden isometries, the metric elements


carry energy and momentum and thus are active forces. The inactive
metric behaves as a matter field with a specific form determined by the
field equations. The inactive metric field carries inertia. It can be bent by
Persistent Behaviors 135

strong gravitational sources. The existence of this field was not foreseen
by our initial analysis in Sec. 1.7. It is not uncommon however in
theories based on those of Kaluza [1921] and Klein [1956].
The last set of field equations obeys the massless wave equation,
which determines the cooperative payoffs in terms of sources:

1
2 g ab a  jl b lk    T j k  1 4 F j ab Fk ab . (4.21)

Again, the first term arises from the organizational fields restricted to
inactive components. These terms are for now arbitrary, since they don’t
have a clear analogy from physical models other than the perfect fluid.
This may not be a good guide. The second term is non-zero only if there
is a non-zero overlap between players’ payoffs: there must be common
cause.
In the absence of all sources, Eq. (4.21) is a massless wave equation
(a Klein-Gordon field in physics texts). These fields (along with other
possible fields) in Einstein’s theory could describe gravitational waves.
Thus even in empty space there may be influences that propagate with
the speed of light.
In general, the active and inactive field equations determine metric
potentials that are not Minkowski. Thus there is no a priori justification
for requiring the active and inactive metrics to be flat. Even in spaces
with no organizational sources, the payoffs contribute as sources and
provide the source for curvature. Thus given the equations, the next
logical step are to apply the decision process theory using these
equations for decision problems, solve the equations for such problems
(numerically), analyze the results and based on the results, and refine the
theory. In the process we expect to grow a quantitative understanding of
the decision process.
With suitable computing power, we have sufficient information to
apply the decision process theory to a wide variety of problems. As
anyone in engineering knows however, it is not just the theoretical ability
to solve problems that is of singular importance, it is also the practical
ability to find sets of problems that can be solved and from them, the
knowledge that gives the ability to build structures (bridges, circuits,
etc.) that do desired things.
136 Geometry, Language and Strategy—Vol. 2

We demonstrate such a set of solvable and useful problems exist, in


analogy to the set of problems that have been found in electrical
engineering. In that discipline, great strides are made using the insights
gained from electrostatics and magneto-statics, coupled with introducing
time dependence using phasors (in mathematical terms, using Fourier
series). The same mathematics of static solutions can then be applied to
phasor solutions to gain insight into dynamic steady state behaviors. We
are able to carry out this program for a class of models (Chap. 5). It is
our belief that the local equations in the normal-form coordinate basis
don’t always give a clear large scale global picture; but insight might be
gained in other frames. In the next section, we return to the notion of
persistency and express it in more general terms so that we can consider
the field equations in other frames. We will look at the equations in the
orthonormal co-moving reference frames in Sec. 4.6.

4.5 Isometry and Hidden Symmetry

Our intuitive definition of persistency (isometry) required the existence


of the normal-form coordinate basis, in which the metric is independent
of the inactive dimensions: these dimensions are hidden reflecting a
hidden symmetry. Translations along each of these inactive dimensions
leave the metric unchanged. In this section, we formulate this hidden
symmetry covariantly so that we can transform the concept to other
frames.
In differential geometry, a transformation that leaves the metric
elements unchanged is called an isometry. A necessary and sufficient
covariant condition for there to be an isometry is that there is a vector
field K satisfying the “Killing” conditions:

K  ;  K ;  0 . (4.22)

Since this is a covariant relation, to prove this we choose a holonomic


coordinate system to which we can always transform the coordinates so
that the vector field is unity along the dimension  and zero along the
other coordinates. The above condition implies that
Persistent Behaviors 137

K   g  K   g  ,
K  ;   g    ,
K  ;  K ;   g     g  2 ,
K  ;  K ;   g     g    g     g   g  ,
K  ;  K ;   g   0 . (4.23)

Therefore given the condition Eq. (4.22), there is a frame in which the
metric is independent of the dimension  .
Conversely if there is a frame in which the metric is independent of
the dimension  , we pick K  to be the vector field that is unity along
that direction and zero otherwise. The above argument again holds and
we deduce Eq. (4.22). Since it is a covariant relation that holds in one
frame, it therefore holds in all frames. This relationship is attributed to
Killing.
If there are two isometries, there will be two Killing vectors K  and
L . From two vector fields we create a new vector field called their
commutator or Lie product Eq. (4.4):

K,L  K L;  L K; . (4.24)

We leave as an exercise that the commutator of two Killing vector fields


is itself a Killing vector:

K,L;  K,L ;  0 . (4.25)

This has far reaching consequences. The set of all Killing vectors, which
includes their commutators, form a Lie Algebra, with the Lie product.
The isometries of the theory therefore generate a local symmetry group
(i.e. a symmetry group at each point in space), which has the same
structure at every point. This provides substance to the notion that
isometries are persistent and reflect the local symmetry group.
Isometries reflect properties of what we term an internal group,
which gives substance to invariance and persistency associated with the
inactive strategies. In particular, we see that the concept of independent
players has in fact been framed as a group theoretical statement that there
138 Geometry, Language and Strategy—Vol. 2

is a commutative subgroup at every point with exactly the same group


structure. Because the operators commute, we can always find a frame in
which the metric components are simultaneously independent of all the
associated inactive dimensions. We now have a firm theoretical
foundation. We next investigate the consequence of these isometries in a
specific frame, a co-moving frame, using the harmonic gauge.

4.6 Co-Moving Orthonormal Coordinate Basis

Field equations of the type Eq. (4.1) and their solutions have been
extensively studied, for example [Wald, 1984, p. 252ff], and [Hawking
& Ellis, 1973]. The essential point for us is that solutions to these field
equations exist and have reasonable properties if the metric, orientation
potentials, matter fields and gradients of these are specified initially on a
surface “orthogonal to time”. The field equations then determine the
evolution of the metric, orientation potentials, matter fields and flows on
all surfaces that are “later”, Sec. 6.6. Our goal is to use these solutions to
decision process theory to better understand the decision process. We
believe this understanding must be based on a quantitative understanding
of the consequences of these field equations. To this end it is natural for
us to adopt and adapt the approach to solving equations of this type that
has proven useful in the past (Cf. Sec. 4.3).
For these reasons, we look at the equations in a frame that is not
necessarily diffeomorphic to the normal-form coordinate basis, but is
one that is obtained from a linear transformation. The frame we want is
one in which one direction is along the energy flow vector that arises
naturally in considering the inertial fields Eq. (3.37). The energy flow
direction is a natural “time” direction. In this frame we are co-moving so
the vector has only a time component and the remaining directions are
orthogonal unit vectors with respect to the metric. We further look for a
system in which this metric is globally Minkowski. Because the metric is
constant, in general the coordinates will not be holonomic.
We call this frame the co-moving orthonormal coordinate basis or if
there is no ambiguity, the co-moving basis. Note that in general, it may
not be possible to physically transform to this frame simultaneously at
Persistent Behaviors 139

every point in space. We only indicate that there is a linear


transformation that accomplishes this. Our goal in obtaining solutions in
this space is to then use them to obtain solutions in the normal-form
coordinate basis. In this section we will use the term gauge
transformation for linear transformations even if they are not
diffeomorphisms.
The Minkowski metric we denote as m . It is diagonal with a value
1 for the timelike indices and 1 otherwise. For our discussion in this
section3, we use Greek letters in the beginning of the alphabet
 ,  , , to represent the dimensions in the co-moving basis and the
Greek letters in the middle  , ,  , to represent the dimensions in the
normal-form coordinate basis when we don’t want to specify whether
the dimensions are active or inactive. Thus in this latter case the indices
span both active and inactive dimensions.
In each frame there will be orientation potentials, which transform not
as tensors but as specified in the frame transformation Eq. (2.37). In each
frame the covariant derivative depends on these orientation potentials.

We use the notation X ; for example, to indicate the covariant
derivative that depends on the potentials   in the normal-form
coordinate basis, Eq. (2.26). Ordinarily, the notation would be adequate
for any frame, however when we switch back and forth between two
frames, it is helpful and less ambiguous to have a special notation for the
covariant derivative in the co-moving frame as well as a notation for the
potentials in that frame. We use X  | for the covariant derivative and
  for the orientation potentials in that frame. To carry out our
program of solving the field equations, which are frame covariant, we
use the transformation properties of these potentials based on their gauge
properties Eq. (2.37).
We obtain the transformation properties of the orientation potentials
from the transformation properties of the vector X  in the normal-form
coordinate basis and its covariant derivative. The transformation is

3
Later, we introduce the player fixed frame model in which we make further distinctions
between some of these orthonormal coordinates, which we associate with “proper” active
strategies and some we associate with “proper” inactive strategies (players).
140 Geometry, Language and Strategy—Vol. 2

determined by the gauge transformation matrix E   that takes the vector


in the normal-form coordinate basis to the co-moving frame:

X   E  X 
X   E  X 
. (4.26)
E  E    
E   E    

We also provide the inverse transformation. We raise and lower indices


using the metric in the appropriate basis:

E m E  g 


. (4.27)
E g  E  m

The content of the first of these equations is that if the frame


transformations E are known in the co-moving orthonormal
coordinate basis, then the normal-form coordinate basis metric
components are determined. Moreover, since the normal-form metric has
the property that gaj  0 for all mixed tensors between the active and
inactive space, the frame transformation will have to enforce this
property.
The frame transformations determine not only the metric elements
from Eq. (4.27), but the payoff matrix Eq. (2.47) using the defining
equations for the 1-forms:

dU j  1 2 F j ab Ua  Ub . (4.28)

The differential 2-form d U j can be computed from the transformation


from Eα and the corresponding 2-form based on zero torsion Eq. (2.28):

U j  E j  Eα
. (4.29)
dU j   E j  |  E α  E β

Comparing the two expressions we find that each player field E j acts as
the vector potential for the payoff in the co-moving basis:
Persistent Behaviors 141

F j ab E a E b   E j  |  E j |  f j
 . (4.30)

Thus whenever the transformations are determined in the co-moving


basis, the co-moving payoffs f j are determined. We note that the
vector potential as used here is in the full model consisting of both active
and inactive coordinates. In the normal-form coordinate basis, the
potential above for each player j , is the covariant vector  j . In
mathematical form, this is the statement above that the payoff is
associated with the single 1-form U j with unit coefficient.
As we develop the theory, a physical, geometric and global meaning
of these transformations will be helpful. We make progress in this regard
by looking at the implications of the transformations on the acceleration
potentials, Eq. (2.37). Consider the set of directions along  in the co-
moving frame and the corresponding covariant vectors E  . The
transformation rule for the acceleration potentials leads to the following
form for the covariant derivative:

E ;    E     E    E  E  . (4.31)

We can use Eq. (4.31) to determine the orientation potential in one frame
from the covariant derivative.
An alternate form is the inverse:

E   |    E       E      E  E   . (4.32)

Here we consider the (inverse) transformation matrix E  to be a vector
field (timelike and spacelike) in the co-moving orthonormal coordinate
basis labeled by the normal-form dimensions  . As noted, these
expressions reflect the transformation properties of the acceleration
potentials:

      E  E   E   E   E   E  . (4.33)

Equivalently, we have the inverse relationship:

      E   E  E   E  E    E   . (4.34)
142 Geometry, Language and Strategy—Vol. 2

These equations provide helpful relations as we formulate models.


In the normal-form coordinate basis, we expressed the field equations
based on the harmonic gauge. We finish this section by expressing the
harmonic gauge condition and exactness statements in the co-moving
basis. The consequence of the harmonic gauge follows from Eq. (4.31),
writing the active and inactive metrics out explicitly using the normal-
form potentials:

g ab E  a ;b   jk E  j ; k    
. (4.35)
1
g
a  
 g g ab E  b    

We use the harmonic gauge condition Eq. (4.13) to obtain:

g ab  a E  b    . (4.36)

We will be interested in specifying frames that are torsion free, Eq.


(2.28), that are orthonormal so that g ab is determined by the frames from
Eq. (4.27), that have the orientation potentials determined by the field
equations, that satisfy the frame Eq. (2.28) and that satisfy Eq. (4.36), the
wave equation. The advantage of this wave equation is that it is
formulated in terms of the exact active strategies x a in the normal-form
coordinate basis.
As an example of the use of this wave equation, consider a model in
which there is a direction in the co-moving orthonormal coordinate basis
Eυ that is also exact. Applying Eq. (4.3) to this, we have the condition
E ;  E ;  , which has the following solution:

 a E j  0
. (4.37)
 a Eb   b Ea

The first equation requires each component to be constant; we pick the


gauge in which the constant is zero, E j  0 . Based on the second
equation, there will be a potential field y whose constant surfaces
define a coordinate and whose associated vector field is determined by
the potential, Ea   a y . That the space might be curved will limit the
Persistent Behaviors 143

number of such coordinates. However, when there is such a coordinate,


the condition Eq. (4.35) provides the wave equation:

g ab  a  b y    . (4.38)

If the metric is known in terms of the harmonic coordinates and if the


same is true of the orientation potentials, this equation provides the
coordinate potential y in terms of the harmonic coordinates.
An alternate and useful approach is to couple Eq. (4.32) with the
harmonic gauge condition, considering the active and inactive
dimensions separately:

m E a |   a bc g bc   a kl  kl 
1
g
b  
 g g ab  0
. (4.39)

m E j
 |   j
bc g 
bc j
kl  0
kl

The active components are zero because of the harmonic gauge


condition Eq. (4.13) and the inactive components are zero because of the
properties of Eq. (2.48). Because the active directions are exact, we have
in addition, Eq. (4.3):

E a |  E a  | . (4.40)

This determines the exactness condition in the co-moving frame in terms


of the orientation potentials. In this frame there is no guarantee that the
differential operators commute, so we don’t expect that in this frame the
transformation matrix will be expressed in terms of the gradients of a
potential field. This is an example where our intuition from Newtonian
physics is not helpful. We anticipate that the directions that don’t
commute will be the flow direction with any direction whenever there is
frame rotation.
The geometric significance of the influence of frame rotation can be
illustrated with the simple example of a sphere [Gockeler & Schucker,
1987], where the latitude and longitude are holonomic coordinates  ,
but they are not orthonormal. The coordinates on a sphere that are
orthonormal are sin  d  , d  , however these coordinates are not
144 Geometry, Language and Strategy—Vol. 2

holonomic. From Einstein’s theory, time and a space direction t , u


may be holonomic but not orthonormal. The coordinates that might be
orthonormal are  
g tt dt , du , but in general these will not be holonomic
because of frame rotation effects.
We will have occasion to use both types of frames in subsequent
chapters. For example we distinguish the co-moving basis as short for
the co-moving orthonormal coordinate basis. This frame in general is
not holonomic. Following the nomenclature and analysis in Vol. 1, there
is also a central holonomic frame (Cf. Sec. 6.10, Ex. 31). In general this
frame will not be orthonormal. We extend the previous analysis by
solving the field equations for a class of models whose behaviors will be
displayed in both frames.

4.7 Decomposition of the Orientation Potentials  

We have a general decision process theory that leads to field Eq. (4.1)
that allow us to investigate decision processes using the scientific method
to expand our understanding and our approach. We have players
identified in a persistent way with inactive strategies along with the
decisions they make that are active. We treat time as an active
dimension. We have argued that there will always be a timelike unit
vector that describes the flow of energy as well as a metric tensor such
that these vectors and tensors depend only on the active strategies
available to each agent. Decisions occur in this space-time x a that ranges
over time and the active strategies. In this normal-form coordinate basis,
none of the tensors depend on the inactive space components  . In this
j

frame, the metric components between the active and inactive space are
zero. In the last section, we extended this general discussion to the co-
moving orthonormal coordinate basis.
The vector Va in Eq. (3.37) representing the flow of energy occupies
a central role in the theory. Some of the orientation potentials related to
the flow in the co-moving frame have large scale geometric significance
that makes it easier to understand the anticipated behaviors. We take one
of the orthonormal coordinate basis vectors to be the flow of energy
E   V . We indicate this dimension by the Greek letter  .
Persistent Behaviors 145

The following choices have proved insightful in the study of fluids.


We believe they will be helpful in the study of decision processes. The
application of the transformation rule Eq. (4.31) can be used to define an
antisymmetric vorticity tensor    , a symmetric expansion
tensor    and an acceleration vector q :

E ;   V ;   E  E    E  E   q E  V
      
. (4.41)
    q
    0

There is no term proportional to V E  because unit flow is orthogonal


to its covariant gradient; equivalently we use the antisymmetric nature of
the orientation potential Eq. (2.34) in the first two indices to prove this.
Thinking of the flow as a fluid, one considers a unit cube of fluid and
follows its path in the co-moving frame. The cube can rotate where the
rotation is specified by the vorticity tensor and the cube can expand,
contract or shear as specified by the expansion tensor. The acceleration
of the cube is given by q in the co-moving coordinate basis since

DV
 V ; V   q E  . (4.42)

Because the orientation potentials are antisymmetric in the first two
indices, these same parameters determine  , so that an expansion of a
vector transverse to the flow is:

E ;    qVV        V E      E V     E  E  . (4.43)

In this expression none of the indices are along the flow,  ,  ,  ,   .


We define a rate of change of the transverse vector as

DF E  DE 
  qV     E  . (4.44)
 
146 Geometry, Language and Strategy—Vol. 2

We have introduced the Fermi derivative [Hawking & Ellis, 1973] to


define the rate of change:

DF X  DX  DV  DV 
  X V   V  X . (4.45)
   
This rate of change, which is defined in terms of the covariant derivative,
has the following useful properties that help us understand the geometric
meaning of the orientation potentials. Any pair of vector fields that have
zero Fermi derivative will maintain lengths and angle. The Fermi
derivative measures the amount of rotation of the basis. Using the Fermi
derivative, we deduce that the significance of the antisymmetric
orientation potential coefficients      is that they specify
the frame rotation along the path.
We note that transformation to the flow vector has both active and
inactive components V a and V j , which is a transformation to a vector
field whose 1-form does not in general vanish. In other words, the flow
corresponds in general to a 1-form that is not exact. In Einstein’s theory
of relativity, this is stated as the twin paradox: twins traveling along
different paths may age differently as a result of their paths seeing
different accelerations: the paths are not inertial. This is true in this
theory as well. This makes sense because we expect the flow to reflect
the influence of sources that generate acceleration. We address the
question about whether the transverse components of the orthonormal set
needs to be exact in the next section.

4.8 Player Fixed Frame Model

We see that a key characteristic of behaviors in this theory will be the


vorticity and frame rotations. We focus our analysis on a specific model
in which the expansion tensor plays little role along the lines of the
centrally co-moving hypothesis of Vol. 1. We recall our assumption that
space-time locally has the property that in an open set of any local
region, the distinction that strategies have values is expressed
mathematically by the existence of a scalar function x a that defines a
hypersurface in that region (of dimension one less than the total
Persistent Behaviors 147

dimension of the space-time). We say that the coordinate 1-form is exact,


meaning its 1-form vanishes, d U α  0 . In the normal-form coordinate
basis, the inactive strategies are not exact whereas time and the active
strategies are exact.
To understand the consequences of the theory, we need computable
and relevant models. Let us suppose that there are na active strategies in
the normal-form coordinate basis, n i inactive strategies or players
giving a total of n  n a  ni space dimensions. We consider a wide class
of models formulated in the co-moving orthonormal coordinate basis,
which we call the player fixed frame model, in which there are na
proper-active strategies y in which the corresponding 1-forms are
exact. We denote these active strategies by the Greek letters
 , , , . In addition to the proper-time coordinate  , which is
the coordinate along the flow and these proper-active strategies  , we
assume the remaining n i strategies are inactive and denote them by
 ,  , , and call them proper-inactive strategies. In general, the
proper-time coordinate and the proper-inactive strategies will not be
exact.
This is a very broad class of models in that we allow any number of
players, any number of strategies and any specification of inertial and
applied forces. Though this class of models appears to us currently as
natural, we will need to consider a broader class of models as we gain
better tools with which to solve the equations. For now however, we
think that this class of models is of sufficient interest to form the basis
for quantitative examples in this book.
We analyze the consequences that the proper-active strategies are
exact Eq. (4.37). We also write in the co-moving basis, the condition that
the inactive strategies satisfy the Killing conditions Eq. (4.22). The
consequences follow from the assumption of exactness, Eq. (4.37):

E j  0
. (4.46)
 a Eb   b Ea

The consequence of the exactness condition dE υ  0 can also be


imposed on the torsion-free requirement Eq. (2.28), with  ,  ,    , :
148 Geometry, Language and Strategy—Vol. 2

dE υ  
       Eυ  Eο    E υ  Eυ
  
     E  Eα         Eυ  Eα
  Eα  Eβ  0 . (4.47)

We have expanded the wedge product using the notation from the
previous section, where we separate the behaviors along the energy flow
direction, the proper-active strategy directions and the remaining proper-
inactive strategy directions.
We convert the consequence of exactness to conditions on the
orientation potentials. We write these orientation potentials using the
decompositions Eqs. (4.41) and (4.43):

           0
           0
       . (4.48)
 
      0
 
   

With this notation the five distinct terms give five sets of conditions,
which can be simplified:

    
   0
      0
. (4.49)
    0
             
    

These conditions along with the additional conditions Eq. (4.46) on the
frame implies that there is a coordinate potential y that is a scalar
function of the harmonic coordinates and Ea   a y .
We add the following additional assumptions to the player fixed
frame model that we have found by trial and error to be consistent and
Persistent Behaviors 149

provide us a subclass of solvable models that articulate one notion of


large scale persistency:

q  0
      0 . (4.50)
    0

Unless otherwise stated, we take this subclass to be the player fixed


frame model, though we can’t exclude other possibilities. We will justify
that these conditions are indeed consistent and articulate this notion of
persistency. To obtain further relationships, we pursue the consequences
of persistency as viewed in the co-moving basis. We do that in the next
section.

4.9 Persistency under the Player Fixed Frame Model

In the last section we took a step closer to what we claim will be a


complete set of numerically solvable equations with which to determine
the properties of the decision process theory based on a general
theoretical framework. We articulated an approach by considering a wide
class of models, the player fixed frame model. These models, with a few
tweaks from the next chapter, are shown to satisfy the centrally co-
moving hypothesis, Ex. 24, Sec. 6.10.
That we expect these models are solvable numerically requires some
explanation. There are theorems that show that the field Eqs. (4.1) have
solutions for any model. The general theorems don’t yet provide
tractable ways of doing the numerical analysis for these coupled non-
linear partial differential equations for all models. Instead, solvable
models based on simplifications are constructed that help one to
understand particular features of the theory. What we would like to do is
find such a set of solvable models. We believe the current approach
comes close to meeting these criteria, as well as providing additional
insight into the meaning of the equations. To proceed we have three
remaining steps.
 We impose the consistency of the orthogonality relations Eq. (4.27).
 We analyze the Killing relations Eq. (4.22).
150 Geometry, Language and Strategy—Vol. 2

 We provide the field Eq. (4.1) in the co-moving orthonormal


coordinate basis.
In this section we address the Killing relations. In the next chapter
we address the field equations and provide a set of solvable models in
Chap. 6.
We flesh out the consequences based on the player fixed frame model
from the orthogonality relations (4.27) for the mixed tensor components
that are zero:

Ea E j  Ea E j   0 . (4.51)

There are no terms with E j since these transformation components are


zero because the exactness condition of Eυ implies Eq. (4.46). Implicit is
our assumption that the matrix E j is not singular, which implies that
the transformation vector E a is proportional to E a :

E a  e E a . (4.52)

Substituting this into the expression Eq. (4.51), we get a projection of the
flow onto the inactive directions:

Ea  E j  e E j   0 . (4.53)

Because at least one component of the timelike flow is not zero, the
implication is:

E j  e E j . (4.54)

A consequence of this rule is that if we interpret E j  as the charge for


player j , then e is the proper charge, or player engagement for proper-
player  .
The orthogonality relations also demonstrate that the derivative
operator associated with the proper inactive strategy is not zero but
proportional to the proper-time frame derivative:

  E a  a  e  . (4.55)
Persistent Behaviors 151

We call  the variation along the flow, the variation along the proper-
time coordinate. We also see that the operator  j is zero based on Eq.
(4.54), as it should be on the space of functions of interest:

 j   E j E a  E j e E a   a  0 . (4.56)

We get further information on the charge by considering the


consequences of gauge invariance using the wave Eq. (4.38) and its
companion Eq. (4.36) for the flow:

g ab  a  b y   q   
. (4.57)
g ab  a E  b  0

Here we make use of the assumptions above concerning the values of the
orientation potentials. We apply Eq. (4.52) to the gauge condition Eq.
(4.36), and get:

g ab b Ea  g ab b  e Ea   0 . (4.58)

This shows that the “charge” is conserved along a streamline.

g ab b  e Ea   0
e g ab b Ea  g ab Ea  b e  0
. (4.59)
e g ab b Ea   e  0
   e  0

The result follows directly from the second line in Eq. (4.57).
Since the coordinates x a are exact, this coordinate vector E a in a
new coordinate system satisfies Eq. (4.40): E a |  E a  | . For the player
fixed frame model, we apply these conditions and find that there are two
distinct classes of equations:

 E a   E a  q E a  2  e E a
. (4.60)
 E a     E a   2   2    e  E a
152 Geometry, Language and Strategy—Vol. 2

This shows that in this basis, the vector fields are determined. The first
equation determines the “curl” of the vector field in terms of the
acceleration and what we show below is the electric field in the co-
moving frame. The second equation determines the “curl” in terms of
what we show later to be the payoff matrix in the co-moving frame.
We have determined equations for all the vector fields Ea associated
with the active components. We now turn to the vector fields E j
associated with the inactive components. In the normal-form coordinate
basis, the Killing vectors are the inactive vectors that lie along each of
the coordinate directions. So for example the Killing vector field K  j 

along the j direction has components K  j    j . We transform this to


 

the co-moving basis to get:

K  j   E  K  j   E j  .
 
(4.61)

We interpret the transformation field E j  for each fixed inactive strategy


j to be the Killing vector field in the co-moving frame.
Given our decomposition of coordinates into proper-time, proper-
active and proper-inactive, there are six cases of the Killing relation Eq.
(4.22) to consider, using the notation of the previous section that
 ,  ,    , :

E j |  E j |  0
E j |  E j |  0
E j |  E j |  0
. (4.62)
E j |   E j |  0
E j |  E j |  0
E j |  E j |  0

We summarize the result for each case in order, leaving to the exercises
at the end of the chapter the derivations.
(1) We start with the first case and write out the covariant derivative in
terms of the orientation potentials with the player fixed frame model
that the transformations depend only on the proper-time coordinate
and proper-active strategies:
Persistent Behaviors 153

e  E j  e  E j  0 . (4.63)

(2) The second case provides the proper-time variation of the proper-
active strategy of the Killing vector field components:

 E j  q E j  2  E j . (4.64)

(3) The third case provides the time variation of the proper-inactive
strategies of the Killing vector field components to be zero and
shows that Eq. (4.63) is identically satisfied:

 E j  0 . (4.65)

(4) The fourth case has no time derivatives and is identically satisfied
based on Eq. (4.46):

 E j    E j  0 . (4.66)

(5) The fifth equation gives the spatial dependence of the proper-
inactive components:

 E j  2 E j    E j  . (4.67)

(6) The sixth case is the time derivative of the component of the Killing
vectors along the energy flow:

 E j  q E j  q E j  0 . (4.68)

(7) We can use the above case with Eq. (4.54) to determine the
“charge” gradient:

 e  q e  2  2  e e   e  . (4.69)

The important overall result is that we get first order partial


differential gradient equations for the proper-time evolution of the flow
component, proper-active components and proper-inactive components
of each Killing vector field. We get two sets of constraint equations
involving gradients and one algebraic equation. The frame
154 Geometry, Language and Strategy—Vol. 2

transformations E j , E j  are independent of proper-time and the


remaining components E j are zero.
These equations determine the inactive metric and show that it is
independent of proper-time, since each of the components is similarly
independent:

 jk  E j m Ek  E j m Ek  . (4.70)

From the harmonic gauge conditions Eq. (4.39) and exactness conditions
a a
Eq. (4.60) we obtain the equations that determine E  E  :

1  e e   E



a
  E a   q     E a
. (4.71)
 E a   E a  q E a  2  e E a

Given this solution to Eq. (4.71), the active metric is determined:

g ab  1  e e  E a E b  E a E b . (4.72)

The active metric may depend on proper-time. Even in the case that it is
independent of proper-time, it does depend on the acceleration and so
there will be effects due to acceleration in the solution.
We have pointed out previously, Eq. (4.30), that the co-moving
payoff is determined once the transformations are known. We can also
proceed as follows to determine the payoff. The payoff matrix is an
acceleration-potential in the normal-form coordinate basis, Eq. (2.48) as
seen for example in  jab . Since the orientation potentials are related, we
use Eq. (4.32) to determine the payoff fields:
1
2 F j ab    Ea Eb  E j    Ea Eb  E j
  q E j   e E j   Ea Eb  Ea Eb 
      E j  e E j  Ea Eb  Ea Eb  . (4.73)

The expression is informative. It shows that terms such as  


determine the payoff matrix and that the electric field is determined by
   . This shows that we are on the right track and that we can gain
Persistent Behaviors 155

insight into the equations in the co-moving frame. From this form we
determine also the various co-moving payoff components:

f j
 0,
f j
 0,
f j
  e f j
 ,
f j
 
 2 q E j       E j  e E j  1  E k  Ek  
2 e E  j
1  E k
 Ek 
,
f j
   2  E j  2  E j . (4.74)

Zero payoffs in the co-moving basis are required between proper-


inactive strategies and the proper-time coordinate and between proper-
inactive strategies.
We need the field equations using Eq. (4.1) and its explicit form in
terms of orientation potentials Eq. (2.41) applied to the co-moving frame
to provide the complete set of equations. The field equations will provide
first order partial differential equations for the orientation potentials. We
examine these equations in the next chapter, where we see the vorticity
effects are made explicit as well as additional gradient effects. We reduce
the equations to linear coupled differential equations that provide the
underlying mechanisms for vorticity behaviors and gradient behaviors
of decision processes.

4.10 Outcomes

After completing this chapter, the student should understand that


persistency in decision process theory defines what it means to be an
agent or player. The underlying mathematical structure is the group of
isometries, showing the relationship between agents, players and
measurement. There are quantitative consequences from this group of
isometries that are expressed by differential equations, which have
unique solutions in the harmonic gauge, given initial conditions specified
on a spacelike hypersurface orthogonal to time. The equations have
analogs to equations from physics, though these equations describe a
156 Geometry, Language and Strategy—Vol. 2

different space, time and phenomena. The student should learn that the
large scale behavior of the theory differs from the local behaviors
because of the possibility of acceleration: i.e. dynamic changes of
strategic choices. We illustrate the possible large scale behaviors by
introducing the player fixed frame model in both the normal-form
coordinate basis and the co-moving coordinate basis. This model refines
our notion of large scale persistency in a way that is consistent with and
follows from the local principles of the theory.
The attainment of the outcomes is facilitated by doing the exercises at
the end of this chapter. To further help the reader, we list more detailed
outcomes from this chapter by section.
 From Sec. 4.1, the student will have learned that decision process
events at each point in time can be described locally in a coordinate
(holonomic) basis in which each coordinate is a surface of constant
potential for a potential field whose normal provides the coordinate
direction. This underlies the assumption that we can specify the
distance between any two decision process events at separate points in
time.
 In Sec. 4.2, the student learns that in the normal-form coordinate
basis, gauge invariant results are directly obtained. This provides an
example of a “global” basis that has advantages over the “local”
coordinate basis. Though not holonomic, the basis is effectively
holonomic when transformations are restricted to the active subspace.
 In Sec. 4.3, we frame the field equations in the normal-form
coordinate basis using the harmonic gauge that removes the gauge
degrees of freedom. These field equations form the basis for the
quantitative predictions of decision process theory.
 In Sec. 4.4, the field equations that result are derived, supported by
results from the exercises. These equations provide the student a
guide to the behaviors expected in the theory and can observe the
similarities and differences between decision process theory and
physical theories.
 In Sec. 4.5 the student will have learned that the Killing relations
from differential geometry provide a covariant articulation of
isometry, which is a common group structure at each point of space
Persistent Behaviors 157

and time. In decision process theory, the set of players form the basis
of this symmetry and so are persistent attributes of any solution.
 In Sec. 4.6, the student will be introduced to the co-moving
orthonormal coordinate basis. The student should be able to
transform between this basis and the normal-form coordinate basis
and understand how to express the covariant derivatives, exactness
conditions and harmonic gauge conditions in each basis.
 In Sec. 4.7 as preparation for creating models in the co-moving basis,
a geometric decomposition of the orientation potentials is given.
Certain of the orientation potentials describe how the frame rotates
when moving along particular directions.
 In Sec. 4.8 the player fixed frame model is defined and the
consequence of exactness developed.
 In Sec. 4.9, the consequences of persistency are developed, along with
the further constraints of the commutation constraints and harmonic
gauge constraints. The student should understand the basis of this
class of models and the defining equations of this model in the theory.

4.11 Exercises

(1) Using the transformation to normal-form coordinates Eq. (2.45),


show that the Lie product of any two inactive vectors is zero,
 U j , U k   0 .
(2) Using the transformation to normal-form coordinates Eq. (2.45),
show that the Lie product of an active and an inactive vector is zero,
 Ua , Uk   0 .
(3) Using the transformation to normal-form coordinates Eq. (2.45),
show that the non-zero components of the Lie product of any pair of
active vectors are

Ua , Ub    Ua , U b  j  j   F j ab j ,

which correspond only to gradients on the inactive space. If one


restricts attention to actions on fields that are independent of the
inactive coordinates, then the active vectors commute.
158 Geometry, Language and Strategy—Vol. 2

(4) Use the covariance property of the “covariant derivative” to develop


the formulae Eqs. (4.31) and (4.32). Further, show for fixed choices
of coordinate values  ,   that the transformations E  , E   
represent the corresponding coordinate directions in the new frame.
Show that their Lie product is:

Eμ , E ν           E  .

(4.75)

(5) Using Eq. (4.75) show that the Lie product of the coordinate
directions vanishes if and only if the orientation potentials    are
symmetric in the last two indices. Using the inverse transformation
below, what do you conclude about the Lie product of the
transformed coordinate directions:

Εα , Eβ            E  .

(4.76)

(6) Demonstrate the following form for the curvature tensor


components from R a b Eq. (3.28), where Rabcd are the curvature
components as computed in a geometry of only the active
coordinates:

Rabcd  Rabcd  1 2  jk Fabj F k cd  1 4  jk Facj Fbdk  1 4  jk Fbcj Fadk ,


Rabcj   1 4  b kj Fack  1 4 Fbck  a jk  1 2  jk Fabk  ,
;c

Rabjk  1
4   F g F   jl km F g F
jl km
l
ac
cd m
db
l
bc
cd m
da

 ml  a jm  b kl   ml  b jm  a kl  . (4.77)

(7) Demonstrate the following form for the curvature tensor


components from R i j Eq. (3.28):
Persistent Behaviors 159

R ij ab  1
4  lk  a jk  b il  1 4  lk  a il  b jk

 1 4 g cd Faci Fbdj  1 4 g cd Fbci Fadj ,


R ijk b  1
4 g ac a ik Fcbj  1 4 g ac  c jk Fabi ,
Rijlm  1 4 g ab a im  b jl  1 4 g ab a il  b jm ,
R ijkl  1
4 g ab a il  b jk
 1 4 g ab a ik  b lj . (4.78)

(8) Demonstrate the following form for the curvature tensor


components from R a j Eq. (3.28):

Rajcd   1 4  jk ;c Fadk  1 4  kj ;d Fack  1 2   jk Fcdk 


;a

Rajbk  1 4 g  jl  km F F  1 2  jk ;ab  1 4   kl ;a mj ;b .


cd m
ac
l
bd
lm
(4.79)
Rajlk  1 4 g  lm jk ;d F  1 4 g  km jl ;d F
bd m
ab
bd m
ab

(9) Using R j abc from Eq. (4.79) and R j abc  R j bca  R j cab  0 from Eq.
(3.71), show Eq. (1.20), which implies that the payoff matrix is
derivable from a potential.
(10) Using the explicit forms from the previous exercises for the
curvature tensor in the normal-form coordinate basis, show that the
Einstein Eq. (4.1) yields the generalization of Maxwell Eq. (4.14) in
the harmonic gauge. In a general gauge, show that the result is:

1
2
1

   jk g bc Fabk  ;c
  T ja . (4.80)

(11) Using the harmonic gauge and the additional “covariant” gauge
condition Eq. (4.15) develop the wave equations for the player
vector potentials Eq. (4.16) from Eq. (4.80).
(12) Using the explicit forms from the previous exercises for the
curvature tensor in the normal-form coordinate basis, show that the
Einstein’s Eq. (4.1) yields the field Eq. (4.17) for the active
strategies.
(13) Using the explicit forms from the previous exercises for the
curvature tensor in the normal-form coordinate basis, show that the
160 Geometry, Language and Strategy—Vol. 2

Einstein Eq. (4.1) yields the field Eq. (4.21) for the inactive
strategies.
(14) Write and discuss the field equations in the absence of a matter field
and with inactive and active metric components equal to their
Minkowski (flat) values.
(15) Show that Poisson’s law, Eq. (4.18), results as a static, weak field
limit of Eq. (4.17).
(16) Demonstrate that the commutator (Lie product) of two Killing
vectors is again a Killing vector.
(17) Show that the results Eq. (4.49) follow from Eq. (4.48).
(18) Using the expression for the Lie product Eq. (4.76), for the player
fixed frame model and using the notation for that model from Sec.
4.8, show that the Lie product for co-moving basis proper-active
strategies are:

Ευ , Eυ   2    e    E a


a

. (4.81)
Ευ , Eυ   2  E j  2  E j
j

(19) Using the expression for the Lie product Eq. (4.76), for the player
fixed frame model and using the notation for that model from Sec.
4.8, show that the Lie products for co-moving basis proper-active
strategies with the co-moving flow vector are:

Εο , Eυ     q  2 e  E a
a

. (4.82)
Εο , Eυ   q E j  2 E j
j

(20) Using the expression for the Lie product Eq. (4.76), for the player
fixed frame model and using the notation for that model from Sec.
4.8, show that the Lie products for co-moving basis proper-active
strategies with the proper-inactive strategies are:

Ε υ , Eα    e  E a  2 E a
a

. (4.83)
Ε υ , Eα    E j   2 E j
j
Persistent Behaviors 161

(21) Using the expression for the Lie product Eq. (4.76), for the player
fixed frame model and using the notation for that model from Sec.
4.8, show that the Lie products for co-moving basis proper-inactive
strategies and of these strategies with the co-moving flow are:
 
Εα , Eβ   Εο , Eβ   0 . (4.84)

(22) Deduce the seven cases that are provided that follow from
Eq. (4.62).
(23) Derive the following for the gradients of the inactive metric:

  jk   2q E j Ek  2      E j Ek  E j Ek  


 2 E j Ek  ,
  jk  0 . (4.85)

(24) In the beginning of this chapter we recalled the centrally co-moving


frame hypothesis concept from Vol. 1. It can be generalized to the
idea that time is a mutually commuting isometry with the player and
coded of conduct isometries and the energy flow is stationary. Do
we get the same hypothesis as before by adding the additional
constraint that the active components of the flow vector in the
holonomic basis are zero, except for the time component? Show that
this means we obtain that for those components K a  V a . The
open question is whether the time components K j must be zero.
(25) In later chapters we explore models satisfying the centrally co-
moving frame hypothesis in Ex. 24. Extensions of such models
depend on picking specific models for the energy momentum
tensor, such as the conductivity model in Eq. (3.76). The
longitudinal and transverse conservation laws are given by Eq.
(3.77). Show that with the centrally co-moving frame hypothesis
that the transverse conservation law can be expressed as follows for
the active dimensions (so time is excluded here):

a p 1 
 2  a jkV jV k   1 2 h jk Fabk  bV j . (4.86)
p p
162 Geometry, Language and Strategy—Vol. 2

Show that the inactive conservation laws, which determine the


inactive flows (including the time), are:
1
4 h jk Fbdk Vl Fcal g ab g cd  1 2 h j k  a kl h l i  aV i


h jkV m
g

 b   aV k  1 2  ki  a ilV l Vm g ab g   0. (4.87)

Show that the longitudinal conservation law is not consistent if the


active flows are zero, since the result is:

V aa 
 p

1
g
a  
g V a  12
p

 aV j h jk  aV k . (4.88)

In this case the left hand side is zero if the flows are zero, whereas
the right hand side is not zero since the shear is not zero. What does
this imply? Can these equations be solved in a consistent fashion?
Chapter 5

Deconstructing the Theory

We have identified the mathematics of decision process theory, Eq.


(4.1). The expression is concise and self-consistent. Despite its merits of
brevity, to most readers unfamiliar with this mathematics, it does not
give great insight into the behavior or consequences. For that we have to
expend significant effort to analyze these equations. We need to identify
the concepts and distinctions of that theory as applied to decisions. We
need examples. We turn to engineering for guidance. In electrical
engineering, whose equations are in fact similar, a large effort is devoted
to learning about DC circuits since the concepts can then be applied to
AC circuits. The general case of transient behavior can then be
understood in that context. Similarly in mechanical engineering we can
start with statics, and then proceed to a description of standing waves,
steady state waves and finally general dynamics. We would like to do the
same for decision process theory. We want to identify a simple set of
cases that are steady state and then extend them to more interesting
cases.
We think steady state models may help illuminate how we use the
decision process theory for engineering models in general. For those
familiar with electromagnetic theory, we note that our approach is
analogous to starting with magnetostatics in which the currents are static
and the magnetic fields don’t vary in time. We then extend such
solutions to harmonic solutions (phasors). Such electrodynamics models
provide significantly more insight than electrostatic models and with
concepts of phasors (harmonics), provide an important step towards a
full understanding of the theory. We go from DC circuits to AC circuits.

163
164 Geometry, Language and Strategy—Vol. 2

The only step missing is the study of transient effects, which is also
illuminated by this approach.
The simplest situation for us however is not strictly analogous to an
electric circuit; however it does have a lot in common with transmission
lines and traveling waves. In both cases it makes sense to work in a
frame of reference in which the behavior is observed to be steady state. If
we travel along with the wave, then there will be no observed flow. In
this chapter we consider the case of decision processes which have that
property: it is possible to find a co-moving frame of reference in which
all of the dynamic attributes of the problem are steady state. Technically,
this means it should be possible to identify a frame of reference that is
co-moving, holonomic and one in which time and the inactive player
strategies are persistent and mutually commuting. We could then use the
principle of least action to provide us with the equations of motion. With
appropriate numerical techniques we could then solve these equations.
We then use these solutions to obtain the normal-form coordinate basis
solutions using the harmonic wave equation and the initial conditions.
There is a practical barrier however. There are a large number of
equations, which represent concepts that we have not explored. The
purpose of this chapter is to carry out this exploration and come up with
a set of equations that we can both solve and understand. To achieve that
goal, we find it useful to take a bottom-up approach rather than a top-
down approach. We start with a specific class of models that leads to our
goal, the player fixed frame models. We write the consequences of each
of the equations of motion and see what distinctions are implied. This is
a challenging task because it demands a fair amount of mathematical
juggling. To help the interested reader follow the thread, much of the
work is summarized in tables and the details are left as exercises. Toward
the end of the chapter we return to the goal stated here and demonstrate
that we have indeed solved the class of models of interest: models that
have steady state wave behaviors.
The detailed plan of the chapter is as follows. We first rewrite the
field equations in the co-moving orthonormal coordinate basis, or co-
moving basis for short. We then review the frame equations in this co-
moving basis, followed by the field equations making the assumptions of
the player fixed frame model. The basic results of the chapter are
Deconstructing the Theory 165

summarized in Tables 5.1, 5.4, 5.5 and 5.6. The expressions for the
frame are provided in Sec. 5.1. In Sec. 5.2 we show how we generated
the field equations in the co-moving frame and in Sec. 5.3, we simplify
the expressions using the player fixed frame model. To interpret these
results for decision process theory, in Sec. 5.4 we articulate the
distinctions that result from an examination of this model.

5.1 Player Fixed Frame Model

Before diving into the equations and the model, let’s recall how we got to
the equations. We have emphasized the importance of having a common
ground for the discussion of decision processes based on a theoretical
and scientific framework. The framework should be based on generally
agreed principles. We suggest the common ground be based on the
mathematics and empirical basis of physical theories as well as the
foundations of game theory and related works in economics. In decision
process theory, we take a unified and consistent view, which has led us
to replacing static theories and equilibrium frameworks with the physical
principle of least action articulated as the field equations, Eq. (4.1). We
have demonstrated that this mathematical expression, though compact, in
addition to providing the self-consistent framework summarizes a great
deal of information that we believe is commonly understood to be part of
the decision process. In Sec. 4.4, using the normal-form coordinate
basis, we expanded the field equations into three sets of equations and
sketched the information that might be derived from each set.
To make further progress in extracting predictions from decision
process theory, we consider models that have attributes in common with
static models; this is in the same sense that DC circuits provide insight
into AC circuits. We can use the same mathematics as long as we extend
the interpretation of what the mathematics represents. One main
departure is that we go from a static equilibrium or point to a
consideration of steady state behavior that extends throughout the
strategic space.
Our bottom-up approach is to proceed with the program set out in
Sec. 4.9 to find solvable models based on simplifications that are
166 Geometry, Language and Strategy—Vol. 2

constructed to help us understand the features that are particular to


decision process theory. We explore the properties of the player fixed
frame model introduced in the last chapter, which highlight the role of
agents and players as manifested by their vorticity, gradient and
persistency properties. This is a general class of models that articulates
large scale structure that we think will be attributes of every decision
process. We emphasize that this approach is designed to help us
understand the content of the theory. It does not exhaust all possible
model behaviors expected in such a theory.
We want to impose no practical barriers to considering small or large
number of players, to including competition or to include variations
depending on whether or not the games are fair. Furthermore, at the
outset we want to consider that active decision flows are dynamic, which
is to say that they change in time. We have to deal with some challenges
however due to the complexity of the equations. The equations become
increasingly more complex as the number of active strategies increases.
These considerations have influenced our choice of assumptions with the
player fixed frame model in Sec. 4.8. Without limiting the number of
active strategies, we impose the restriction that in the co-moving frame,
the active strategy preferences are holonomic.
We find that the field equations and frame equations can be solved
more easily than the wave equations that lead us back to the harmonic
coordinates. We therefore break out some of the complexity of the
problem by moving to the co-moving orthonormal coordinate basis. We
think the field equations in the co-moving frame provide significant
initial insight. Implicit in our move to the co-moving frame is an
emphasis on those solutions that have steady state aspects. It will turn
out that implicitly we have imposed the condition that there is a time-like
isometry vector, which means we have the analogs of gravity and
gravitomagnetism.
For the reader with some knowledge of theories of this type, they may
know that the study of the field Eq. (4.1) is a large and ongoing project
that may take considerable effort and time to flesh out. Much work has
already been done. However, our applications are different from those in
physics and so the program that has been carried out in that domain is not
entirely applicable here. Indeed without some analysis such as we
Deconstructing the Theory 167

suggest here, it is not clear what aspects can be usefully carried over. It is
also worth noting that discussions in other disciplines assume the reader
already understands a good deal, so there is some advantage in creating a
self-contained discussion of the solutions.
It is important therefore to specify a program in our domain that
identifies what is relatively simple from what is relatively complex. We
start that project. Our goal is at least to get to the point that we can see
that it is possible to have a complete dynamical numerical solution for
problems in which there are any number of strategies and any number of
players. We should be able to carry out this program for one, two, three
or four active strategies. As numerical techniques improve, it should be
possible to extend to a larger number of strategies: such approaches have
been used in other fields of study.

5.1.1 Dynamic Components

The key idea behind the player fixed frame model from Sec. 4.8 is that
in the co-moving frame of reference, the active strategy variables are
exact, which imposes dynamic constraints. We made additional dynamic
assumptions Eq. (4.50) to those of exactness, Eqs. (4.46) and (4.49). We
did not motivate those assumptions, a task we take up here:

q  0
      0 . (5.1)
    0

The first equation makes the plausible statement that the energy flow has
no acceleration along the proper-inactive directions. The player exercises
some persistency in his or her decision process.
In the second and third set of equations the interpretation is that the
co-moving frame has a fixed orientation with respect to the players and
that the only player bonding will be along the active strategy directions
and will be captured by  . In particular, all compression or expansion
coefficients such as  and the symmetric part of    are zero.
Along the flow direction, the frame rotation  is zero and so the Fermi
derivative of the frame rotation also vanishes in Eq. (4.44).
168 Geometry, Language and Strategy—Vol. 2

Though we explore the distinctions that arise from these concepts in


more detail in Sec. 5.4, here we give some initial insights about
additional frame rotations. First, we extend the notion of the Fermi
derivative Eq. (4.45), which was defined relative to the flow and
provided physical insight into the rotational behavior of the frame along
the flow. We take the directions to be along the orthonormal directions
and below, indicate with the Greek letter  both the component index
and the scalar variable. In this paragraph, we temporarily suspend our
notation from the last chapter and allow this Greek letter to represent
both proper-active and proper-inactive strategies:

DF X  DX  DV  DV
  X V  V  X
   
. (5.2)
DX  
 X  ; E 


We express these Fermi derivatives of the orthonormal vector fields


E , E  , using the expansions Eqs. (4.41) and (4.43) for the
orientation potentials:

DF E
  E 

. (5.3)
DFV
0


Not only is the Fermi derivative of the flow zero along the direction of
motion, it is zero along any of the orthonormal coordinates. The Fermi
derivative of any frame component along any other frame direction is a
rotation set by the orientation potential restricted to the transverse space.
We return now to our notation from the last chapter and state the
consequence of our model assumptions Eq. (5.1) for the variation of
E j :
Deconstructing the Theory 169

DF E j 
  E j    E j  0

. (5.4)
DF E j   
  E j    E j  0


Based on the player fixed frame model, the orthonormal vectors for the
inactive space (the Killing vectors) remain fixed when moved along any
of the transverse directions. In particular we see the value of having
  0 and   0 .
We have already commented about the frame rotation  being zero
along the flow direction. Our assumptions amount to the requirement that
the frames do not rotate along the flow or along any of the transverse
orthonormal axes. Our assumptions can be summarized as saying that
with this hypothesis, the co-moving orthonormal coordinate basis is a
player fixed frame model, justifying its name. The orientation of the
proper player does not change. This class of models enhances the notion
of player persistency and extends the notion of persistency to be an
attribute that applies to the large scale structure. We do expect to see
some aspects of game theory in our solutions. We will also see
significant differences. In particular we see strong vorticity effects
evidenced in the vorticity tensor components    .

5.1.2 Persistent Components

Persistency is a consequence of the Killing equations, Eq. (4.62), which


we summarize in Table 5.1. These results were derived in the last chapter
and provide the essential results for the frame evolutions. For further
details see Sec. 4.9. In Table 5.1 we provide equations for the
transformations that take us from the normal-form coordinate basis to
the co-moving orthonormal coordinate basis. One of the goals is to
obtain equations for the transformations from the normal-form
coordinate basis to the orthonormal coordinate basis.
It is noteworthy that the inactive Killing vectors E j E j E j 
are steady state for the player fixed proper-time through a wave equation,
assuming we can define a proper-time variable. We see that the
170 Geometry, Language and Strategy—Vol. 2

transformations from active to inactive involve the player engagement


e , which is steady state. The Killing conditions provide sufficient
gradient equations for each steady state field so that these fields are
determined, given appropriate initial conditions.

Table 5.1. Player Fixed Frame Model—Persistent and Dynamic Components

Distinction Variable Properties


The active flows are not in general steady state and
Ea satisfy Eq. (4.57) based on the exact character of the
Energy active flow and on the wave equation:
Flow E j
g ab  a E  b  0
.
The wave Eq. (4.57) determines the proper active
Proper- Ea strategy in terms of the harmonic coordinates:
active
strategy E j  0 g ab  a  b y   q    .

We have the exactness conditions Eq. (4.60) for x a


a
E 
and the harmonic gauge condition Eq. (4.71)
Active
Strategy E a
  e E a  E a   E a   q E a  2  e E a 
E a  E a     E a   2   2    e  E a  .
1  e e   

 E a
   E  a
  q    
 E a

The proper-time variations Eq. (4.65)  E j  0 and


Inactive- E j Eq. (4.68)  E j  0 , are zero. The gradients are
Strategy Eqs. (4.67) and (4.64); the Killing vectors determine
E j the payoffs Eq. (4.30) using Eq. (4.74):

Killing E j  E j  0  E j  2 E j    E j  ,


Vector
E j  E j   q E j  2  E j ,
Player- E j f j
  f j
 0,
Potential f j
 2  E j  2  E j , f j
 e f j
,
   
1
2 f j
  q E 
j
  e E

1  E E 
j k
 k

      E j
 e E 1  E
 j k
 Ek   .
Deconstructing the Theory 171

Orthogonality leads to Eqs. (4.54), (4.55), (4.59) and


Player E a  e E a (4.69), showing in particular that the player
engagement engagement e is conserved and the charge gradient
determined:

E j   E j  e
  e 
.
 e  0
 e   q e  2  2 e  e   e 

5.2 Field Equations—Co-Moving Frame

We gain insight about decision processes by working in the co-moving


frame. The field Eqs. (4.1) in the co-moving orthonormal coordinate
basis are computed using the flux 2-form Eq. (3.27). The exercises from
the last chapter addressed the technical aspects of this computation for
the normal-form coordinate basis. In this chapter we outline the steps for
the same calculation in the co-moving basis.
Our approach will be to take the expression for the orientation flux in
terms of the potentials and expand them. We provide a guide for all the
time evolution equations in Table 5.2. We list all the space evolution
equations in Table 5.3. The calculations in this section are for any model.
In the next section we specialize to the player fixed frame model. By
examining the calculations here and in the next section, along with the
corresponding exercises at the end of the chapter, the reader will gain a
deeper understanding of the field equations and their consequences. In
Sec. 5.4, these results will be applied to the identification of distinctions
in the player fixed frame model.
In this section, we find it convenient to use more compact
expressions, so we restrict Greek letters to not be along the flow:
 ,  ,  ,   . We start with the defining relationships for the
orientation flux Eq. (3.27):
172 Geometry, Language and Strategy—Vol. 2

R α ο  dω α ο  ω α β  ω β ο
R α β  dω α β  ω α ο  ω ο β  ω α γ  ω γ β . (5.5)
 ,  ,  

For the appropriate orientation potentials we use Eqs. (4.41) and (4.43):

  q
    
. (5.6)
    
,  

The orientation potential 1-forms that result from these definitions are:

ω α ο     V     Eβ  q V          Eβ
. (5.7)
ω α β     V     E γ     V     E γ

Note the suggestive notation V  Eο for the flow 1-form.


To obtain the differentials of the orientation potentials we use the
differentials of the frames based on the condition of no torsion Eq.
(2.28):

dV        Eβ  E γ  q Eβ  V
dEα     Eβ  E γ              V  Eβ . (5.8)
 ,  ,  

We have sufficient information to compute the curvature components in


Eq. (4.1). We take the mixed components in Eq. (5.7) and compute in
Ex. 7 the curvature components R and R , Eq. (5.67). From these
curvature tensors, we obtain a variety of results that are detailed in the
subsections that follow.
Deconstructing the Theory 173

5.2.1 Time Evolution of Vorticity and Expansion

Based on the symmetry condition R  R , Eq. (3.70), we get our
first result, the time evolution of the vorticity:

   1 2  q ||  q ||                      . (5.9)

The double bar notation for the components of the covariant derivative is
a short hand for the terms that involve only the transverse components:

q |    q     q     q  q ||     q  q || . (5.10)

In other words, this is the covariant derivative defined on the


hypersurface orthogonal to the proper-time coordinate. It makes some of
the intermediate algebra simpler, though in this case it makes no
difference.
Next, we look at the symmetric curvature tensor R component
Eq. (5.67), called the tidal force by Hawking & Ellis [1973], which
determines the time evolution of the expansion that appears in the
literature:

           R  q q


 1 2  q |  q |            . (5.11)

We have used Eq. (5.9) to eliminate the time dependence of the vorticity.
We return to Eq. (5.11) later, though we note that this is a useful result in
its own right.
We obtain a field equation by contracting the tidal force using Eq.
(3.74) with n space components that total the number of active and
inactive strategies. We obtain an expression in which only the energy
density and average pressure appear:

  p
R  R        p  . (5.12)
 n 1 

We combine this equation with the time evolution of the expansion to get
the evolution of the volume compression coefficient   m  :

174 Geometry, Language and Strategy—Vol. 2

  q |              
  p
2           p  . (5.13)
 n 1 

This formula in the  3  1 dimensions of physical space and time is


attributed to Landau and Raychaudhuri by Hawking & Ellis [1973].

5.2.2 Codacci Equations

Not all equations that we get are time evolution equations. The
possibility of field equations being independent of time exists with
Maxwell’s equations. For example, Coulomb’s law provides a set of
equations that are independent of time, with only spatial partial
derivatives. For equations of the Einstein type, such relations are named
after Codacci. In a space of n space dimensions and one time
dimension, there is a theorem that there will be n  1 Codacci equations.
We are able to identify n of these equations here.
The mixed component R is  p , which is zero. This contracted
curvature tensor is:

R  R   m R . (5.14)

We use this with Eq. (5.67) from the exercises at the end of this chapter
to obtain the n Codacci equations, [Hawking & Ellis, 1973, Eq. (7.17)]:

   ||     ||  ||  2 q  . (5.15)

These constraints are on the surface orthogonal to the flow and contain
no time derivatives. The constraints are only spatial partial differential
equations that must be satisfied on the surface transverse to the time
flow. Since there are n  1 Codacci relationships, we need to find one
more.
Deconstructing the Theory 175

5.2.3 Remaining Equations

The remaining set of equations (lines 3-5 of Table 5.2) are computed
from the transverse components of the orientation potential 1-forms ωα β ,
Eq. (5.7). From Ex. 8 at the end of the chapter, the differential 1-form
Eq. (5.68) contains no proper-time partial derivatives of the frame
rotation  , suggesting that it is set by other constraints. It does not
appear in the Codacci constraints Eq. (5.15), suggesting that there are no
time evolution equations for the frame rotations  . Indeed, we don’t
expect evolution equations for q or  , since their proper-time
derivatives don’t occur in the curvature tensor components, though we
may expect contributions from   q  e  q .

Table 5.2. Time Equations

Determination Source Number in Number


dimension D D=4
3
q Flow equation D 1

3
 2  D  1 D  2
Gauge choice 1
(e.g. exactness)

6
 Field Eq. (5.17) 1
2 D  D  1

3
  2  D  1 D  2
Symmetry R  R 1
Eq. (5.9)

9
 Symmetry R  R 1
2  D  1  D  2 
2

Eq. (5.71)

24
2 D  D  1
Total equations Orientation potentials 1 2

We do obtain proper-time differential equations for the transverse


orientation potentials  , Eq. (5.71) in Ex. 11 as a result of the
symmetry relations R  R based on the computation of the
curvature components Eq. (5.70). We have computed the time evolution
176 Geometry, Language and Strategy—Vol. 2

of the vorticity, Eq. (5.9) and the time evolution of the transverse
orientation potentials  , Eq. (5.71). We have yet to compute the
evolution equations for the expansion parameters  in order to have a
complete set of evolution equations of the orientation potentials
expressed in terms of the other potentials and the sources.
We obtain the complete equation for the expansion parameters by
starting with the contracted curvature tensor in terms of the sources, Eq.
(3.74):

 p 
R  R  R      p  ph  h  . (5.16)
 n 1 

The result is based on the expression for the curvature tensor Eqs. (5.67)
and (5.70) with  ,  ,  ,   :

  q q  1 2  q |  q |              


                 1 2        1 2      
 1 2      1 2       1 2       1 2     
 p 
  p  ph  h   0 . (5.17)
 n 1 

The essential point is that these equations determine the time evolution
of the expansion parameters entirely in terms of the sources and the
orientation potentials and their derivatives along directions in the surface
normal to the flow.

5.2.4 Summary and Final Codacci Equation

In summary, the complete set of evolution equations for the orientation


potentials is:

   1
2 q
 ||  q ||                      , (5.18)
Deconstructing the Theory 177

       ||     ||     

              b  
q      q     
2q      q  , (5.19)

  q q  1 2  q |  q |              


                 1 2        1 2      
 1 2      1 2       1 2      1 2    
 p 
  p  ph  h   0 , (5.20)
 n 1 

  q ||  q q              
  p
2           p  . (5.21)
 D2

These are Einstein’s field equations. Exercise 16 demonstrates that the


last two equations are consistent, using Eq. (5.75). As a consequence we
identify the remaining Codacci relation, with  ,  ,    :

     1 2  2  1 2       1 2      
       1 2      1 2     . (5.22)

This equation shows explicitly that the energy density is the source of
curvature since a non-zero value of the energy density requires non-zero
values of at least some of the orientation potentials.
The exposition of the theory is complete in the co-moving
orthonormal coordinate basis. We make the following observations,
summarized in Table 5.3, for a space-time of dimension D  n  1 . The
other set of relationships are constraints, partial differential equations on
the surface orthogonal to the direction of proper-time and are listed in
Table 5.3.
178 Geometry, Language and Strategy—Vol. 2

Table 5.3. Gradient Equations on Surface Normal to Flow

Determination Source Number for D dimensions Number


D=4
Field Eq. (5.22) 1 1

Field Eq. (5.15) 3


R  0 D 1

Symmetry Eq. 3
R  R (5.73) 1
8 D  D  1 D  2  D  3

Bianchi identity1 0
R  R  R  0 Eq. (5.76) 1
24  D  1 D  2 D  3 D  4

Bianchi identity 1
R  R  R  0 Eq. (5.77) 1
6  D  1 D  2  D  3

Total spatial equations Gradient 8


equations

It is important to count the number of unknown potentials needed to


specify a solution to the equations. In a holonomic frame, such as the
normal form coordinate basis, it is enough to find a solution for the
metric potential g  . Because of the gauge freedom, the number of
unknown potentials is
1
2 D  D  1  D  1 2 D  D  1 .

It can be shown that the curvature tensor computed from these metric
potentials will then automatically satisfy all of the symmetry relations,
Sec. 3.11, Exs. 9-13.
The situation is quite different in the Co-Moving Coordinate Basis.
The metric is a constant, the Minkowski metric. The unknowns are the
orientation potentials in Table 5.2, and the symmetry relations are no
longer automatically satisfied. Structurally, the curvature tensor is

1
See [Gockeler & Schucker, 1987] and Cf. Sec. 5.6, Ex. 19.
Deconstructing the Theory 179

antisymmetric in the first two indices and the last two indices, so the total
number of components is

D 2  D  1 .
2
1
4

The number of independent components is further reduced by the


symmetry properties. The number of symmetry equations is
1
8 D  D 2  1  D  2  ,

plus the Bianchi identities,


1
24 D  D  1 D  2  D  3 ,

which is
1
6 D 2  D  1 D  2  .

The net number of independent components is therefore (Cf. Sec. 3.11,


Ex. 13):
1
12 D2  D 2  1 .

In general, there are many more curvature components than field


equations.
The orientation potentials are set by equations from three sources: the
field equations, the symmetry equations and a gauge that fixes any unset
values. We start by looking at the field equations. The total number of
field equations is
1
2 D  D  1 .

Of these, the time equations in Table 5.2 determine the compression


matrix  ; lines 1-2 of the gradient equations in Table 5.2 determine
the Codacci equations.
In the co-moving orthonormal frame, the field equations are not
sufficient to determine the time evolution of the remaining orientation
(acceleration) potentials; however they are the only equations that
180 Geometry, Language and Strategy—Vol. 2

depend on the inertial properties. The symmetry equations determine the


time evolution of the remaining orientation potentials and provide
gradient equations, lines 3-5 in Table 5.3.
All the equations and potentials are now accounted for except  ,
which are unspecified and hence can be chosen at will and the
acceleration transverse to the flow q . We turn next to a specification of
this acceleration using the conservation laws on the energy momentum
tensor that describes the organizational dynamics.

5.2.5 Conservation Laws

We end this section with the conservation laws Eq. (3.42) in the co-
moving orthonormal frame. These conservation laws are imposed on the
sources because of the field equations, see Ex. 16 in Sec. 3.11, Eq.
(3.75). We have both a transverse and longitudinal part:

m 
 p  q  h  p  | h 
d . (5.23)
 V  |  V | p  0
d

Transformed to the co-moving frame the first conservation law has no


components along the flow direction: The acceleration is purely
transverse, V  q . The transverse conservation law along with the
longitudinal conservation law can be written (Ex. 17), with  ,  ,    :

 m
  p   q     p    p     p
. (5.24)
      p  0

The first equation is algebraic and linear for q  in which all the indices
are transverse to the flow. The second equation gives us the time
evolution of the energy density.
Deconstructing the Theory 181

5.3 Field Equations—Player Fixed Frame Model

To fully articulate the consequences of the player fixed frame model,


Eq. (5.1), we apply all of the equations identified in Sec. 5.2 and
summarized in Tables 5.2 and 5.3. We must in fact verify that the model
assumptions are self-consistent as well as show that the equations have
solutions. We do that in this section. These calculations provide deeper
insight into our model assumption as well as further insight into the
general theory. If questions arise about these results, their origins will be
clear and subject to verification and challenge.
The results are summarized in Tables 5.5 and 5.6. We provide
exercises at the end of the chapter as guides for these results. Equally
important are the distinctions about decisions that are the outcome of
these calculations, which we apply to the orientation potentials. We
discuss the distinctions in the next section. In this section, we revert to
the notation in which Greek letters at the beginning of the alphabet
 ,  ,  , represent proper inactive strategies and the Greek letter
upsilon,  , , , along with primes represent the proper active
strategies.

5.3.1 Expansion Equations

We start with the first set of equations in Table 5.2, the field Eqs. (5.17)
for      , the expansion coefficients. Based on the model
assumptions, the expansion coefficients are zero in both the active and
inactive spaces, but not necessarily the mixed. The time variations for the
first two must be set to zero in Eq. (5.17) based on these model
assumptions:

   0
. (5.25)
  0

When both components are both active, the first equation simplifies, with
 ,  ,   , :
182 Geometry, Language and Strategy—Vol. 2

 1 2  q  1 2  q   1 2     1 2     q q 


2   2     2         
  p 
  p   ph   h    0. (5.26)
 n  1 

This equation provides partial differential equations for the components


of the acceleration of energy at each point in proper-time.
When both components are inactive, the resultant Eq. (5.17) is a
divergence condition for    :

      q       2    2  


 p 
        p  ph  h  . (5.27)
 n  1 

In general the divergence is not sufficient to determine the potentials.

5.3.2 Six Symmetry Classes

To determine the potentials   we need the consistency equation


R   R  from Table 5.3 in which two components are active and
two inactive, Eq. (5.72), which simplifies to:

e    e      


     2  2 
2   2   0. (5.28)

This will pose an interesting constraint on models as it involves the


commutator of the strategic compression/rotation matrices.
In addition there is a second consistency equation R   R  that
follows from Eq. (5.72), which again after simplification is:

e   e     


  2   2 
2   2    0. (5.29)
Deconstructing the Theory 183

These two equations combine to form a single equation, which gives


back each equation when extracting the symmetric and antisymmetric
part in  ,  :

2e         


   4  4 . (5.30)

If the time dependence is zero, this equation determines the “curl” of


 .
In general the divergence and curl are sufficient to determine the
potentials (see Ex. 18). The additional constraint on the matrices
emphasizes that these potentials are matrices in the inactive space.
The results of the symmetries Eq. (5.72) from Table 5.3 were helpful
in providing equations that determine the orientation potentials. We
should consider all such relations. There are six classes of consistency
equations involving active and inactive indices that follow from these
symmetries:

R   R 


R   R 
R  R
. (5.31)
R  R
R    R  
R     R   

We have covered the first two. The third and sixth are identically
satisfied in the player fixed frame model. The fourth is:

e   e  . (5.32)

The fifth gives the “curl” of the payoff in the co-moving basis, after the
usual simplifications:
184 Geometry, Language and Strategy—Vol. 2

     2          


    2          
    2            0. (5.33)

This “curl” is analogous to no magnetic monopoles in Eq. (1.20)


expressed in covariant notation. In the holonomic frame, these conditions
imply that the payoff field can be written in terms of a potential, Eq.
(1.21).

5.3.3 Time Evolution of the “Mixed” Expansion Coefficient

To complete the analysis of the time evolutions Eq. (5.17) for the player
fixed frame model, we write the result, after simplifications, for the
expansion coefficient  , which in general provides the time evolution:

       q   2  


           
 1 2 e  q  1 2 e    1 2 e     p  0. (5.34)

If we think of the potential  as the payoff matrix for a proper-player


 in the co-moving frame, then we see that it is analogous to the
magnetic field in Maxwell’s equations, Eq. (1.19).
We can view Eq. (5.34) as the analog of Ampère’s law suggesting
that the expansion coefficient  is analogous to the electric field and its
contribution in Eq. (5.34) is the analog of the displacement current in
electrical engineering2. We then view  p  as the current along the
strategic direction  for proper-player  . We shall firm up these
distinctions in the next section.

2
Electric and magnetic fields depend on the frame of reference. Our terminology in this
section is specific to the co-moving frame of reference.
Deconstructing the Theory 185

5.3.4 Codacci Equations

We tackle item two next in Table 5.3, which provides the consequences
of the field equations that result from R  0 that are named the
Codacci equations, Eq. (5.15). After simplification, in the player fixed
frame model, the proper inactive components give the divergence of the
mixed sum of expansion and vorticity components:

             2 q


               . (5.35)

The active component gives the divergence of the active vorticity


components:

          2 q    2      


 e      . (5.36)

We have a total of n such relations including both the active and


inactive components. The inactive set Eq. (5.35) consists of terms
reminiscent of Coulomb’s law Eq. (1.14), though we are missing the
charge and have the extra divergence    .
To complete the items in Table 5.3 we need the first element Eq.
(5.22), which after model simplification, gives the energy density in
terms of the potentials:

           3   


2     3 2       1 2   
 1 2      1 2     . (5.37)

We already have an expression for the divergence, Eq. (5.27), so an


alternate expression is one in which the partial derivatives are absent:
186 Geometry, Language and Strategy—Vol. 2

 p 
  p   ph     h       
 n 1 
2           q  3 2     
 3 2     1 2      1 2      0 . (5.38)

We have an algebraic set of relations between the energy density and


diagonal stress components with the orientation potentials. This
completes the analysis of the items in Table 5.3.

5.3.5 Vorticity Equations

We turn to row four of the time evolutions, Table 5.2, which gives the
time evolution of the vorticity components, Eq. (5.9) that result from the
symmetry R  R . We write out these conditions for the various
cases,     , starting with the first case that both indices are
active. After simplification the result for the player fixed frame model is:

    1 2    q   q    2     2   . (5.39)

The second case is the vorticity with mixed components, whose


simplified expression is:

   1 2 e  q . (5.40)

The third case gives the time evolution for the inactive components,
which after simplification is:

   0 . (5.41)

We set these vorticity components to zero on the initial hypersurface and


this shows that they remain zero at all later proper-times along an
appropriate path.
Deconstructing the Theory 187

5.3.6 Remaining Potentials and their Evolution

The last orientation potentials to analyze from Table 5.2 are the time
evolutions of the last row. There are six cases to consider:

          . (5.42)

We start with the first case  and simplify Eq. (5.71) using the model
assumptions:

  e       e     


         . (5.43)

In the model, we have the symmetry    .


The antisymmetric part of this equation is:

      2         . (5.44)

If this starts at zero it stays at zero over time. The symmetric part of
Eq. (5.43) is:

  e   e  . (5.45)

This assumes that the antisymmetric part of  is in fact zero. It then
follows from this and Eq. (5.32) that:

e   e  . (5.46)

This allows for a simplification of the analog of Ampère’s Law,


Eq. (5.34):

    q  2


     
 1 2 e  q   p . (5.47)

Sticking with the same elements of Table 5.2, we consider the second
potential in Eq. (5.42). In the model, after simplification it yields:
188 Geometry, Language and Strategy—Vol. 2

     e                  


                
 q        q      . (5.48)

The equations are antisymmetric in the active indices demonstrating


consistency with the model assumptions Eq. (4.49).
The third case in Eq. (5.42) with all three indices active yields a non-
zero time evolution, which after model simplification is set to zero based
on the exactness constraints Eq. (4.49):

        2     q   


    2     q  
    2     q    0. (5.49)

We obtain a set of “curl” equations for the vorticity potentials   .


The fourth case,  in Eq. (5.42), simplifies to:

   e   e   e    e    0 . (5.50)

It is identically zero (i.e. no new constraint) using Eqs. (5.46) and (5.40).
The fifth term  computes to an identically zero time evolution:

  0 . (5.51)

The sixth term   we compare with the time evolution of


     Eq. (5.48) and take the difference, since the two are equal
by the constraint Eq. (4.49). We make simplifications, obtaining the
result:
1
2         e      
 q      
   q        0 . (5.52)

In this case the result is not automatically zero, so that we obtain the
“curl” equation for  . Coupling this equation with the time variation
Eq. (5.48), we achieve the following simpler expression:
Deconstructing the Theory 189

      q  q


     . (5.53)

Comparing this with Maxwell’s Eq. (1.14), these equations are the
generalization of Faraday’s Law to decision process theory.

5.3.7 Conservation Laws

We have now computed all of the items in Table 5.2; we have not
explicitly checked Eq. (3.71), but leave to an exercise to check that they
are satisfied. The remaining equations result from the conservation laws
on the inertial sources, Eq. (5.24). The longitudinal conservation law in
Eq. (5.24) simplifies for the player fixed frame model:

   2 p . (5.54)

In addition, there are two transverse conservation laws to consider, one


for accelerations along  and the other along  .
After simplification the first of these is:

 p  p q  e  p    p    p . (5.55)

The second is:

 p  q  pq   p


 p  2 p  e  p . (5.56)

These are the differential equations for the stress components.

5.4 Player Fixed Frame Model—Distinctions

The results from the last section are summarized in Tables 5.4, 5.5 and
5.6 in this section. We discuss the distinctions that arise from these
results. In thinking about the results from the previous sections, it is
tempting to base distinctions solely on concepts that arise in game
190 Geometry, Language and Strategy—Vol. 2

theory, economics or other decision processes. In engineering however


we are familiar with the problem. There are many types of engineering
problems in each of the engineering disciplines. Historically, there were
specialized distinctions associated with each particular application. The
power of the engineering discipline and the scientific principle in general
is to see the underlying connections that span multiple applications. Thus
engineering terminology is often based on the terminology of the
underlying physical laws, which are abstract and mathematically based.
This makes the concepts somewhat more objective.
We see some advantage for decision process theory to adopt the same
stance and rely more on the physics nomenclature than application
specific distinctions. We see an advantage with distinctions that are more
neutral as they help to provide the common ground for discussion.
However, we balance this stance with the understanding that there is an
advantage to use names that appear to apply to the subject matter as they
make the arguments clearer. Thus we feel justified that in introducing
decision process theory, we relied on foundational aspects of game
theory, Sec. 1.4. We believe it useful to discuss effects with examples,
where we propose how the theory applies. In general, we indicate by
bold italics distinctions that follow directly from the theoretical
foundation and that should be generally applicable.
We start our discussion of the results of Secs. 5.1 and 5.3 based on
these caveats. We find that there are four classes of forces that determine
decisions: electric forces, magnetic forces, tidal forces and stress or
inertial forces due to external sources. In the co-moving frame we are
able to isolate the global characteristics of these four classes.
These four forces contribute to the change in flow through the
transverse conservation of energy and momentum, Eq. (3.42). We recall
the form of this with no inertial effects, Eq. (2.49):

dV b
gab   abcV bV c  Vk Fabk V b  1 2 V jV k  a jk  0 . (5.57)
d
This shows that the acceleration (the first term) is determined by the
active orientation potentials  abc , the electromagnetic forces Vk Fabk V b
and additional forces due to the charges and the gradient of the inactive
Deconstructing the Theory 191

metric. It is worth noting that this latter term can be written in terms of
the co-moving frame orientation potentials:

q e e     1 Ee e 

 2
 1 2 V jV k  a jk     e e  a
 2

E a
2       e e e  . (5.58)
1  e e 
2

This shows how the co-moving frame orientation potentials contribute to


the acceleration. To these effects we add the inertial fields.

5.4.1 Distinctions—Persistency

The origin of persistency is the local group invariance of a commutative


group of isometries and leads to the concept of a player in decision
process theory. A possible global concept is based on the idea that the
transformation from the normal-form coordinate basis to the co-moving
basis doesn’t rotate the axes that correspond to the players. This is made
explicit in the player fixed frame model from Sec. 4.8. We provide a
concept that is a global property, corresponding to real world behaviors.
This notion of player transcends the transactions over time. As players
make choices and change their decisions, we are still able to distinguish
different players. Yet players may change their payoff matrix and change
the frequencies with which they pick strategies. The player fixed frame
model provides a mathematical basis for this type of view. It is a model
or approximation that the general decision process theory does not
require. We investigate this model and see what consequences it has and
whether reality conforms to this view. So this model is an inquiry into a
particular form of persistency that is a global property rather than just a
local property. The consequences of these ideas are summarized in Table
5.1 in which we identify four key distinctions:
 Energy flow
 Active strategy and proper active strategy
 Inactive strategy, proper inactive strategy, Killing vector and player
potential vector field
192 Geometry, Language and Strategy—Vol. 2

 Charge, player engagement, player interest


We discuss each of these distinctions in turn.
We have defined a formal unit vector field V  Eο to be the direction
along which the energy of the system flows at each point. The energy
flow is an important attribute of any physical system. What is the
meaning of energy and energy flow for decisions? For decision processes
we identified in Sec. 1.4 the energy flow direction as characterizing the
strategy chosen: the relative frequencies of the flow are the relative
frequencies with which the players collectively make their choice.
We now identify additional attributes that are associated with the
flow of energy. In game theory and many economic theories, decisions
are transactions in which something of value is exchanged. Game theory
for example rests on the idea of a utility that each player gives or
receives in a transaction, summarized by the payoff matrix of values. In
decision process theory, we exchange energy. How that occurs depends
on the nature of the forces and is dictated by the principle of least action
(Cf. Sec. 3.5).
Energy is convertible and described for all forces by their
contribution to the energy momentum tensor. Because we assume the
foundational aspects of game theory, our notion of energy and the game
theory notion of utility are consistent. We start with the same notion of a
payoff matrix including a notion of some initial game value. The payoff
field contributes to the energy momentum field so the specified payoff
values also specify the initial energy and momentum of the system.
The basic attributes of and equations for persistency in the player
fixed frame model are specified in Table 5.1. The energy flow consists of
the active strategy components E a  that reflect the strategic frequency
choices. In the co-moving frame the proper active strategy components
are along the  axes. We distinguish the inactive strategy flow
components E j as those that reflect the basic coupling of the player to
the decision process. We also call these basic components the charge of
player j. The projection of the charge onto the co-moving direction  we
call the player engagement of player  . In the co-moving frame the
proper inactive strategy components are along the  axes. We see from
Table 5.1 that the projection is E j   E j e , so it makes sense to call
e the proper charge of  .
Deconstructing the Theory 193

We made the formal distinction that inactive strategies correspond to


isometry transformations and hence provide the theoretical foundation
for what we mean by a player. We made this notion covariant by noting
that the necessary and sufficient condition for an isometry transformation
is the existence of a Killing vector field. We identify the various
components of this field as E j E j E j in the co-moving basis.
Based on the relations that must be satisfied in general for such vectors,
we obtain the equations in Table 5.1.
Our notion of a player and the underlying isometry mechanism was
introduced to create the payoff field F j ab . The results in Table 5.1 show
that the linear combinations of the Killing vectors E j E j E j are
the player potential vector fields that define the payoffs Eq. (4.30):

,   
. (5.59)
f j
  E j  |  E j |  F j ab E a Eb 

Thus the equations that determine the transformations of the inactive


flows also determine the persistency and payoff properties of the theory.
In the co-moving basis, the payoffs are specified by orientation potentials
in Table 5.1.
Below, we identify the payoffs with the electric and magnetic forces.
With that identification, the charge or player engagement for a given
player is then the strength of the coupling of sources to the field for that
player in the normal-form or co-moving coordinate basis respectively.
There will be an additional coupling specified by the player current
(Sec. 5.4.4). There is no correspondence of these notions in game theory
since two payoffs that differ by an overall constant factor have the same
strategic behaviors. We do note however that the idea of engagement is
an important concept in the real world. In a game such as poker or a real
world situation such as a world war, the size of the stakes impacts how
people play the game. So we might think of charge in practical
applications as engagement.
194 Geometry, Language and Strategy—Vol. 2

5.4.2 Distinctions—Electromagnetic Fields

The compelling reason to identify the payoffs with the electromagnetic


field is the form of the equations: either Eq. (4.16) in the normal-form
coordinate basis or Table 5.4 in the co-moving coordinate basis. The
results in the co-moving basis are striking. Comparing these equations
with Maxwell’s Eq. (1.14), we identify the electric field for player 
with  and the magnetic field for player  with   . Though
similar, the equations are by no means the same. The Maxwell-type
equations for decision theory contain corrections due to tidal forces
discussed below and reflect the existence of multiple independent player
fields.
These equations as well as all the other equations that we obtain in
this theory are of three general types that we characterize as follows:
(1) Time evolution equations: for example we have  on the left-
hand side of the equation and only field values or the spatial
derivatives on the right-hand side.
(2) Divergence equations: the spatial derivative is contracted with the
tensor on the left-hand side of the equation, for example    and
on the right-hand side there are only fields or other divergences if
they have not been specified in the theory.
(3) Curl equations: for vector fields we will have antisymmetric tensor
combinations on the left-hand side such as      and on
the left only fields (or in this case time derivatives). For two
dimensional tensor fields such as   there will be an equation
that determines the totally antisymmetric cyclic combination
            .
By analogy with electro-dynamic equations, we expect that if the form
matches Maxwell’s equations there will be well-defined solutions.
Typically one needs both the divergence and curl equations as well as, or
in combination with, a time evolution equation.
That obtains here. The concepts of electric and magnetic fields are
precise; we expect that the consequence of the equations to be as rich and
complex as those in the field of electrical engineering. One can skim any
textbook to see the diversity of effects that such equations generate. For
example, see the classic text on electro-dynamics [Jackson, 1963].
Deconstructing the Theory 195

The question is whether these consequences accurately describe the


effects of decision making. Decisions are often framed in more colorful
language. Because of the origins of the distinction payoffs from game
theory, we anticipate that the magnetic field reflects self-interest
behaviors, egotism, competition and conflict. We recall the use of
electric field components in converting a game into normal form, for
example using time as the hedge strategy in Eq. (1.12). Based on this, we
anticipate that the electric field reflects altruism, bias and persuasion.

Table 5.4. Player Fixed Frame Model—Electromagnetic Fields

Distinction Variable Properties


Electric field  Ampère’s law Eq. (1.19) is generalized by Eq.
(5.47), which gives the time evolution. Eqs. (5.35)
and (5.46) provide the generalization of Coulomb’s
law:

        q   2   


           
 1 2 e  q   p ,
e   e  ,
            2 q
               .

Magnetic field   Eq. (5.33) gives the “curl” of the co-moving
magnetic field. The time evolution is Eq. (5.53)
generalizes Faraday’s law Eq. (1.14):

     2          


    2          
    2           0,
          q 
q            .

Democratic institutions argue the need for strong self-interests


working together to find the common good. Autocratic institutions
196 Geometry, Language and Strategy—Vol. 2

demonstrate the power of coercion to create conformity and agreement.


We see that the electric and magnetic forces fuel these opposite
extremes. Rather than argue in favor of any of these positions, decision
process theory makes visible the specific mechanisms indicated in Table
5.4, which provide a more precise set of distinctions than used in the
popular press.

5.4.3 Distinctions—Tidal Fields

Although the inactive frames don’t rotate along a particular proper


strategic direction  , the size of the space can increase or decrease. The
measure of this change is the component of the orientation potential  .
As an example, the volume of an element of ordinary space in spherical
coordinates is r 2 sin  drd d  . Effectively as one increases the radius,
the directions of the latitude  and longitude  don’t rotate. However
the differential volume increases by r 2 reflecting the fact that the space
is getting flatter. We term effects that reflect the curvature of the space
tidal.
Generically the orientation potentials in the co-moving frame can be
distinguished as spin or strain depending on whether potentials are anti-
symmetric, for example    or symmetric, for example
   , respectively. For these examples, the spin components are
assumed to be zero in the player fixed frame model. These symmetric
player bonding components have no analogy in Newtonian physics.
Strictly speaking, they are not attributes of Newtonian gravity since they
are tensor forces. They are attributes of Einstein’s general theory of
relativity and do explain small but measurable effects such as the
deviation of Mercury’s path around the Sun compared to Newtonian
calculations.
In addition to these strain components, the tidal spin components 
are not zero in the player fixed frame model. Because the gradient of the
charge Eq. (4.69) is non-zero whenever these spin components are non-
zero, we think of these spin components as the tidal player interests.
They contribute contributions to Coulomb’s law Eq. (5.35) as seen also
in Table 5.4.
Deconstructing the Theory 197

Table 5.5. Player Fixed Frame Model—Tidal Fields

Distinction Variable Properties


player  “Curl” Eq. (5.52) and time evolution Eq. (5.40), where
interest the time dependence of the tidal rotation is determined by
Eq. (5.39)

       q    q 


          e    ,
   1 2 e  q .

Player  Time evolution is Eq. (5.45). The divergence Eq. (5.27).
bonding The matrix commutator and curl are determined by Eq.
(5.30), since the time dependence of the magnetic field is
determined through Eq. (5.53) by the electric field;

   2e   ,


      q            2  
 p  .
2      p  ph  h  ,
 n 1 
2e               

     4   4  .

Tidal   Eq. (5.36) gives the divergence, Eq. (5.49) gives the
magnetic “curl” and Eq. (5.39) gives the time evolution.

          2q    2     


 1 2 e e  q  e   ,
     q     2    
    q    2    
    q    2     0,
    1 2    q   q    2     2   .

The tidal magnetic components   are so named as they contribute


to the payoffs Eq. (5.59) in the normal-form coordinate basis since by
Eq. (4.74):
198 Geometry, Language and Strategy—Vol. 2

f j
   2  E j  2  E j . (5.60)

The dynamics of the tidal magnetic field differs from   as seen in
Table 5.5. For example we see that the time variation depends on the curl
of the acceleration and distinguishes the electric field from the tidal
player interest. These effects need more detailed study. We may hope to
learn more from these dynamics in decision processes than what we
currently have learned from applications in physics.
What effects do these tidal strains and tidal spins represent in decision
processes? We have been able to draw little from game theory as these
effects would be absent at equilibrium. A preliminary study by Thomas
& Kane [2010] suggests that the player interest provides the possibility
for players to demonstrate interdependent and independent behaviors.
We will review that work in Chap. 7, updating it to conform to the
decision process theory outlined here using the player fixed frame model.

5.4.4 Distinctions—Inertial Fields

The orientation potentials describe tidal spins and tidal strains, the
orientation flux fields, which relate directly to the geometry of the
behavior. Decision process theory provides not only for strains but for
the organizational system components such as p  of the stress tensor
that, along with the energy density and energy flow, characterize the
energy momentum of organizational fields. The orientation flux fields
are keys to decision process theory. The effects of frames that change
with the orientation flux fields provide the basis for an understanding of
how energy attracts or interacts with energy. Since we take energy to be
more fundamental than utility or value, we are in fact hoping to provide a
deeper understanding of economic value. We see in Table 5.6 that the
orientation flux field is directly related to the components p  of the
stress tensor along the active directions. With the physics analogy as a
guide, we see that the metric components g 00 , Eq. (4.72), in the time
direction will be determined in part by the equations for the
transformations E a , which are determined by the exactness conditions
Eq. (4.60) and found to depend critically on the orientation flux field.
Deconstructing the Theory 199

This mathematical result produces the physical result of (scalar)


gravitational attraction that does conform to Newtonian gravity.
Inertial effects occur not only in the acceleration of the flow of
energy, but contribute to the magnitude of the energy density  . The
energy density Eq. (5.37) is the sum of the contributions from the
electromagnetic fields and the other tidal fields. This conservation of
energy provides the basis for converting value from one form to another.
The average stress p we characterize as the pressure. The field
equations yield a particularly interesting result Eq. (5.38) that
algebraically relates the diagonal stress components and energy density
to the orientation potentials. These diagonal components and the energy
density are typically thought to be non-negative so this result imposes
strong constraints on the behavior of the electromagnetic and tidal fields.
Of particular interest in applications will be the behavior of the
divergence of the acceleration of energy, Eq. (5.26), where we eliminate
the divergence of the player bonding Eq. (5.27) and make use of the
algebraic relationship for the diagonal pressure components Eq. (5.38),
(Cf. Sec. 6.10, Ex. 47):

  p
 q      p    2  2 
 

 n 1 
     q  q q .
 
(5.61)

In numerical examples we may start at a point of no acceleration, so the


last two terms start at zero. We see that the inertial and player bonding
contributions generate a positive contribution, whereas the two vorticity
contributions are negative. It is noteworthy that the magnetic
contributions have cancelled and don’t appear. It is not the payoffs that
drive the system to equilibrium. The forces that drive the system towards
the equilibrium would be the inertial or compression forces. This is
indicated by the divergence.
Divergence  q is a measure of flow out of a small volume
centered at the point of interest. Thus the inertial effects lead to an
increase of acceleration as we move away from the point of interest. If
the actual flow is steady state so that the forces balance, there must be a
compensating force that is negative. This compensating force is the
200 Geometry, Language and Strategy—Vol. 2

gravitational attraction. The vorticity contributions on the other hand


cause the acceleration to decrease as we move away from the point of
interest. Such contributions provide a centrifugal effect that is opposite to
gravity.
For applications, a model must be provided for the stress tensor and
energy density. We should specify whether these scalar fields change in
time, whether they are related etc. The simplest choice is one in which all
the stress components are determined by the average stress, the
organizational dynamics model, Vol. 1:

p   ph 
p  ph . (5.62)
p  ph  0

This stress tensor defines a perfect fluid. However, we find arguments


from the field equations (Chap. 6) that extend this model.
We also need a model for the energy density, which we also take
from the organizational dynamics model. It reaches the same conclusion
that is used for physical systems: the energy density is proportional to the
pressure:

  p. (5.63)

We call the proportionality constant the structural coupling or resilience


 as the player’s response to stresses or forces that lead to actions. The
more resilient the system, the higher is the value of  and hence the
higher is the energy density. Even such simple assumptions provide
insight. We can create separately a model for the inactive space
corresponding to the players and a model for the active space
corresponding to the strategic choices.
In addition, from Vol. 1, we anticipate a relationship between the
strain observed in the geometry and the stress on the inertial fields. In
many physical systems the stress is proportional to the strain. So in a
near perfect fluid, we might have the conductivity model in which

p   . (5.64)


Deconstructing the Theory 201

In a fluid the proportionality constant is the viscosity. In statics it is


Young’s modulus. In electromagnetic theory, the strain  is the electric
field and suggests we call p the current of player  along the
direction  .

Table 5.6. Player Fixed Frame Model—Inertial Components

Distinction Variable Properties


 The stress tensor for both components active, Eq. (5.26)
Energy q
provides the gradients of the acceleration of the flow of
energy:

1
2    q   q    q q   2   
 1 2       1 2    
2          
2             
  p 
  p   ph   h   .
 n 1 

Pressure p 1
n p

  p  Algebraic relation between pressure p , which is the
energy  average stress and the stress tensor components; Eq.
density (5.38) is an alternate to Eq. (5.37) for the energy
density; Eq. (5.54) provides the time evolution:

 p 
  p   ph     h    1 2    
 n 1 
 1 2            2       

  q  3 2      3 2     0,

   2 p  .

Stress p  Eqs. (5.56) and (5.55) must be satisfied by the sources.
tensor p The stress components internal to the players are not
p constrained.

  p    q  p  q      p 


  p  2  p   e  p ,
 p  p q  e  p    p    p .
202 Geometry, Language and Strategy—Vol. 2

Therefore the stress-strain relationship is an entirely reasonable


assumption: we call the proportionality constant the conductivity of the
medium since the law is a version of Ohm’s Law where the electrical
resistance is inversely proportional to the conductivity. Ohm’s Law is
that an applied voltage generates a proportional current. More generally,
we keep the idea that p is the current of player  even with zero
player charge E j .
This is consistent with electromagnetic phenomena in which both the
charges and currents specify the coupling of the fields to the sources. In
particular a system can be electrically neutral and still display a current:
we shall see that this is the case for decision process theory. The player
charges can vanish yet there can be player currents that generate payoff
fields.
The inertial fields indicate the reluctance or willingness of the players
to make choices in a given direction. They suggest to the player whether
or not there is an impending event requiring them to change their
behaviors. This is separate from whether the choices made are right in
for example a game theory sense. When players are willing to move in a
given direction we have an effect alluded to in Chap. 3. When there is
great reluctance to change course, we have an often seen behavior to stay
the course, no matter how disastrous. We see that the inertial field
behaviors are not arbitrary and obey equations that must be considered as
part of the field equations. Though we might in isolation be able to
describe where a particular behavior might lead, we can’t trust that
behavior to be part of the overall solution because of the massive
coupling of the effects. We suggest that the mechanisms identified are
nevertheless real, as are their resultant coupled behaviors.

5.5 Outcomes

In this chapter, we accomplished our goal of deconstructing the theory


into its component distinctions. We demonstrated that there is a class of
models that may be self-consistent. We show in the next chapter that to
achieve that self-consistency we impose the requirement that the
orientation potentials don’t depend on proper-time: they are steady state.
Deconstructing the Theory 203

The resultant model then corresponds to the steady state behavior of AC


circuits in electrical engineering: these are models in which time is
inactive and mutually commuting with the player inactive strategies.
We show in the next chapter that the resultant equations are a
complete set of partial differential equations that are amenable to
numerical solution for any number of players and any number of
strategies and admit unique solutions.
For the Player Fixed Frame Model, the student will have learned that
the compact mathematical expression Eq. (4.1) for decision process
theory expands into a large number of separate and testable components
summarized in Tables 5.1, 5.4, 5.5 and 5.6. By going into this detail, the
student learns the essence of the theory. Based on these detailed
expressions, the student will understand how to test the many aspects of
the self-consistent framework against observed behaviors in decision
processes. The results come together in the player fixed frame model
with the steady state hypothesis. In Chap. 7, we apply this model to the
prisoner’s dilemma, a model that has sparked interest in the game theory
literature.
The attainment of the outcomes of this chapter is facilitated by doing
the exercises in the following section. Based on this investment, the
student should achieve the more detailed outcomes below.
 One of the goals is to obtain equations for the transformations from
the normal-form coordinate basis to the orthonormal coordinate
basis. In Sec. 5.1, the student should understand the time evolution
and gradient equations for these transformations and their basis.
The inactive Killing vectors E j E j E j  will be steady state
for the player fixed frame model, with the changes added in
Chap. 6. By contrast, the active strategy transformations
E a E a  e E a E a  depend on the proper-time through a wave
equation. The player engagement e however will be steady state.
Sufficient equations exist for the gradients of each steady state field
so that these fields are determined given initial conditions. The results
are summarized in Table 5.1.
 In Sec. 5.2, the field equations are expanded in the co-moving
orthonormal coordinate basis. The student will see the resultant
equations group into evolution equations, constraint (Codacci)
204 Geometry, Language and Strategy—Vol. 2

equations and conservation laws, characterized in Tables 5.2 and 5.3.


The Codacci constraints involve only partial derivatives that lie in a
surface transverse to the flow of proper-time.
 In Sec. 5.3, the field equations are expanded in more detail in the
player fixed frame model, producing the results listed in Tables 5.4,
5.5 and 5.6. The student will see the origin of Maxwell like equations
as well as new equations reflecting the player bonding forces of the
model.
 The student will learn new distinctions based on the field equations of
decision process theory from Sec. 5.4. In particular the student will be
able to identify the persistency associated with players,
electromagnetic like fields associated with payoff fields, inertial fields
associated with orientation flux fields and new effects associated with
such tidal fields.

5.6 Exercises

(1) Show that the Fermi derivative as defined in Eq. (5.2) leaves the co-
moving orthonormal metric (which is Minkowski) unchanged.
(2) Provide the frame equations in the player fixed frame model starting
from Eq. (4.62).
(3) Derive the time dependence of the charge, Eq. (4.59).
(4) Derive the space dependence of the player interest, Eq. (4.69).
(5) Demonstrate that the differential of the mixed 1-form potential is,
with  ,  ,  ,   (the flow is in the direction indicated by the
Greek letter omicron ‘  ’)

dω α ο    

 
      q q    q V  Eβ

                b   V  Eβ

 
                     Eβ  E γ
 q      Eβ  E γ . (5.65)

(6) Use Eq. (5.65) to show that the mixed curvature 2-form is, with
 ,  ,  ,   :
Deconstructing the Theory 205

R αο    


      q q    q  q    V  Eβ 
  
  
  
  


b  V E β

b          V  Eβ


                    Eβ  Eγ 
   


       q      Eβ  Eγ .  (5.66)

(7) Extract the following curvature tensor components from Eq. (5.66),
with  ,  ,  ,   :

R               


          q 
             
          q  ,
R        q q    q
                
 q           . (5.67)

(8) Use Eq. (5.7) to show that the transverse components of the 1-form
potential have the following differential 2-form, with  ,  ,  ,   :

dω α β    

             V  Eγ 
       q            V  Eγ


                   E  E γ . (5.68) 
(9) Use Eq. (5.68) to show that the transverse curvature 2-form is
(  ,  ,  ,   ):
206 Geometry, Language and Strategy—Vol. 2

R α β  d ω α β  ω α ο  ω ο β  ω α   ω β ,
R αβ   

            V  Eγ 
        q  q         V  Eγ 
       q                          V  E γ


                          E  E γ 
  
            E  E γ .  (5.69)

(10) Extract the following curvature tensor components from Eq. (5.69),
with  ,  ,  ,   :

R             


 q      q     
    q                      ,
R                   
              
                . (5.70)

(11) Use Eqs. (5.70), (5.67) and the symmetry relations R  R to
establish the time evolution of the transverse orientation potentials
with  ,  ,  ,   :

                      


| |
  
               q 
 q      q       2 q  . (5.71)

(12) Use the symmetry relation R  R and Eq. (5.70) to obtain the
constraints that are spatial partial differential equations for the
orientation potentials, with  ,  ,  ,   :
Deconstructing the Theory 207

           


2    2   2  
2   2   2 
                 
                   0. (5.72)

(13) Show that multiplying Eq. (5.72) by m  gives (  ,  ,  ,   ):

                   
2    2    2 
            2     2    0. (5.73)

(14) From the form of the energy momentum tensor Eq. (3.74)
demonstrate that:

 ,    , h R  R  2 . (5.74)

(15) Use Eq. (5.74) and the contracted forms for the curvature tensor to
show with  ,  ,  ,   :

      1 2  2  1 2     
 1 2        1 2    
 1 2             . (5.75)

(16) Use Eq. (5.75) to show that Eq. (5.13) and the “trace” of Eq. (5.17)
are consistent.
(17) Derive Eq. (5.24) from Eq. (5.23).
(18) Prove that if the “curl” and “divergence” are known, then the fields
are determined. Use the Cauchy-Kowalewsky existence theorem
[Courant & Hilbert, 1962, p. 39].
(19) Check to see whether the conditions Eq. (3.71) in the co-moving
orthonormal coordinate basis provide any new restrictions. Since we
have already checked each of the symmetry conditions Eq. (3.72),
use the fact that we only need to check the cyclic tensor Eq. (3.73)
208 Geometry, Language and Strategy—Vol. 2

and derive the forms below to show that no new constraints result,
with  ,  ,  ,  ,    :

X                    


                         
            
2         2       
2        , (5.76)

1
2 X   q   q   q             
                  
         . (5.77)

(20) We have a complete set of equations that appear to have one


additional constraint that has not been dealt with, Eq. (5.56).
Show (or disprove) that this equation is identically satisfied as a
consequence of the field equations.
Chapter 6

Steady State Harmonics

Our starting point for decision process theory is Eq. (4.1). If we adopt the
centrally co-moving hypothesis from Vol. 1, then in the (holonomic)
normal-form coordinate basis, there is a commuting Killing vector
whose active1 vector field components K a  V a corresponding to the
time isometry are proportional to the energy flow; the inactive strategy
components are zero K j  0 . This means the centrally co-moving frame
is holonomic and the partial differential equations will be independent
of time.
We can solve these equations and then use the harmonic gauge
equations to return to the normal-form coordinate basis and enforce the
initial boundary conditions. An alternate, more restrictive and more
explicit approach is based on the player fixed frame model in Chap. 5.
We obtained partial differential equations, many but not all of which
were independent of proper-time in the orthonormal co-moving frame. If
we impose the additional condition that in this frame, all the orientation
potentials are proper-time independent, we can show that we adhere to
the centrally co-moving hypothesis. We suggest that this hypothesis may
be of longer term value in identifying interesting solutions of the theory.
As we suggested in the last chapter, significant insight is obtained in
the study of electromagnetic fields by starting with a study of electro-
static and magneto-static phenomena, in which all of the fields are steady
state or independent of time. The motivation is that real world systems
often have this property. The success of game theory using equilibrium
solutions suggests that decisions may share the same property. Therefore

1
By active we mean both time and the active strategies.

209
210 Geometry, Language and Strategy—Vol. 2

it makes sense to add this requirement to our player fixed frame model.
Unless stated explicitly otherwise, we make that assumption in the
remainder of this volume.
Because there are currents and magnetic fields, the approximation is
steady state. We emphasize that we are not assuming that the flows are
zero, just steady state. There are currents for example, associated with
each player. This is the analog of AC behavior in electrical circuits as
opposed to DC circuits in which there are no currents. These electro-
static and magneto-static solutions can be extended to cover features that
appear in fully dynamic systems. Wave phenomena are added by
extending these steady state solutions to harmonic solutions (phasors)
that can be solved using the same techniques as used for DC solutions. In
this way, more realistic dynamic solutions are built as superpositions of
harmonics.
The resultant solutions will be for any number of active or inactive
strategies and any number of players in the player fixed frame model in
which the orientation potentials in the co-moving frame are steady state.
The resultant harmonics might be termed electro-gravitational waves, as
the energy can move back and forth from the electromagnetic field to the
gravitational field.
The models we obtain are expressed as coupled partial differential
equations, albeit with a rather large number (potentially hundreds) of
variables, which admit complete numerical solutions. We are able to
obtain solutions if we meet certain criteria. It is a general theorem
[Courant & Hilbert, 1962] that when these criteria are met, there are
well-behaved solutions. That being the case, we can then in later chapters
solve these equations for interesting decision processes. Since
mathematical existence proofs are not the same as practical methods for
solutions, we have to demonstrate both that we meet the appropriate
criteria and that we can indeed carry out this program in practice. We
have to carefully specify all of the equations and constraints and evaluate
whether solutions are not only feasible but practical.
This chapter derives a harmonic wave equation from the player fixed
frame model that allows us to obtain solutions in the normal-form
coordinate basis. With an eye to finding a consistent set of solutions, the
model is modified, Sec. 6.1, so that the orientation potentials are steady
Steady State Harmonics 211

state, though still allowing for the possibility that the harmonic wave
equation produces active coordinates that are not. In Sec. 6.2 we consider
streamline solutions.
We discuss in detail in Sec. 6.3 a model for the player currents that
represent the energy momentum stress tensor for the constraint effects
that may drive solution behaviors. We say more about the inactive
components in Sec. 6.4 and the active components in Sec. 6.5. We then
obtain the frame-wave equation in Sec. 6.6 in the Co-Moving
Orthonormal Basis. We establish in Sec. 6.7 that there is a central co-
moving holonomic frame in which the metric is independent of time. In
that frame, the central time and hence the class of models considered
satisfies the centrally co-moving hypothesis of Vol. 1. For the fixed
frame model, in the central holonomic frame the active flow components
are zero, Ex. 19, Eq. (6.70).
For reference, we provide a special case of the model when there is a
single active strategy, Sec. 6.8. We use this model for two interesting
examples later in the book: see the Prisoner’s Dilemma (Chaps. 7-9) and
Robinson Crusoe economics (Chap. 10).

6.1 Steady State Scalars

We make the steady state hypothesis that in the co-moving orthonormal


coordinate basis, in addition to the assumption of the player fixed frame
model Eq. (5.1), the proper-time derivative of all the co-moving
orthonormal frame orientation potentials, acceleration and inertial fields
are zero. In other words, along the streamline, none of the potentials
change value. We are in a steady state orthonormal frame, though we
may need to add a few more assumptions to insure that we have a
consistent set of equations with a non-trivial set of solutions.
Based on Exs. 1 and 2 at the end of the chapter, since these scalar
fields are independent of proper-time, the differential operators for
proper-time and proper strategy mutually commute when operating on
such scalars. We can then treat the differential equations for the
orientation potentials using standard techniques for partial differential
equations. We go through Table 5.1 through 5.6 to isolate the equations
212 Geometry, Language and Strategy—Vol. 2

that are relevant. Not every equation is a differential equation; some are
algebraic. The algebraic equations require special attention, which we do
in Secs. 6.3-6.4, which leads to the assumption that the shear components
 are zero (which we equate to very large conductivity) and the player
bonding tensors   are diagonal in the inactive indices.
We start by deciding whether the equations have solutions at all. With
this in mind, in this section for Table 5.1, we temporarily assume that
the active transformations are independent of time. The exactness
conditions Eqs. (4.60) and (4.71) reduce to:

 E a   q  2  e  E a
 E a   q      E a . (6.1)
 E a     E a  2    e      E a

These equations have solutions. Because we have the divergence and


curl of E a , we can solve for these fields, Sec. 5.6, Ex. 18. The gradient
equation for E a also has a solution (Cf. Sec. 5.6, Ex. 18). We deal next
with the acceleration fields.
In Table 5.4, based on the steady state hypothesis, we take both the
electric and magnetic fields to be constants in proper-time. Taking the
electric field independent of time we get the divergence of the magnetic
field:

   q  2            p . (6.2)

We impose the constraint that the magnetic field is independent of


proper-time, analogous to Kirchhoff’s Law, from Eq. (5.53):

    q  q       . (6.3)

This is more complicated than Kirchhoff’s law since the curl is not zero.
The curl depends on the acceleration as well as the player bonding
tensors. Because we assume all fields are independent of proper-time, the
time dependent Eq. (5.46) is satisfied identically. This is an example of
the self-consistency issue we raised: we have to satisfy all of the
Steady State Harmonics 213

equations identified in the last chapter. The assumption of the steady


state hypothesis is a great help here.
The tidal fields simplify as well. From Table 5.5, we see that the time
independence of  is consistent with the time independence of the
electric field Eq. (5.45). We obtain a curl condition on the acceleration
from requiring the tidal magnetic field to be independent of time,
Eq. (5.39):

  q   q   4          . (6.4)

This equation along with Eq. (5.26), Table 5.6 determines the differential
equations for the acceleration along the active directions.
There is one equation that is not identically satisfied, Eq. (5.30):

      2   2   2   2  


. (6.5)
         2  2  2  2 

We have written the independent symmetric and antisymmetric parts that


must be satisfied. If the electric field   is zero, we see that the player
bonding fields  have zero “curl” and form mutually commuting
matrices for each active strategy  .
Finally we look at the equations imposed on the inertial forces, Table
5.6, Eq. (5.54), which imposes the condition:

 p  0 . (6.6)

We need more information to impose further constraints, though we note


that this equation is identically satisfied if the electric field  is zero.
For a fixed current, the electric field is inversely proportional to the
conductivity, Eq. (5.64), so a small electric field implies a very large
conductivity. Decisions act like a system in which changes toward
equilibrium occur very quickly, which is similar to a conductor. In order
to better understand this issue as well as the dynamic behaviors, in the
next section, we consider the broader context in which the active
transformations are not independent of time.
214 Geometry, Language and Strategy—Vol. 2

6.2 Streamline Solutions

In general the frame transformations E a E a  depend on proper-time


through the harmonic gauge condition, Eq. (4.71). We propose and here
define dynamic streamline solutions as follows. Because the scalar fields
are independent of proper-time, the coefficient fields in the exactness
differential Eq. (4.60) and in the harmonic gauge condition Eq. (4.71)
are independent of time. We generalize the solutions Eq. (6.1) by
expanding x a , the streamline proper-time scalar field, in powers of  n
with coefficients X n a   that are scalar functions that depend only on
the proper-strategy scalar functions y (see Exs. 3 and 5):

x a  ,    X n a   n . (6.7)
n

This expansion provides time dependent tangent vectors E a and E a


(Ex. 7) that satisfy the commutation rules (Exs. 8 and 10) as long as we
meet appropriate constraints (Ex. 9).
We motivate this approach using the definition of integral curves that
exist for any vector field Z a . The integral curve is defined by specifying
that its tangent vector through a point is set by the vector field through
these first order differential equations:

dxa
 Z a  x  s . (6.8)
ds
These coupled first order equations have a unique solution and provide a
definition of the coordinate s associated with the vector field. The
coordinate s by its definition moves along the streamline and thus
provides a path dependent definition of proper-time. We refer to this
simply as the proper-time, with the understanding that the path is along
the streamline.
We apply this concept to the orthonormal set of vectors and conclude
that at each point there will be integral curves associated with each
orthonormal vector. In particular there will be an integral curve
associated with the flow that defines the proper-time (a streamline):
Steady State Harmonics 215

dxa
 x  E  b x 
a b
 E a  x    .
a
(6.9)
d
The scalar field  is the measure defined along the integral curve with
  1 (Cf. Exs. 5-6).
We compute the tangent vectors from the holonomic scalar function
x a  y ,  using the result of Ex. 6 in terms of the characteristic vector
potential (or simply characteristic potential) a :

E a    x a
. (6.10)
E a   x a   2a     x a

We use these expressions as the basis for determining the holonomic


scalars in terms of the proper active strategies and proper-time. We note
that the time component can initially be set as E t   1 . If we do that,
then the second equation at   0 shows that the initial values of E t are
determined by the characteristic potential2 a and the variations of the
time field  t :

E t   t   0  a . (6.11)
 0

The characteristic potential thus provides the initial values of this frame
transformation. If the “curl” of the vector field is given (next paragraph),
then different frames would be distinguished only by the divergence of
the vector field:

  a    a    . (6.12)

The vector field  is introduced below in Eq. (6.14). Since decision


process results are independent of frame, we are free to choose this
arbitrary scalar field to be zero,   0 :

  a    a   0 . (6.13)

2
The potential here is proportional to the analog of the gravitomagnetic potential of
general relativity.
216 Geometry, Language and Strategy—Vol. 2

This provides a covariant definition of the frame.


To show that solutions exist, we turn to the power series expansion
Eq. (6.7). In the expansion, Eq. (6.7), we specify the initial values of the
coordinate x a along the surface specified by   0 as well as the initial
flows E a  . We argue that the expansion in terms of powers of the
proper-time scalar function  can be solved in terms of these initial
conditions. We require that the tangent vectors E a  and E a satisfy the
commutation rules for the derivatives, Eq. (4.60), which we leave to Exs.
8 and 10. We find that the gradient of the scalar field  along the proper
strategy direction is determined (Exs. 5, 6, 9, and 13), which in addition
to the gradient condition Eq. (6.13), imposes a condition on the “curl” of
the characteristic potential a :

  2a  


  q  2  e   q . (6.14)
  eq a       eq a   e q    e     

We define the term seasonal potential to be3

A  2a e q  2a exp  q  .

Equation (6.14) is a striking result since it shows that the gradient of


proper-time is not zero but is linear in proper-time along the streamline.
The proportionality constant is given by the acceleration and product of
the charge times the electric field (the “force” due to the electric field).
The coefficient 2a eq acts like the potential for the magnetic and tidal
fields and is determined by them. Our choice of “gauge” Eq. (6.13) is
 
 e q a  0 . Finally we see that the curl of  must vanish. We look
for solutions that enforce this condition and are consistent with the field
equations.

q
3
As a subtle reminder, the exponential function is written as e using the “function”

font to distinguish it from the variables e that represent the player charges.
Steady State Harmonics 217

6.3 Models for the Player Current

We show that the conditions on the “curl” of  are related to the


question of a model for the player current p . The scalar field  is a
known function of the acceleration, electric field and charge, so its “curl”
is determined. To require that the “curl” is zero imposes a condition on
these fields.
We will enforce the condition that the “curl” of  , Eq. (6.14), is zero
by assuming that the electric field (in the co-moving frame) vanishes, a
condition that does not violate any of the field equations:

  0 . (6.15)

This condition becomes part of our steady state hypothesis. We


motivated this assumption earlier by suggesting that for finite currents
p , we have very large conductivity. The condition on the “curl”
follows from Eq. (6.57), that gives   q , and Eq. (6.4) that implies
that the “curl” of the acceleration is zero,  q   q   0 .
This argument for the conductivity is based on a physics analogy and
is an example of the non-static behaviors of the centrally co-moving
hypothesis, Vol. 1, in which time is inactive (Ex. 24). A non-zero shear
 is expected to give rise to a non-zero stress p , roughly
proportional to the shear (in fluids, the proportionality scalar is the
viscosity, Eq. A.8, Vol. 1); in this case the proportionality scalar is also
the conductivity. The rate of change of the energy density is then
proportional to   , which must be zero if the energy density is
steady state, Eq. (6.6). This implies that each of the shear components is
zero. If the shear components (electric field components) are zero, then
the coefficient   q is the acceleration vector and from Eq. (6.4) the
“curl” of the acceleration vector is zero. Because the gradients of the
acceleration are not zero, Eq. (5.61), we have a non-zero acceleration
field. The acceleration effects are a consequence of our frame of
reference. We believe this hypothesis is useful because we focus on the
vorticity of the solutions, which we believe to be a significant new
feature of decision process theory.
218 Geometry, Language and Strategy—Vol. 2

We note that our argument does not logically require that the stress
components (player current) p  be zero. We insure that the energy
density is steady state Eq. (5.54), if we require the shear components to
be small (very large conductivity). The corresponding stress components
need not vanish. For each (proper) player  , the stress components p 
represent their view of the player passion to exercise strategy  . The
“divergence” of the payoff for this player, Eq. (6.2), shows that the
player passion is the source of their view of the player payoff in the same
way that the electric current is the source of the magnetic field in
physics:

      q   2                   p . (6.16)

For a single active strategy, every term but the last depends on the
payoffs, which must vanish since the payoffs   are antisymmetric in
the active proper strategies, of which there is only one. This implies for a
single active strategy that

p  0 . (6.17)

For two or more active strategies however, it is possible for the player
passion components to be non-zero.
The next question is to determine the appropriate equations of motion
for the player passion. One possibility is the following player ownership
rule for player passions (Cf. Chap. 10): each player is accountable and
hence owns only his or her own strategies   S , where S is the set of
strategies owned by  . By this we mean there can be no player passion
for a strategy that is not owned by player  , or equivalently p   0 for
every strategy not owned   S . We may have players that own no
strategies: we call them dependent players. A player that owns at least
one strategy is therefore a non-dependent player. To insure that these
distinctions carry a meaning over the space and time of the decision
process, a sufficient assumption is that the “curl” of the player passion is
zero. Then there would be a coordinate surface whose normal is aligned
with each owned strategy. All such surfaces would be in the subspace the
player does not own.
Steady State Harmonics 219

In the next section we show that such equations simplify further


because the player bonding  can be transformed to a diagonal matrix
using a constant transformation. That being the case, we see that the
boundary condition for a dependent player  , is that p   0 on an
initial surface for every strategy  and so must propagate to zero
everywhere. If a player starts with no ownership of any strategy on an
initial surface, that remains true at all other points of space in the special
case that the “curl” vanishes.
In a similar way we investigate strategies that a non-dependent player
owns. When the “curl” vanishes, the integral curves that result from Eq.
(6.18) for each strategy will propagate from the initial boundary surface
to new surfaces defined by the player ownership potential
p : p   p . These potential surfaces act like “coordinates” of what
is possible for the player to influence through her player passion.
Orthogonal to these surfaces will be the strategies that the player does
not own.
Notwithstanding such features, we find these assumptions not totally
compelling and overly restrictive. The fundamental concept is that the
player passion is the idiosyncratic view of that player of the stresses all
other players feel and this gives rise to the player’s idiosyncratic view of
their payoff. However, a conclusion that we can draw is the virtue of
finding such coordinate surfaces: in other words we need to determine
whether the player passion in some sense corresponds to a vector field
with zero curl.
The equations that determine the player current are determined using
the “divergence” Eq. (5.55) for steady state stresses, along with a tensor
X   determined by the “curl” from the conductivity model. That
assumption is that the current is proportional to the electric field, Eq.
(5.64), and the equation for that field, Eq. (5.53) is given below,
assuming a constant conductivity  :

 p  p q   p    p


. (6.18)
 p     p  q p   q  p    p      p  X  
220 Geometry, Language and Strategy—Vol. 2

To get small electric field we must go farther and say that the
conductivity is very large. If the conductivity is not constant there will be
an additional term in the second equation.
We see that the “curl” X   need not vanish, though it might be
instructive to consider a model in which it does. We will denote this as
the ownership model, loosely related to the discussion above. There is no
assumption that the electric field is proportional to the player current: we
simply assert that X    0 . This model does not make the conductivity
assumption.
For the conductivity model, there also exists a potential (Ex. 14) that
can be used to define coordinate surfaces without assuming that X   is
identically zero. The potential is found by identifying an integrating
factor for the player passion.
This leaves us with two quite distinct and useful models for the
player passion, both defined in terms of potential fields: the ownership
model and the conductivity model. We will study both. We assert that in
each case these potentials provide a unique and well-defined set of
definitions of player ownership based on the coordinate surfaces defined
by their potentials. The strength of ownership is a measure of the passion
the player brings to the decision. This description includes the
conductivity model where the “curl” does not vanish (Cf. Ex. 14).
We see the connection to ownership as follows. For the pure
strategies the player has control over, she is indifferent to which one
should be picked. The choice can be made on the basis of what makes
the most strategic sense: in game theory this choice is based on an
optimization strategy. We think of the surfaces of constant player passion
in the same way: they describe choices for which the player is
indifferent. The passion vector for each player is normal to her
indifference surface. The verbiage makes some sense because the
opposite of passion would be indifference.

6.4 Inactive Stress Equations

The player bonding strain components reflect the degree to which


players cooperate as opposed to decisions that a player influences
Steady State Harmonics 221

through their intentionality. The assumption of zero electric field


components simplifies the constraint Eq. (6.5) on the tidal stress
components:

       0
. (6.19)
        

The first line is the “curl” equation and says that the player bonding
matrices are derivable from a potential. The second line consists of
algebraic equations stating that symmetric matrices mutually commute.
Symmetric matrices that mutually commute can be put into diagonal
form with a common transformation.
We identify the class of solutions that result from our steady state
hypothesis in which we must have a diagonal form for the player
bonding with a condition on the stresses:

 p     2       


. (6.20)
      0

Since the bond strain matrices have zero “curl”, they can be derived
from a potential,    . We obtain diagonal player bonding
matrices with a strategic viscosity  that is arbitrary and a reduced
pressure tensor   that is diagonal. The divergence condition Eq.
(5.27) then sets the diagonal components:

 p 
    q              ph  h  . (6.21)
 n 1 

For steady state tensors, we use Ex. 1, Eq. (6.53) that implies we can use
ordinary derivatives. The resultant bond strain matrices are diagonal and
are derived from a potential so both equations in Eq. (6.19) are satisfied.
The result can be generalized by transforming the equations using a
constant non-singular transformation. If this same transformation is
applied to the initial coordinate basis, then we can align each inactive
strategy j with a corresponding proper inactive strategy  j . The
222 Geometry, Language and Strategy—Vol. 2

equations for the frame transformations Table 5.1 then maintain that
choice:

 E j    E j  . (6.22)

For the inactive strategy    j , the coordinate system exhibits


compression or expansion, set by the diagonal components of the player
bonding matrix. The measure or value of the choices depends on the
position in space.
The components    j that start at zero remain at zero. This
provides a useful interpretation of the inactive metric Eq. (4.70) for
distinct players j  k :

 jk  E j Ek  h  E j Ek  E j Ek . (6.23)

The inactive metric represents the degree of cooperation between the two
players and is equal to the overlap of the player interest flows (player
flows or charges). This is after making an appropriate rotation of frames
to identify the “diagonal” players.
We achieve additional insight by looking at the player bonding
shear equation defined by subtracting from the player bonding tensor a
multiple of the diagonal matrix so that the resultant bond shear tensor
has zero trace (Sec. 8.10, Ex. 3):

      q              1
ni h     . (6.24)

We see that the player bonding      h  , acceleration


q   q q   q , strategic viscosity  and reduced pressure tensor
components   determine the shear. The complexities of the other
tensor fields show up explicitly in the expression for the player bonding
Eq. (5.37):

      1 2      3   3 2   


ni  1
 1 2       1 2    . (6.25)
ni
Steady State Harmonics 223

We have a “gravitational” equation for the player bonding potential  .


By its definition, the player bonding potential, written as lnV , describes
the change of a volume element V . As we move in the space of the
active strategies, this volume element changes; the more compression,
the smaller the volume and conversely. This is clearly a property of the
geometry of space in decision process theory, not an analogy.
The positive sources for player bonding are the energy density, which
is determined by the pressure, the magnetic fields (and the electric fields,
but we take these components to be zero), the bond shear components
and the player bonding gradients. The negative sources, which provide a
type of anti-gravity, are the tidal player interest fields and the tidal
vorticity fields.
When the energy density    p is proportional to the average
pressure, the size of the average pressure is set by Eq. (5.38):


  h  
 h  1  h    p
 n 1 
 
      q   2    1 2  
   1

 1 2      3   3 2      0. (6.26)

If there is at least one active strategy, the coefficient of the average


pressure on the left is positive. For the pressure to be positive, the terms
on the right must sum up to be positive as well. The average pressure is
thus determined by the scalar potentials that are determined by the field
equations and functions we can set by model assumptions, the viscosity
vector  and the reduced pressure components   .
Finally, it is worth noting that for a single active strategy, we need not
impose the constraint Eq. (6.20) on the stress since the “curl” and
commutator relations Eq. (6.19) are satisfied identically. We can choose
to make the forms of either the strains or the stresses simple.

6.5 Active Stress Equations

We are free to choose the viscosity vector  and the reduced pressure
components   . We compute the average pressure from Eq. (6.26) and
224 Geometry, Language and Strategy—Vol. 2

the player current p  from Eq. (6.18). We show in this section that the
field equations determine the remaining stress components, which are the
active stresses p  . We use the fact that the acceleration vector q has
zero “curl” (Sec. 6.3) so the vector field is the gradient of a scalar field
q   q . Furthermore, the divergence is determined by the scalar fields
and the average pressure:

  q  q q  q         2 


  p
    p  . (6.27)
 n 1 

We assume solutions satisfy the Cauchy-Kowalewsky existence theorem


[Courant & Hilbert, 1962, p. 39].
In addition to the field equation for the divergence of the acceleration
vector, we have a more general set of relations involving the gradients,
Eq. (5.26), which demonstrate that the acceleration vectors, along with
the other scalar fields determine the active stress components (but see
Sec. 5.6, Ex. 20):

 p      q   q  q             


2     2    2      
  p
  p   h  . (6.28)
 n 1 

We have moved significantly away from the concept of a perfect fluid,


Eq. (5.62). In any given solution we need to verify that the stresses make
sense. The price we pay for having a relatively simple structure for the
strains and vorticity components in the player fixed frame model is that
the energy-momentum stresses deviate from the ideal fluid. The
deviations are mandated by the presence of non-zero payoff fields and
player interest fields, as in the expression Eq. (6.20) for the inactive
stress components. These internal fields would not be expected to have a
perfect fluid behavior. We conclude from this brief analysis that we have
identified the essential equations for the player fixed frame model. They
form a consistent set with what we expect to be non-trivial solutions.
Steady State Harmonics 225

6.6 Frame-Wave Equation

In addition to the steady state scalar equations based on the steady state
hypothesis, we extract a time dependent wave equation that applies to the
player fixed frame model. We believe the result justifies the effort. With
the assumption of zero electric field, the coefficients of the expansion
(6.7) are determined by the harmonic gauge condition Eq. (4.71):

1  e e   E



a
  E a   q     E a  0 . (6.29)

We have incorporated the conditions Eq. (4.60) in the properties of the


proper-time scalar field  . We demonstrate solutions to this equation
using a power series expansion. The basic idea is to expand each term in
Eq. (6.29) and equate terms with the same power:

1  e e   E



a
 1  e e   X s  2 a  s  2  s  1 s ,
s 0

 q  
 
 E a
    q      ,

    X s a s   q X n a s s   2a X s 1a  s  1 s . (6.30)


s 0 s 0 s 0

The divergence (which is the ordinary derivative when acting on steady


state fields) of the tangent vector E a has several terms:

 E a     X    q  X
n 0
s
a s

s0
s
a
s s

  2a  X  s  1 s 1
a s

s 0

    q X s a  s s   q q X s a s 2 s
s 0 s 0

  2a q X s 1  s  1  s
 a 2

s 0

    2a X s 1a   s  1 s


s 0

  2q a X s 1a  s  1 s s
s 0

  4a a X s 2 a  s  2  s  1 s . (6.31)


s 0
226 Geometry, Language and Strategy—Vol. 2

In the sums we changed the summation index in order to identify the


common powers.
As an example of the process, we start with arbitrarily picking the
first two functions X 0a   and X 1a   . Based on matching powers of
 we obtain the next term:

2 1  e e  4a a  X 2 a    X 0a   q      X 0 a


  6a q  2a    X 1a  4a  X 1a . (6.32)

We see that the second term in the series is determined by the first two
terms: the position and the initial value of E a . Though more
complicated than solving differential equations with constant
coefficients, it is clear that the same pattern maintains. The coefficient
X a s  2 will depend on the functions X a s , X a s 1 and their gradients. The
general case is Ex. 16. This demonstrates that given arbitrary fields
 
X a 0 X a1 , all other fields X a n are determined.
The power series demonstrates that there are solutions to the frame-
wave equation Eq. (6.29) based on given initial conditions. We expected
this based on general theorems on elliptic partial differential equations
(Courant & Hilbert, 1962). To summarize, we write the wave equation as
a partial differential equation (Ex. 17):

1  e e


 4a a  4a q  q q 2   2 x a
  x a  2  2a  q    x a
  2 a  2  a  4a q  q     x a
   q  2 q q   x a   q      x a  0. (6.33)

This differs from the usual wave equation because of the time and
strategy dependence of the coefficients. These differences make it
impossible to find solutions that factor into a function of proper-time and
a function of position. We address this problem in the next section.
Nevertheless, we can construct general solutions from linear
combinations of specific solutions. As a step to identify specific
solutions that might be of interest, note that any solution will be
Steady State Harmonics 227

generated once we specify the boundary conditions on a timelike


hypersurface   0 . To make the discussion explicit (Cf. Chap. 14), we
imagine that there are three active (proper) spatial directions
  x y z , though the results hold for all cases of one or more
active strategies. Based on the boundary conditions we then have the
active coordinates x a  x, y, z,  at all space time. In particular we have
the active coordinates along the spacelike hypersurface z  0 . We can
decompose any general function of the remaining three variables
x, y,  into a Fourier or harmonic series in  at each transverse point
x, y . This might be especially insightful if the behaviors in the space
directions x, y are periodic for example.
An analogy from electrical engineering would be a wave guide in
which the transverse space direction behaviors are strongly influenced by
the geometry of the guide. The behavior along the time direction can be
resolved into a continuum of harmonic terms starting with a zero
frequency (linear) contribution of the form U  V  , and including sine
and cosine terms, sin and cos , respectively. A good deal of
understanding of the general solution can be obtained based on
expectations of the behavior of the individual harmonic contributions.
Conversely, we can start with any harmonic contribution along the
spacelike hypersurface z  0 , taking specific harmonics for the
transverse directions such as sin mx sin my and multiplying by the time
harmonic such as sin . We expect that with suitable boundary
conditions, there will be a solution for each of these harmonics and that
the superposition of such harmonics will allow us to reconstruct the
general solution.
Though the two approaches should be equivalent, they may differ in
their ability to deliver accurate results using numerical approximations.
For example, the successive terms in Eq. (6.67) for the power series
solutions (Ex. 16) involve gradients of previous fields. We expect small
errors to grow as we compute higher order terms. We will show below
that a consideration of harmonics can be based on polynomials with the
higher order coefficients set to be small (zero) with the possibility of the
lower order terms being computed without the same loss of accuracy (see
Ex. (18)). Based on these considerations, we examine harmonic
solutions, emphasizing that the harmonics are the usual Fourier series on
228 Geometry, Language and Strategy—Vol. 2

the hypersurface z  0 , but away from that hypersurface the behaviors


follow from the full partial differential Eq. (6.33).
The zero frequency harmonic solution of Eq. (6.33) is linear in
proper-time:

x a  U 0 a  V0 a . (6.34)

The coefficients are functions of position. Using this as a trial solution,


we obtain two coupled partial differential equations for the coefficients:

 U 0 a   q     U 0 a


4 a V0 a  2   a    a  2 a q V0 a 0,
  V0   3q  
 a  
 V
  0
a

  q     q  2 q q V 

0
a
 0. (6.35)

This generalizes the approach we started with in Eq. (6.1).


It suggests to us the following approach. Start with a solution that is
constant in time, linear in time plus a superposition of harmonic
polynomials of degree N . The harmonic polynomials generalize the
notion of phasor solutions. In the limit that the degree N goes to
infinity, assuming that the limit is sufficiently well behaved, we obtain
steady state harmonics. As mentioned, because of the complexity of the
differential equations, these harmonic polynomials will not in general be
a simple factor of e i away from the initial spacelike hypersurface,
which we have chosen to illustrate above as z  0 for three active
strategies. In the next section, we will transform to a frame in which we
obtain phasor solutions defined in the usual way. First however we
complete the analysis in our current frame of reference.
Sticking with this illustrative example, a phasor solution on the
surface z  0 would be linear combinations of sin and cos
multiplied by harmonics V a  x, y  and U a  x, y  respectively for the
transverse strategies along x, y . We can approximate these harmonic
functions with harmonic polynomials of degree N by taking the real and
imaginary parts of the exponential power series expansions of the
Steady State Harmonics 229

complex phase e i and truncating after N terms for


0  x, y , z  0 :

 i   i 
N s s

sin N    Im   sin N  s  Im


s 0 s! s!
. (6.36)
 i   i 
N s s

cos N    Re   sin N   s  Re


s0 s! s!

These power series set the values of the coefficients on the surface  0
by requiring X N 1a  X N  2 a  0 for polynomials PN a  ,  , Ex. (18). So,
on the hypersurface z  0 , we have:

 i   i 
s s

X a
s 0   V a
0  Re  U a
0  Im
s! s!
Y a
s 0    X s  0 
a
. (6.37)
 i   i 
s s

 X a s  0   V a  0  Re  U a  0  Im
s! s!

We have converted to first order partial differential equations using


Y a s     X a s   .
Away from the initial surface, the coefficients will evolve and in
general will no longer reflect a pure decomposition into a single
harmonic. We are led to this complexity because the harmonic shapes
change with z because the coefficients of the partial differential
equations depend on the proper strategies  . A similar complexity arises
in electrical engineering when there are damping effects that depend on
z , which arise from reactance contributions.
Any solution of the partial differential equations can be written as
superposition of these extended harmonic polynomial solutions. In the
limit that the number of polynomial terms goes to infinity, we write any
solution as a superposition of harmonics. In electrical engineering, the
overall time dependence can be constructed from a superposition of
phasor solutions that have defined time dependence  sin , cos 
and a computed spatial dependence. These solutions help characterize the
behaviors expected. For the wave Eq. (6.33), we view the harmonic
230 Geometry, Language and Strategy—Vol. 2

polynomial solutions as providing the analogous insight. For any number


of active strategies, any superposition of appropriate harmonic
polynomials will again be a solution.
The initial values on the hyperplane will be:

xa 0 ,   U0a 0   V0a 0 


U a 0  cos N   V a 0  sin N   . (6.38)

There is a similar expansion for the gradient:

 x a 0 ,   U0a 0   V0a 0 


  U a 0  cos N   V a 0  sin N   . (6.39)

Each harmonic polynomial coefficient will evolve to a different spatial


dependence, one that is determined from the differential equations. We
set the proper-time behavior on an initial surface in (proper) space and
the remaining spatial dependence is determined. An analysis of a
complete set of harmonics will be equivalent to an analysis of the general
solution to the partial differential equations.
We have identified a complete set of coupled ordinary partial
differential equations (6.69) that fully represent the decision process
theory field Eq. (4.1) in the co-moving orthonormal coordinate basis.
These can be solved numerically using known techniques [Courant &
Hilbert, 1962]. We will use the numerical method of lines, [Wolfram,
1992] in Chap. 14. However, this is not always effective for elliptic
equations, which characterize our steady state equation for two or more
active strategies. When there are several inactive strategies, we note that
solutions to these coupled partial differential equations may benefit from
lattice techniques such as finite element analysis [Bhatti, 2005]. In any
event, the harmonic polynomials are extended in a practical way to
functions of both the proper strategies and proper-time,  ,  .
Steady State Harmonics 231

6.7 Centrally Co-Moving Hypothesis

The above analysis is in a co-moving orthonormal frame. We transform


to the central holonomic frame in which central time    e q (Ex. 19)
and proper strategies form the (non-orthonormal) holonomic basis and in
which the fixed frame model is co-moving for the active strategies. The
resultant coefficients of the frame-wave equation are independent of the
central time (Ex. 32):

    x a  1  e e  4a a  e 2 q  2 x a


  q      x a
4a e q   x a  2  q     e q a  x a  0. (6.40)

This wave equation has the feature that linear combinations of solutions
are again solutions: the equations are linear in x a .
In general, solutions to these equations will reflect oscillations as well
as attenuation (and growth). Such behaviors can be studied by
considering phasor solutions that correspond to a fixed frequency
x a  e i , Ex. 33. We call such solutions frame-waves. Even though the
differential equations are equivalent, these bounded solutions will in
general not be equivalent to the harmonic polynomial solutions, which
can lead to unbounded behaviors for the frame transformations.
We can then solve Eq. (6.40) with harmonics (phasors) that are
independent of central-time, Exs. 33-34, where x a  x a ei :

    x a     q  4i eq a   x a


  2 e2 q 1  e e  4a a   2i eq a    q  x a   0 . (6.41)

We obtain solutions in terms of complex numbers; we superpose


solutions with appropriate boundary conditions and take the real part to
obtain our final answer. Numerically, either this approach or the one
from the previous section gives the same answers. If we have the
solution for one such partial differential equation, we can use the
defining equation for the proper-times    e q to get the other. For
232 Geometry, Language and Strategy—Vol. 2

example, we see that the proper-time behavior of a steady state wave in


the central holonomic frame will have a somewhat different behavior in
terms of central time:

sin  sin  eq  . (6.42)

The apparent frequency changes depending on the streamline for


constant values of the proper-time  because of the variation of the
potential q that determines the total acceleration.
Breaking our solution down into harmonics provides us necessary
and practical insight into the behaviors of decisions in decision process
theory. The limitations are the usual sort and are based on the maximum
number of terms practical for numerical calculations. If we work in the
approximation of a finite number of polynomial terms for Eq. (6.36), or
using the central holonomic frame Eq. (6.41), we nevertheless expect to
be able to ascertain (possibly damped) wave phenomena in our solutions.
We expect to observe them more directly in Eq. (6.41). At any point in
space, we expect to see harmonic behavior. The motion of the peaks
(valleys) defines the wave motion, which we expect to propagate at the
maximum speed allowed by the theory (corresponding to the speed of
light in physical theories). If we create a pulse, then we would expect to
see that pulse propagate and possibly dissipate over time and/or space.
The number of independent waves will depend on the number of active
strategies. In particular, we expect to see wave phenomena even for a
single active strategy. In the next section, these general results for any
number of strategies will be applied to a single active strategy.
We expect that the phasor solutions for decision process theory will
have properties somewhat different from those in electrical engineering.
In decision process theory, each phasor solution corresponds to a
different set of initial conditions. To get some guidance on what to
expect, we refer the reader to Hawking & Ellis [1973, p. 231] on
solutions to equations in general relativity, a prime example of the type
of theories that we are considering, which is a gauge theory based on
differential geometry. Distinct solutions to the frame-wave equation in
general correspond to distinct physical solutions, ones that are not
diffeomorphic.
Steady State Harmonics 233

In order to get a unique solution for the metric in the normal-form


coordinate basis, one specifies the initial conditions on a hypersurface
orthogonal to the direction of time. Solutions are considered to be
identical if there is a diffeomorphism between the two solutions: such
transformations are related to each other by a gauge transformation.
When there is such a diffeomorphism, the two spaces are essentially
identical. By fixing the gauge, such as with the above frame-wave
equation (the harmonic gauge condition), one obtains a unique solution
from the set of all such identical solutions.
So far this sounds like any physical theory; what is not expected is
that two solutions with different initial conditions may not be
diffeomorphic. The rather dramatic consequence of this is that solutions
in the central holonomic frame will not be diffeomorphic to solutions in
the normal-form coordinate basis. Further, solutions in the normal-form
coordinate basis with different initial conditions, including phasor
solutions, will not be diffeomorphic. Thus, it is evident that in the central
frame there are no “gravity waves” and no “electromagnetic waves”.
However, it does not follow that such analogs are absent in the normal-
form coordinate basis. In our numerical solutions, we will inquire
whether such analogs are present. We require first that the waves be
present as traveling frame-wave solutions in Eq. (6.40). We then have to
see evidence that such waves are also present in the metric, for gravity
waves, or in the payoff field, for electromagnetic waves. The gauge
theory constraints from the general theory apply when counting the
number of independent components.
Thus the solution to Eq. (6.40) provides the transformation, though
not a diffeomorphism, between two holonomic coordinate bases: the
normal-form coordinate basis and the central holonomic frame
coordinate basis (Sec. 4.6). A phasor solution can be thought of as
reflecting a rotating basis. The rotation effects are transmitted from one
point in space to another as a wave that is possibly attenuated. Since
these effects are present in the coordinate transformation, we should see
these effects in other tensor quantities such as the metric and curvature of
space-time.
234 Geometry, Language and Strategy—Vol. 2

6.8 Single Active Strategy Dynamic Solutions

In realistic solutions with many active strategies and many players, it


may still happen that many of the strategies are effectively not utilized
and yet are behaviors we want to track as opposed to constraints. In this
case the solution is an equivalent decision process with fewer strategies
and more players. For each strategy that is not utilized in the sense that
its strategy is an isometry, that strategy acts like the inactive strategy of a
player, albeit a player that has no active strategies of his own. We will
also refer to such strategies as codes of conduct. The simplest case is that
in which all but one active strategy is replaced with players who are
inactive, but make an impact on the game through their payoffs and other
indirect orientation potentials. We call these single active strategy
solutions.
For streamline solutions with a single active strategy, the partial
differential equations have two variables. The number of field equations
is reduced since there are no magnetic fields   , tidal magnetic fields
  , or “curl” equations, since these are antisymmetric in the active
components. This simplifies the covariant gradients Eq. (6.14):

  1
a  0 . (6.43)
  q

These equations insure that the commutation rules Eq. (4.60) are
satisfied.
We write the partial differential equations that result from the
harmonic wave Eq. (6.33) for the components of Eq. (6.7), where in the
normal-form coordinate basis, t is time, u is the active strategy
direction4 and a  t , u :

4
A similar discussion can be made in the central holonomic frame if we use Eq. (6.40).
See Sec. 6.10, Ex. 51.
Steady State Harmonics 235

1  e e 

 q q 2   2 x a    x a  2 q   x a
  q  

   q  2 q q   x a   q      x a  0 . (6.44)

The single strategy model is ideal for understanding in better detail how
we obtain the harmonic coordinate behaviors from these
transformations. We expect to gain insight by using the harmonic
polynomials PN a  ,  , Eq. (6.36) and the coefficient equations (Ex. 18):

X N 1a  X N  2 a 0,
  X N    2 N  1 q  
 a  
 
 XN a

 N  q     q  q q  N  1  X N a 0,
  X N 1a    2 N  1 q      X N 1a
  N  1  q     q  Nq q  X N 1a 0,
s  N  2, N  3,,0 :
  X s a    2 s  1 q      X s a
 s  q     q  q q  s  1  X s a
  s  2  s  1 1  e e  X s  2 a 0. (6.45)

These equations can be successively solved to obtain all of the


coefficients once the scalar functions q   e   are
determined.
From Table 5.1 we get three coupled equations that determine the
scalar functions, along with equation for the payoffs:

 E j    E j ,
 E j   q E j  2  E j ,
 e   q e  2   e  ,
f j
  f j
 0, f j
 0,
f j
  
 2 q E j    E j  e E j    e E j 1  E k  Ek   . (6.46)
236 Geometry, Language and Strategy—Vol. 2

Given the active transformations determined from x a and the equations


above for the inactive vectors, the metric components are determined,
Eqs. (4.70) and (4.72).
The remaining scalars are given in Sec. 5.4. The tidal player interests
 reflect the vorticity attributes of the solution and are determined
from Table 5.4:

  2 q         . (6.47)

There are no magnetic or tidal magnetic fields.


Because there is only a single active strategy, the player current
p  0 is zero, Sec. 6.3, Eq. (6.17). The tidal forces Table 5.5
determine the divergence of   :

      q       2  
 p 
  p  ph  h  . (6.48)
 n 1 

The inertial forces (using Eq. (6.93) from Ex. 19) determine the active
diagonal stress component:

 p    q     1 2      1 2     . (6.49)

We are free to pick the energy density scalar  and inactive stress
scalars p . The average pressure is determined from the diagonal
stresses, Eq. (3.41).
The inertial forces also determine the divergence of the acceleration,
Eq. (5.61):

  p
 q  q q     q  2       p  . (6.50)
 n 1 

We use this to eliminate the divergence in Eq. (6.44):


Steady State Harmonics 237

1  e e


 q q 2   2 x a    x a  2 q  x a
    p 
  q q  2       p    x
a

  n  1  
  q      x a  0. (6.51)

We thus have sufficient equations to determine all the unknown scalar


fields as well as the harmonic polynomials for any index N (see Ex. 50).
We can also write this in the centrally co-moving frame, Ex. 51:

  x a   2 e 2 q 1  e e  x a     q   x a  0 . (6.52)

This partial differential equation can be solved numerically by standard


techniques. The great advantage is that we don’t need to consider the
proper-time or central time, only the strategic space dependence.

6.9 Outcomes

In this chapter, we have established the general properties of the player


fixed frame model that extend to any model satisfying the centrally co-
moving hypothesis. We established that in the central frame, the
solutions are steady state and that in any other frame, using the harmonic
gauge condition, we can write any general solution as a superposition of
harmonics, in exact analogy to what occurs in electrical engineering.
Thus we expect many of the same general classes of behaviors: DC
phenomena, AC phenomena in the presence of sources (including initial
conditions acting like sources) and resonant phenomena. We numerically
explore the consequences of these expectations in later chapters. We take
up simple scenarios for the Prisoner’s Dilemma (Chaps. 7-9) and
Robinson Crusoe economics (Chap. 10). In Chap. 14, we expand the
discussion to more general two-person games with multiple strategies.
For the purpose of visualization and practical convenience in doing the
numerical computations we focus on two strategies per player. Some
attributes of the theory only show up if there are three or more active
strategies.
238 Geometry, Language and Strategy—Vol. 2

The attainment of the outcomes of this chapter is facilitated by doing


the exercises in the following section. Based on this investment, the
student should achieve the more detailed outcomes below.
 The student should understand the conceptual origins of the harmonic
wave equation and be prepared to apply this equation to decision
processes. From Sec. 6.1, the student should understand the
fundamental model Eq. (4.1). This form is extremely compact and
specifies the basic geometry of the strategic space of decisions in a
complete and self-consistent formalism:

R  1 2 g  R   T .

 From Sec. 6.2, the student should understand streamline solutions in


the co-moving orthonormal coordinate basis.
 From Sec. 6.3, the student should understand how one might create a
model for the player currents that represent the energy momentum
stress tensor for the constraint effects that may drive solution
behaviors. The student in particular should understand the distinctions
between the inactive components, Sec. 6.4, and the active
components, Sec. 6.5.
 From Sec. 6.6, the frame-wave equation is developed and then
expressed in the central co-moving holonomic frame in Sec. 6.7. This
is an important result for the chapter that satisfies the centrally co-
moving hypothesis. The student should also understand that some
generalizations are possible without dramatically changing the
consequences: for example, it can happen that in the central frame in
which time is an isometry, the flow need not be zero. For the player
fixed frame model, however, the active flow components are zero, Ex.
19, Eq. (6.70).
 From Sec. 6.8, the student should understand the special case of
streamline solutions when there is a single active strategy. The
student should see how these equations provide insight into the
decision process. The streamline solutions of the player fixed frame
model provide the first step towards a quantitative understanding of
dynamic behaviors in decision process theory.
Steady State Harmonics 239

 The student is expected to see that the various distinctions and their
related field equations of Part 1 are brought together in this chapter
and provide a basis for harmonic solutions along streamlines. These
solutions are electro-gravitational waves of decision process theory.
On the transverse surface, the resultant equations are partial
differential equations that are mathematically similar to those of
electromagnetic theory. The student should gain an appreciation of
how these equations might be solved.

6.10 Exercises

(1) For any scalar field  , prove that in the co-moving basis,
|   |  and hence for the player fixed frame model:

        2    e        . (6.53)

(2) For any scalar field  , prove that in the co-moving basis, |  |
and hence for the player fixed frame model:

       q  2  e   . (6.54)

(3) Show that for any scalar function    that is a function of the
proper-active strategies y  only, that such a function is independent
of proper-time. In particular demonstrate for each coordinate that

 y  0 . (6.55)

(4) Show that the gradients of proper-active strategies with respect to


proper-active strategies are orthonormal:

  y   . (6.56)

(5) Demonstrate that if there is a scalar function   x  such that


    1 , then we have the following commutation relation and
implications for any arbitrary vector function a    of the proper-
active strategies:
240 Geometry, Language and Strategy—Vol. 2

        q  2  e 


        q  2  e 
   2a  
   q  2  e . (6.57)

(6) Using Ex. 5, show the following relationship between the


differential operators   and  with the partial derivatives   and
  when operating on a scalar function   y  x  ,  x   :

    
. (6.58)
     2a     

(7) Starting with a power series expansion for the coordinate x a in


terms of the scalar function  defined in Ex. 5 and coefficients X n a
that are functions only of the scalar fields y (which are scalar
functions of the harmonic coordinates), use the commutation rule
result (6.57) to show the expansions for the vector fields E a  for
the flow and E a for the active transverse directions, defining the
scalar field   q  2  e :

xa  X  ,
n 0
n
a n

E a   X n
n 0
n
a n 1
,

E a    X 
n 0
n
a n
   X n a n n   2a X n a n n 1.
n 0 n0
(6.59)

(8) Use Ex. 1 to show the first commutation rule of Eq. (4.60) is
satisfied:

 E a   E a    x a    x a   q  2  e  E a . (6.60)

(9) Use Ex. 1 with the scalar  and the assumption that the scalar fields
are static to show that a is constrained by the magnetic and tidal
magnetic fields and that the curl of  must be zero:
Steady State Harmonics 241

        2    e      ,


2  a   2  a  2 a 
2 a           2    e     ,
 a    a   a    a     e    ,
       0. (6.61)

(10) Use Ex. 1 to show the second commutation rule of Eq. (4.60) is
satisfied with the vector field a satisfying Eq. (6.61):

 E a     E a     x a     x a  2    e      E a . (6.62)

(11) Show that the co-moving coordinate basis field equations from Sec.
5.4 for the player fixed frame model satisfy the normal-form
coordinate basis longitudinal flows, Eqs. (2.49) and (2.50),
modified to include the inertial acceleration:

g ab V b   abcV bV c  Vk Fabk V b  1 2 V jVk  a jk


 qa ,
Vk  qk . (6.63)

(12) Show that imposing a gauge condition, such as the following, on the
vector field a in Eq. (6.61) corresponds to a frame transformation,
taking the most general form of  to be     q in terms of a
scalar field:

  a e q   0 . (6.64)

(13) Show that the “curl” of the new seasonal vector potential field
A   2 e q afrom Eq. (6.61) simplifies and determines the
seasonal payoff and acceleration fields. Show that the field
equations for the “curl” of the magnetic fields satisfy these
equations and further, show that with the gauge from the previous
exercise, the vector field depends on the player current in the
second order differential equation shown below:
242 Geometry, Language and Strategy—Vol. 2

  eq a       eq a   eq    e    
 f  ,
 f      f     f   0,
 e q     eq a    e p     e       q  
2   e      
2  q   2   e .
(6.65)

(14) The ownership model, Eq. (6.18) can be solved in terms of a


potential, p   p . Show that this is also true for the
conductivity model, with the addition of an integrating factor. Show
that the integrating factor can be obtained in terms of the
acceleration potential and the transverse player bonding potential.
Show that the gradient term is modified as shown. You will have to
make use of the fact that the player bonding matrix commutes with
its gradient to show that there is a single matrix integrating factor
that works for all stresses p . In the matrix form of the equation
(here we use bold to indicate a matrix), show that the form is as
shown.

p  P   p
 P     q      P 
. (6.66)
λ P  exp  qI  φ 
  p        2    p

(15) Discuss the significance of the source term and using Ex. 14, show
that the passion source term is a scalar multiple of the gradient
 p . So in a sense, this conductivity model gradient is as related to
ownership as is the one in the ownership model.
(16) Verify Eqs. (6.30) and (6.31). Use these equations to determine the
general form of the power series in terms of the initial functions
X a 0 and X a1 :
Steady State Harmonics 243

  s  1 s  2  1  e e  4a a  X s  2 a
  X s a    2 s  1 q      X s a
 s   s  1 q q  q     q  X s a
2  s  1  2  s  1 a q  a     a  X s 1a
4  s  1 a  X s 1a  0. (6.67)

(17) Show that the partial differential equation corresponding to the


wave Eq. (6.29) is:

1  e e


 4a a  4a q  q q 2   2 x a
  x a  2  2a  q    x a
  2  a  6a q
 q     q  2 q q   x a   q      x a  0. (6.68)

(18) Use the power series solution to show that the wave Eq. (6.68) has
polynomial solutions PN a  ,  with X N 1a  X N  2 a  0 and show
that the order of solving for the coefficients starts with the equation
for X N and then equation for X N 1a using this solution and so
forth:

  X N a    2 N  1 q      X N a
 N  q     q  q q  N  1  X N a 0,
  X N 1a    2 N  1 q      X N 1a
  N  1  q     q  Nq q  X N 1a
4 Na  X N a  2 N   a   2 N  1 a q  X N a  0 ,
244 Geometry, Language and Strategy—Vol. 2

s  N  2,,0 :
  X s a      2 s  1 q   X s a
 s  q     q   s  1 q q  X s a
4  s  1 a  X s 1a  2  s  1     2 s  3 q  a X s 1a
  s  2  s  1 1  e e  4a a  X s  2 a  0 . (6.69)

(19) In Ex. 6, we express the frame derivatives    in terms of


holonomic partial derivatives    . In effect we have made a
frame transformation from the co-moving orthonormal coordinate
basis to the central holonomic frame. We include in the definition
of central holonomic frame a redefinition of proper-time    e q
(which is just a scale change along a streamline). Show that the
transformation W   W   W   from the co-moving orthonormal
frame to the central holonomic frame has the following values for
central holonomic frame coordinates     k , where the  
first two coordinates are active and holonomic and the last
coordinate is inactive and non-holonomic:

W    eq W    e q e W    2a e q
W  0 W  0 W     . (6.70)
W k  Ek W k   Ek W k  0

(20) Show that the transformation W  W  W  from the central
holonomic frame to the co-moving orthonormal frame has the
following values and is the inverse of the transformation Eq. (6.70):

e q e  q e
W   W   W   0
1  e e 1  e  e
2a 2a 
W   W   e W    . (6.71)
1  e e 1  e  e
Wk   Ek  Wk   Ek  Wk   0
Steady State Harmonics 245

(21) Using the frame transformations from Ex. 19, show that the
following provide the frame transformation E  from the normal-
form coordinate basis to the central holonomic frame basis. In
addition show that the inverse frame transformation E  is as
indicated below.

E a  e  q E a E a  E a  2a E a E ka  0
E j  0 E j  0 E kj   kj
. (6.72)
Ea  1  e e  e q Ea  2a e q Ea Ea  Ea Ea k  0
E j  0 E j  0 E j k   kj

(22) In the central holonomic frame, show that the inactive metric
elements  jk   jk are equal to the corresponding inactive metric
elements in the normal-form coordinate frame, show that the mixed
metric elements g kb  0 are zero and show that the active
contravariant metric elements g ab have the values given below.
Furthermore, show that the determinant of the active metric
elements are as indicated:

g   1  e e  4a a  e 2 q g   2a eq
g   2a eq g       . (6.73)
det g ab  1  e e  e2 q

(23) In the central holonomic frame, show that the active covariant
metric elements are:

e 2 q 2a e  q
g  g  
1  e e 1  e e
. (6.74)
2a e  q 4a a 
g    g     
1  e e 1  e e

(24) Using the fact from Ex. (23) that the metric elements in the central
holonomic frame are independent of central time  , show that
246 Geometry, Language and Strategy—Vol. 2

central time is indeed central (Thomas G. H., 2006), i.e. it is


inactive and commutes with all inactive strategies. Furthermore
show that the curl of the seasonal potential 2a eq determines the
seasonal payoff

f   2 f   2eq    e     , Cf. Eq. (6.65).

(25) Use the frame transformation to obtain the player payoff matrices in
the central holonomic frame, considered as tensors in the space of
time and active and inactive strategies:

F j  e q f j


F    f
j j
   2a f j
   2a  f j
 . (6.75)
F k  F k  F
j j j
ik
0

(26) Based on the form of the payoff matrix Eq. (6.75) in the central
holonomic frame and using the definition Eq. (4.30) transformed to
the central holonomic frame, show that the payoff matrix for the
active components is the curl F abj   a Abj  b Aaj of some potential
function Aaj by showing:

 a F bcj  b F caj   c F abj  0 . (6.76)

(27) To determine the potential function Aaj implied by Ex. 26, both
the curl and divergence need to be specified. Show that the gauge
condition g ab  a A j b  0 in the normal form coordinate frame, which
we call the harmonic gauge for the vector potential, Vol. 1, can be
expressed in the central holonomic frame by establishing the
following relations, where the covariant derivatives are here
restricted to active components only:
Steady State Harmonics 247

1
g
b  
g g ab A j a  g ab  b A j a  0

g ab A j a ;b 
1
g

 b g ab g A ja  . (6.77)
ab
g A j
a ;b  1
2 g A a  b ln   0
ab j

1
g
b  
g  g ab Aaj  g ab  b Aaj 
1
g
b  
g  g ab Aaj  0

(28) Show that in the central holonomic frame the active flow

components are V   W    e q , V   W    0 and the inactive 
flow components are V k  W k   V k . If we call the covariant-flows
the “momenta”, show that these momenta are

 e q 2a 
V  
, V   
 1  e e 1  e e 

for the active components and Vk  Vk for the inactive components.


This shows that a contributes to the momentum.
(29) Show that in the central holonomic frame the acceleration
components are 
q   2q a eq , q   q . The covariant 
components (the “forces”) are q  0, q  q , qk  0 . These
acceleration components are for the full geometry of active and
inactive coordinates. We can define the active geometry
acceleration in frame in which the active coordinates are holonomic
and in which the inactive coordinates are orthogonal using only the
active acceleration, illustrated for the central holonomic frame in the
first equation below. Show that in that frame, the active geometry
acceleration consists of three contributions, shown in the second
equation below. The first contribution is the competitive
acceleration, the second is the cooperative acceleration and the
third is the absolute acceleration. Show that the active geometry
acceleration in the central holonomic frame is given by the third
and fourth equations. Also show the identity of the active geometry
acceleration as computed from the first equation and computed
248 Geometry, Language and Strategy—Vol. 2

from the sum of the three contributions. In the process show that the
competitive acceleration and cooperative acceleration have the
forms provided in the last two equations. You will also have
demonstrated that the three acceleration contributions are each
independent of proper-time as is the active geometry acceleration in
the central holonomic frame.

Qa  g ab V c  c V b   abc V b V c
Qa  Vk Fabk V b  1 2 V j V k  a  jk  qa
Q  0
q  2 e   e e 
Q 
1  e e 

 2
. (6.78)
2q e e  2  e e  1  e  2 e e
    

Ek fk 
1  e e 
 2

q e e e e   2 e e e   e e 


1 E  E    lm 
l m

1  e e 
2 2

(30) Using Ex. 29 for the active geometry acceleration in the central
holonomic frame and the transformation properties Ex. 21, show
that the active geometry acceleration in the normal-form coordinate
basis is Qa  Ea Q . So in particular the time dependence is
provided entirely by the transformation Ea . Furthermore, show
that the covariant magnitude of this acceleration g ab Qa Qb is
  Q Q  and so a scalar independent of proper-time: like the
pressure p , energy density  and cooperation potentials  jk , the
covariant magnitude is constant along a streamline. Show that for
any scalar  that is constant along the streamline, the gradient is
 a  Ea   and so show that in the normal-form coordinate basis
such scalars will in general be time dependent if there are non-trivial
harmonics.
(31) In our numerical analysis in Vol. 1, in addition to an exactly solved
single strategy model, we computed a variety of examples in which
Steady State Harmonics 249

all strategies were active. In these models, we assumed that there


was a central holonomic frame in which time was inactive and
commuted with all active strategies. For the player fixed frame
model with the steady state hypothesis, we satisfy that assumption
(Ex. 24). We also made several practical simplifications to evaluate
these models: for example we did not use the field equations to
obtain the metric and payoff tensors, but estimated their behaviors;
we averaged the player payoffs, which set the value of the composite
payoff since we assumed the composite payoff was zero and the
frame was fixed; and we assumed that the relationship between
energy density and pressure was that of a perfect fluid. In this
chapter with the player fixed frame model with the steady state
hypothesis, we have removed those simplifications and estimates
and provided exact solutions. In particular we have explicitly kept
each player’s inactive strategy (no common inactive strategy) and
didn’t assume that the composite payoff is zero. Show that the
normal-form coordinate basis behaviors seen from that initial
numerical analysis now follow directly from the free fall behavior
of E a (Cf. Sec. 18.9). Show that the gradient of the pressure and
the gradient of the cooperation potentials  jk are computed from
Ea (Ex. 30).
(32) Show that the wave Eq. (6.29) in the central holonomic frame is
transformed from the partial differential Eq. (6.33) to the following
partial differential equation in which the coefficients are all
independent of the central time  :

1  e e


 4a a  e 2 q  2 x a        x a
4a e q   x a  2e q a  q      x a
  q      x a  0. (6.79)

(33) Show that one can apply the phasor approach directly to the partial
differential equation in Ex. 32, namely one can superpose solutions
of the type x a  x a ei where the phasor component x a satisfies
the following equation that is independent of the central time:
250 Geometry, Language and Strategy—Vol. 2

    x a     q  4i eq a   x a


  2 e2 q 1  e e  4a a   2i eq a    q  x a   0 . (6.80)

(34) Show that the equation in Ex. 32 also has linear solutions:

x a  U 0 a  V0 a ,
   U 0 a   q     U 0a
2 e q a  q    V0 a  4a e q V0 a 0,
   V0 a   q     V0 a  0 . (6.81)

(35) The solutions in Ex. 33 of the wave equation are in terms of


complex numbers. Show that the solutions can also be written in
terms of sines and cosines as the two coupled sets of differential
equations below. Note that the freedom to choose the coefficients
rests entirely in specifying the boundary conditions. Once the
boundary conditions are set, the remaining behaviors are coupled as
shown.

x a  U a cos  V a sin ,
   U a   2 e 2 q 1  e e  4a a U a
  q     U a
2 e q a  q    V a  4 e q a V a 0,
   V a   2 e 2 q 1  e e  4a a V a
  q     V a
2 e q a  q    U a  4 e q a U a  0. (6.82)

(36) An estimate of the behavior of the solutions to the phasor equations


is that the coefficients are constants in Ex. 33 and the phasor

solution is a wave x a  e k y . This match is well-defined for the
initial values and is a reasonable approximation in the nearby
region. Show that the propagation vector k satisfies the following
Steady State Harmonics 251

algebraic equation. Discuss solutions to this equation if the


propagation vector were entirely along a single axis.

k k      q  4i e q a  k


 2 e 2 q 1  e e  4a a   2i e q a    q   0 . (6.83)

(37) Investigate the behavior of the propagation vector in Ex. 36 by


assuming that the initial conditions are that q  a  0 and that the
propagation vector is along a single direction z with magnitude
   k k  . Show that the equation and solution are below. Infer
that for sufficiently small frequencies there are two solutions to the
wave equation, which have no oscillations and are attenuated along
the  z direction or along the  z direction. Show that for
sufficiently high frequencies there will be a traveling wave, but one
that is still attenuated in general. Write the propagation magnitude
as  p   p  i  p and show that the velocity of the wave is   p .
What is the velocity of the wave for sufficiently large frequencies?
The source of the attenuation in both cases is the presence of a non-
zero player bonding  z  . Discuss how these insights carry over to
solutions that take into account the additional spatial dependences.

 p 2    z  q z   p   2 1  e e   0
. (6.84)
2   z  q z   i  1  e e   4   z  q z 
 2
  1 2 1

(38) In the central holonomic frame the player payoff fields are
determined by a current, Eq. (4.14). Show that the current
components are given in terms of the following expressions
involving the co-moving orthonormal scalars:

p j  E j p 
. (6.85)
p j  eq E j  p e  2eq E j a p 

(39) Show that the gauge condition implicitly chosen in the central
holonomic frame is
252 Geometry, Language and Strategy—Vol. 2

b g ab  0 . (6.86)

(40) Assume that there is a Killing vector K a  V a proportional to the


flow V a , for time and the active strategy directions. Assume that the
inactive strategy components are zero, K j  0 . Assume that this
Killing vector commutes with all the inactive central strategies,
those that mutually commute with each other. Show that these
mutually commuting set of Killing vectors, along with the
remaining active coordinates form coordinate potentials that can be
defined for all coordinates and form the central holonomic frame.
This frame has the advantage that there are no player biases due to
the charges. Transform to a new frame, the active-co-moving frame,
k    in which only the active spatial flow components are
zero. Using the transformation below show that one can transform
to the fully co-moving frame in which all flow components are zero.
Show that in that frame that time and the active strategy directions
are still holonomic, though the inactive directions may not be.

Also show that in the central holonomic frame, you can implement
the player independence hypothesis from Vol. 1 by explicitly
defining the player ownership structures to be the unit diagonals
Q  , which are unity when the strategy (active or inactive) is owned
by that player and zero otherwise. This gives an alternative to the
strong ownership concept in Sec. 11.5.2. Show that different
boundary conditions for the flows are addressed by the
transformations below and the harmonic wave equation.

 k   
 
k  k
V k
0
W   k
,
 0 1 0
 
 0 0  
Steady State Harmonics 253

 k   
 
k  kk V k 0
W  1 


 0 1 0
. (6.87)
 
 0 0  

(41) Assume we start with the potential A  e q a and its equation of


motion Eq. (6.65). Assume further that we have the gauge
conditions    A  0 and    0 . Show using the equations
of motion from the full set of field equations that we have
    0 , and so if the gauge conditions are satisfied on a
surface they are satisfied everywhere. That means we only have to
solve the equations for the potential:


   A  e q  e p     e       q  
2  e        2  q   2   e  . (6.88)

(42) Assume that we have the assumptions stated in Ex. 41 and that z is
a coordinate direction that is constant on the initial surface. Show
that the gauge condition  z   0 and the equations of motion leads
to the following equations, which form an elliptic partial differential
equation on the initial surface.



  z A  0   z  z Az     z A  0 ,
 z

   A      A
z z    z A  ,
  z

 e  q     Az    z A 
z

 e p z  z   e  z      q  

2 z e   z      2z q   2 z   e  0 . (6.89)

(43) The technique of using an integrating factor can also be applied to


the equations of motion for Eq. (5.33), so that for any set of vector
254 Geometry, Language and Strategy—Vol. 2

potentials b and for the matrix function λ B determined below,


this equation is always satisfied.

 
            dx  dx   dx   0
   B    b    b 
 
 B

   B     b    b   dx  dx   dx   0 . (6.90)
 B     B   0  λ B  exp  φ 
      q   2                   p

(44) Continuing with the vector potentials from Ex. 43, show that the
equations of motion Eq. (6.16) can be determined with gauge
conditions     b  0 and    0 defined on the surface:

     p  2       q 


          
  b  B 1   p  2    
  q         b      b 
2      b       b   . (6.91)

(45) Show that for Ex. 44 and an initial surface defined by a coordinate
axis z being constant, the elliptic partial differential equations on
the surface for the gauge condition are:

     b z    z b 
 z

B 1   p z  2  z  


  q      h  2     b z

     b  z   0 . (6.92)

(46) For the single active strategy streamline solutions of Sec. 6.8, show
that the inertial forces, Table 5.6, provides the diagonal components
of stress, Eq. (5.38):
Steady State Harmonics 255

  p  
  q    p   ph     h 
 n 1 
   1 2      1 2      0. (6.93)

(47) For the single active strategy streamline solutions of Sec. 6.8, show
that the inertial forces, Table 5.6, provide the divergence of the
acceleration field given by Eq. (5.61).
(48) For the single active strategy streamline solutions of Sec. 6.8, show
that Eq. (6.48) can be used to determine the divergence of the player
bonding:

   3   1 2       1 2       . (6.94)

(49) For the single active strategy streamline solutions of Sec. 6.8, using
the divergence of the acceleration Eq. (6.50), show that the
harmonic wave Eq. (6.44) can be expressed as:

1  e e 

 q q 2   2 x a    x a  2 q   x a
    p  a
  q q  2       p    x
  n  1  
  q      x a  0. (6.95)

(50) For the single active strategy streamline solutions of Sec. 6.8, using
the divergence of the acceleration Eq. (6.50), show that the
harmonic polynomials Eq. (6.45) can be expressed as:

X N 1a  X N  2 a  0,
  X N    2 N  1 q  
 a  
 
 XN a

    p  a
 N  Nq q  2       p   XN  0,
  n  1  
  X N 1a    2 N  1 q      X N 1a
    p 
  N  1   N  1 q q  2       p    X N 1
a
 0,
  n 1 
256 Geometry, Language and Strategy—Vol. 2

s  N  2, N  3,,0 :
    p  a
  X s a  s  sq q  2       p   Xs
  n  1  
   2 s  1 q      X s a   s  2  s  1 1  e e  X s  2 a  0 . (6.96)

(51) For the single strategy model, show that the wave equation (Ex. 32)
simplifies to:

  x a   2 e 2 q 1  e e  x a     q   x a  0 . (6.97)

(52) An alternate approach to the player fixed frame model is to solve


decision process theory directly in the centrally co-moving frame
based on the centrally co-moving hypothesis. To proceed we need to
set the gauge in a way that is practical for numerical computations.
Ideally, this means setting the gauge on an initial surface and
verifying that the field equations in this gauge maintain this
condition without explicitly imposing the gauge conditions away
from that surface. We intend to use the method of lines in
Mathematica and so it is convenient to identify one active strategy
z in the central co-moving frame along which we “evolve” the
equations. In other words we start on a surface z  0 , write
appropriate boundary conditions on the other strategic directions  .
Although the central time is holonomic in this frame, none of the
scalars depend on the central time based on our hypothesis. Show
that this approach is possible with the seasonal gauge defined
below. We assume the gauge condition is imposed on the surface at
z  0 and assume further that the derivatives along the normal to
this surface vanish. Verify that this is possible by showing that the
field equations that result from Rzz involve only first order
derivatives of the metric elements in question and hence the values
established on the initial surface will remain for all values of z .
This gauge and this property are similar mathematically to the
radiation gauge [Wald, 1984, pp. 80-81], though we are imposing
the condition along a spacelike direction not a timelike direction.
Steady State Harmonics 257

g z  g z  0
. (6.98)
 z g z   z g z   z g zz  0

(53) For the model and steady state gauge in Ex. 52, show that for the
metric, a reasonable set of starting values on the initial surface
would be:

g    
. (6.99)
 z g   0

(54) Show that the player fixed frame model solutions in the centrally
co-moving frame not only satisfy the seasonal gauge in Ex. 52 but
also the gauge Ex. 39. Show that these two conditions however are
not equivalent and that in general, there is no reason to believe that
the gauge for the player fixed frame model will continue to hold for
more general models. Start with the seasonal gauge and on the
initial surface z  0 , impose the additional constraints as written
below. Discuss whether these constraints will be maintained by the
field equations. We have demonstrated in Ex. 52 that the last
equation is maintained, but did not draw any conclusions about the
other equations.

  g    z g z  0    g   0
  g    z g z  0    g   0 . (6.100)
 z
  g  z g  0  z g  0
zz zz

(55) With the assumptions of Ex. 52 for the seasonal gauge, show in
particular that the curvature components Rzz Rza Rab   , Eq.
(3.36), that enter into the field Eqs. (3.31) for the transverse active
strategies ,,  z, are given below. The notation is that the
metric splits into a block diagonal form with a z -block g zz and a
non- z -block g ab  : a , b   , . Do these equations imply that the
seasonal gauge conditions are maintained?
258 Geometry, Language and Strategy—Vol. 2

g az   z g az   z g zz  0,
g zz  g zz   , g ab   g ab   , z  ,
Rzz  1
2   g   g zz   1 4 g zz g   g zz  g zz
 1 4 g   g zz  ln g
 
 1 2  z  z ln g  1 4 g de g cf  z g ce  z g df , (6.101)

 

Raz   1 2  gb  z g ab   1 4 g zz gb  g zz  z g ab 


 1 4 gb  z g ab   ln g  1 2  a  z ln g
 
 1 4 g zz  a g zz  z ln g  1 4 g cf g de  a g ce  z g df , (6.102)

  
Rab   1
2  a  b ln g  ab

 1 2 ab  ln g  acd bdc 
  

 1 2 ab

g zz  g zz  1 4 g zz g zz  a g zz  b g zz
 1 2 g zz  a  b g zz  1 2 g zz  z  z g ab 

 1 4 g zz  z g ab   z ln g  1 2 g zz g cd  z g ac   z g bd  . (6.103)

(56) Given the results from Ex. 55, show that an appropriate gauge for
the vector potential associated with the time isometry is the
following, which can be satisfied by starting with a z   z az  0 on
the hypersurface z  0 . Show that the gauge will be maintained by
the field equations at all other z .

g  a g
g z  a z g . (6.104)
az  0
Chapter 7

Prisoner’s Dilemma: A Code of


Conduct Application

In Part 1, we provided an in-depth development of decision process


theory. In Part 2, we shall test the properties of this theory by analyzing
various scenarios. There are four major types of scenarios: steady state
scenarios, reflection and transmission scenarios, resonant scenarios
and transient scenarios. In Vol. 1 we focused on resonant scenarios. In
Part 2, we focus on the other three, starting with steady state scenarios.
We start with the steady state aspects, which allow us to delve deeper
into the connection of this theory to earlier game theory descriptions.
Game theory takes a static view that the system has reached equilibrium.
Moreover, game theory assumes that the payoffs that govern equilibrium
don’t change in time or vary as a function of strategic choices. Hence
game theory views the structure and geometry as flat as well as static.
There may be other possible states, in the sense of statistical mechanics
to which the system might “jump”, but the description of that jump is
outside of the theory. We take a quite different view that the behavior is
not statistical but a deterministic steady state process. The jumps are to
be dealt with as a continuous process. The theory is deterministic in the
same sense that fluid flow may be considered steady state yet reflects a
physical flow subject to Newtonian mechanics.
A very close analog is the behavior of circuits and transmission lines
in electrical engineering. There, one starts with a study of DC circuits,
which are static. One then moves to AC circuits that with harmonic
source terms are steady state flows. The AC circuit description also
works for transmission lines, though in this case there is a travelling wave

261
262 Geometry, Language and Strategy—Vol. 2

that can move along the transmission line. Thus there are two parts to the
story.
First there is the co-moving behavior that provides the relationship
between the frequency of the driving amplitude and the wavelength of
the standing wave. Second there is the circuit like relationships between
the amplitude and phase of the travellng wave at the source and the
amplitude and phase of the wave at the load. The difference in decision
process theory is that the co-moving behavior is specified by a detailed
steady state surface geometry, some aspects of which are suggested by
game theory. In addition the circuit relationships are replaced by
complex relationships between conditions on this (hyper)-surface and
equations of motion that determine the behaviors away from that (hyper)-
surface. These decision engineering relationships we explore in Part 2.
The steady state surface geometry is characterized by four major
categories of steady state scenarios: codes of conduct, competition,
cooperation and inertial flow. We illustrate these in a variety of ways,
starting with the prisoner’s dilemma with a single active strategy and
three additional codes of conduct. In Chap. 7, the codes of conduct
provide a new and interesting perspective not often covered in game
theory. This perspective helps to resolve what is sometimes considered a
dilemma when game theory is applied to this problem. In Chap. 8, the
steady state geometry is examined in more detail. In Chap. 9, dynamic
scenarios including determinism and chaos provide further insight into
the dynamic mechanisms.
In Chap. 10, we consider an even simpler game with a single active
strategy and no codes of conduct: Robinson Crusoe economics. We
discuss a taxonomy or organization for the theory. In Chap. 11 we set the
context of these examples with a general review of game theory and
recent economic approaches. A key difference with our approach and
game theory approaches is the way we deal with risk, which is covered in
Chap. 12. Our approach is similar to the systems dynamics approach
[Forrester, 1961], which we cover in Chap. 13. In both our approach and
the systems dynamics approach, it is important to take a global
perspective where we consider and close all of the feedback loops,
especially those that are often missed: those occurring over long time
intervals or distant strategic distances.
Prisoner’s Dilemma: A Code of Conduct Application 263

7.1 Introduction

There are a rich range of possible steady state geometry descriptions of


the prisoner’s dilemma. There may be no codes of conduct and all of the
strategies available from each player are active. In the co-moving central
frame, the competitive nature of the steady state geometry plays a
prominent role. At the other extreme, we can have dynamic behavior
based on a single active strategy in the co-moving central frame. In this
case, there are no payoffs or competitive behaviors and the remaining
prisoner strategies are codes of conduct. The cooperative and inertial
aspects of the steady state geometry play the prominent role. We
consider the general competitive analysis later in Part 3. However here,
we focus on the single strategy possibility.
Recall the discussion from Sec. 1.7 and the connection to game
theory. Decision process theory developed in the intervening chapters
substantially diverges from those game theoretic origins. It is our goal
here to illustrate this by quantitatively analyzing the choices prisoners
face using a single strategy player fixed frame model. The model
highlights how far we have come theoretically from our starting point
and demonstrates how the approach from Chap. 6 can be applied to give
new insights.
For example, we argue that the player fixed frame model applied to
the prisoner’s dilemma demonstrates the stability of solutions that rest on
an adopted code of conduct (Sec. 11.2) of the players, which we believe
is an important departure from game theory. In this sense the prisoner’s
dilemma is the “hydrogen atom” for decision process theory since it
provides a solvable model illustrating the new features of the theory, just
as the “hydrogen atom” provides an important and solvable model for
classical and quantum mechanics in physics. The prisoner’s dilemma is
an example of a contract (Ex. 5) or code of conduct, which forms the
basis of the invisible hand [Smith, 1776]. The difference from game
theory and current economics is that the invisible hand is not a
consequence of competition but of a code of conduct, which has different
dynamics.
The prisoner’s dilemma has been extensively investigated by game
theorists since the late 1950s. The common sense solution to the
264 Geometry, Language and Strategy—Vol. 2

prisoner’s dilemma is to argue that the situation is an example of the


tragedy of the commons (Sec. 11.2). The tragedy occurs when both
prisoners take from the commons, causing both to suffer. However, this
tragedy of the commons is avoided if the prisoners adhere to a common
code of conduct, which in this case would be to remain silent when
questioned. We demonstrate that codes of conduct are natural attributes
of decision process theory. They come into being whenever we identify
strategies as inactive. Codes of conduct are conserved and therefore are
persistent attributes of the theory. They persist therefore only if they are
adopted by every player.
There are other aspects of the theory that are stable. The solutions that
result from the field equations can be proved to have a mathematical
stability in the sense that small changes to the initial conditions lead to
small changes in the behaviors of the solutions. This result is a
consequence of the differential geometry for Einstein type hyperbolic
partial differential equations [Hawking & Ellis, 1973, p. 254]. In
particular, there is a frame in which the steady state hypothesis (Sec. 6.1)
results in scalars, vectors and tensors that are independent of time. Given
these values and given initial conditions, behaviors in other frames of
reference have such a mathematical stability.
We observe that structural stability is also important and distinct
from mathematical stability. Whether a bridge stands or falls, the laws of
physics provide a mathematical stable description; but for it to stand we
require the bridge to be structurally stable. Since the player fixed frame
model with the steady state hypothesis has similarities to a bridge in that
there are significant steady state aspects to both, we will also have
occasion in our discussion to speculate on whether or not our solutions
are also structurally stable. As with bridges we look for the tell-tale
creaks and groans from the stresses and strains to imagine what might
happen if our mathematical assumptions about the stresses were to break
down.
We treat the problem of the prisoner’s dilemma by first elaborating
on the steady state scenario aspects: we start by discussing the relevance
of altruism and egotism (Sec. 7.2), followed by analyzing the prisoner’s
dilemma into normal form (Sec. 7.3). We then provide an equivalent
formulation of the initial conditions (Sec. 7.4) and introduce the code of
Prisoner’s Dilemma: A Code of Conduct Application 265

conduct behavior (Sec. 7.5). We turn to solutions based on known


behaviors (Secs. 7.6-7.10), which describe the initial conditions. A key
attribute is that the theory has cooperative behaviors, which manifest as
stresses (Sec. 8.1) and strains (Sec. 8.4), which are addressed in the next
chapter, along with a discussion of the related characteristic of the
theory, its persistent behaviors (Sec. 8.8). We then discuss dynamic
scenarios in the following chapter, including a sensitivity analysis
(Sec. 9.5) and a summary of normal-form behaviors (Sec. 9.8) and chaos
(Sec. 9.9). Dynamic scenarios include reflection and transmission
scenarios, transient scenarios and resonant scenarios.

7.2 Egoists and Altruists

The results in this section extend previous work by Thomas & Kane,
[2008] and Thomas & Kane [2010], who applied a single strategy player
fixed frame model to the prisoner’s dilemma. They modified the forms
for the payoffs introduced below in Eqs. (7.3) and (7.4), to this model.
We shall see that reducing the number of strategies in this way to a single
active strategy is equivalent to choosing a code of conduct (Sec. 11.2). In
this case, public-interest supplements self-interest.
These authors found similarities to the decision process theory with
the incomplete games proposed by Harsanyi [1967-1968], in his trilogy
on game theory. Decision process theory goes one step further than
Harsanyi in proposing that actions are characterized not only by
competition as measured by payoffs, but by inertia and charge. As with
physical systems, inertia implies that there must be enough force to
overcome inertia in changing a system from a given course. Ordinary
competition-forces—i.e., the forces that arise from one’s beliefs and past
experiences—may not be sufficient. The second is the property of
charge. In their model of public goods games, Eshel, Samuelson,
& Shaked [1998] described players (subjects) as either egoists (who
maximize self-interest) or altruists (who maximize other-interest).
In contrast, Thomas & Kane [2010] suggested that Eshel et al.’s
egoist/altruist distinction corresponds roughly to a well-known and
empirically-tested distinction made in psychology regarding independent
266 Geometry, Language and Strategy—Vol. 2

and interdependent worldviews. They refer the reader to Markus &


Kitayama [1991] for an extensive review. The suggestion is that subjects
who possess independent worldviews correspond roughly to Eshel et
al.’s egoists, while subjects with interdependent worldviews correspond
roughly to altruists. Based upon this insight, the suggestion was to
incorporate the independent/interdependent distinction and liken it to the
physical property of charge.
We feel that a more satisfactory conclusion (Sec. 11.2) is that the
worldview of independence is associated with the players’ inactive
strategy whereas interdependence is associated with a code of conduct
that treats as inactive what would otherwise be active strategies. We
associate charge (Sec. 12.4) with the player’s interest flow, which we
characterize as either that of a giver or taker. This has some of the same
sense of the above-mentioned authors, though we take the words to be
more akin to seller and buyer or producer and consumer. Alternatively
we could use the terminology accommodating and greedy for positive
(giver) and negative charges (taker), respectively. We find for the
prisoner’s dilemma that the more aggressive player is greedy (Cf. Sec.
8.8.4).
Unlike game theoretic approaches, such as [Harsanyi, 1967-1968] we
do not consider separate subjective and objective payoff matrices. In
decision process theory, each player’s payoff matrix is a personal payoff
matrix. That is, each player constructs a payoff matrix that represents
what that player currently believes will result as payoffs to him or her
and all other players. These personal payoff matrices are not public
knowledge within the game: each player constructs and can trust only his
or her personal payoff matrix. However, players may learn or infer about
the action of others as the game is played and over time, may come to a
more accurate view based on their own outcomes and the outcomes they
observe for others. It is reasonable and should be expected that this
payoff matrix will change over time.
There are two additional properties of personal payoff matrices that
are important. First, although each personal payoff matrix is private
information rather than public information, according to our theory, each
personal payoff matrix is observable and measurable, though it may not
be easy to measure. An analogy may be helpful. The field of empirical
Prisoner’s Dilemma: A Code of Conduct Application 267

psychology attempts to understand people’s (and occasionally animals’)


observable behaviors in terms of mental processes that cause them.
These mental processes are very difficult to observe, however. The
underlying properties of the mental processes are usually not possible to
observe under circumstances that present themselves in the everyday
world. In order to understand mental processes, a person’s environment
needs to be controlled in ways that the everyday environment usually
does not allow [Stanovich, 2004, p. 92]:
“The occurrence of any event in the world is often correlated with many other
factors. In order to separate, to pry apart, the causal influence of many
simultaneously occurring events, we must create situations that will never occur in
the ordinary world”.
When the environment is controlled in this way, mental processes may
become observable. This observation about psychology provides insight
to the personal payoff matrices. The personal payoff matrices are
potentially observable and measurable, but it would be difficult to
measure them, especially in the context of a game already in play. This is
why players’ personal payoff matrices are usually not public information
within the game. Nonetheless, these personal payoff matrices are still
considered objective and measurable.
The second important property of personal payoff matrices is that
each player’s personal payoff matrix is idiosyncratic: that is, the players’
personal payoff matrices do not have to be identical to each other and
most likely depend on what strategies have just been played. If each
player generates his or her personal payoff matrix based on his or her
past experiences (which are by definition particular to each player), then
each player will most likely generate a different personal payoff matrix
that depends on strategy and time. Decision process theory proposes
rules for how such personal payoffs change based on past behaviors.
In decision process theory, to specify a solution, we provide the
initial strategies and initial payoffs for each player; the field equations
then provide unique solutions. To compare and contrast our results with
game theory, we choose as our initial strategies those that game theory
might propose as the equilibrium strategies. We choose as our initial
payoff matrices those that game theory might propose. If these were to
stay the same at all other strategies and all later times, we would recover
268 Geometry, Language and Strategy—Vol. 2

the game theory Nash equilibrium result. To the extent they change we
obtain the deviations of decision process theory from game theory.

7.3 Prisoner’s Dilemma—the Story in Normal Form

In this section, we put the prisoner’s dilemma into normal form. We


identify the codes of conduct in Sec. 7.5. From Sec. 1.7 we recall that the
essential details of the story are that two prisoners are being held for a
crime where it is suspected they have acted in concert. Each is given a
choice to confess or not confess with penalties that are supposed to
induce confession of their guilt. However if neither confesses they will
get off lightly. If both confess they will be penalized but not as severely
as the case in which one confesses and implicates the other.
We are interested in understanding the decision process that takes
place for each of the prisoners. We note that the game theory analysis
argues that the players will act only based on their self-interest, which
leads each of them to conclude that they should confess. The paradox or
dilemma is that common sense suggests that there are reasons why both
prisoners might choose not to confess, yet this choice is not supported in
game theory. We need to show that both scenarios are possible in
decision process theory as is also true in real life. The relevant task is to
determine the initial conditions and what happens next.
We start with the formulation of the prisoner’s dilemma as a game in
normal form between two prisoners specified by the following payoffs
to player 1.
1
G12 N2 C2
N1 0.1 1 . (7.1)
C1 0 0.9

There are identical payoffs to player 2:


2
G21 N1 C1
N2 0.1 1 . (7.2)
C2 0 0.9
Prisoner’s Dilemma: A Code of Conduct Application 269

We suppose that at some initial value of time and at some special


strategic point in the strategic space, there are quantitative payoffs for the
matrix G k of possibilities as illustrated above for each player k  1,2 .
We differ from game theory in that these payoffs are not constants for all
points of space and time.
At any point in space-time, we describe the elements of the matrix in
detail for player 1, noting a similar description holds for player 2. For
player 1, the rows are labeled (C1) if player 1 confesses and (N1) if player
1 does not confess. The columns are labeled in a similar manner (N2) and
(C2). Our initial choices for the payoffs translate as follows. At the initial
point of space and time, the payoffs reflect quantitatively the posed
problem: if both confess, player 1 loses 9 10 units; if both don’t confess
player 1 loses a much smaller value 110 . If player 1 confesses and
implicates player 2 who does not confess, then player 1 loses 0 units. On
the contrary, if player 1 does not confess and is implicated by player 2,
player 1 loses 1 unit. At other values of time and space, the payoff values
are computed from the player fixed frame model, though we have yet to
identify which strategies are inactive.
This formulation has several general properties in common with game
theory. We have payoff matrices and we have mixed strategies. We
differ however in our method of computing the mix of strategies from
which each player picks. In game theory, the mix of strategies is
associated with equilibrium.
For example, the payoffs for the two players need not add up to zero;
when they do [Von Neumann & Morgenstern, 1944)], such zero-sum
games have an equilibrium value that is computed as follows. Suppose
the payoffs Eq. (7.1) for player 1 represented a zero-sum game. The most
conservative strategy for player 1 would be to determine the minimum
outcome for each of its pure strategy choices. If player 1 chooses (N1),
the minimum case is 1 . If player 1 chooses (C1) the minimum is  9 10 .
The maximum of these two is  9 10 . The most conservative strategy is to
choose this max-min. For pure strategy choices, there may not always be
a max-min solution. What Von Neumann & Morgenstern [1944] showed
however, is that for a zero-sum game with any number of players, each
with any number of strategies, there is always a max-min using mixed
strategies. This equilibrium strategy determines the mix of strategies in
270 Geometry, Language and Strategy—Vol. 2

game theory. For two players for a non-zero sum game, an analogous
Nash equilibrium determines the mix of strategies. For the prisoner’s
dilemma it is that both players choose not to confess. They optimize their
self-interest by acting defensively.
To articulate the difference, we don’t need to work directly with the
individual payoffs such as Eq. (7.1) but can construct our argument using
the symmetric game Eq. (1.12):

Fab1 N2 C2 N1 C1 t
N2 0 0 1
10 0 0
C2 0 0 1 9
10  9
10 0
. (7.3)
N1  110 1 0 0 1
0

C1 0  9 10 0 0 9
10 0

t 0 9
10 0
 1
0
 9
10 0
0

We identify the “hedge” strategy with time t and take this payoff to be
the value at an initial value of time and at an initial strategic point.
We construct the similar payoff for player 2, again evaluated at the
same point:

Fab2 N2 C2 N1 C1 t
N2 0 0  110 1 1
0

C2 0 0 0  9 10 9
10 0
. (7.4)
N1 1
10 0 0 0 0
C1 1 9
10 0 0  9
10 0

t  1
0
 9
10 0
0 9
10 0
0

These forms for player 1 and player 2 define new payoff matrices that
are each symmetric. In each case, a two-person game is defined in which
each player has the same set of strategies. The max-min of each
symmetric game is the same as the original game from which it was
constructed and so it represents exactly the same strategic possibilities as
the original game.
Prisoner’s Dilemma: A Code of Conduct Application 271

In game theory, it is a property of the Nash equilibrium that the


product of the payoff for player 1, Eq. (7.3) and the mixed strategy
 Nash  0,1,0,1,  0  is zero: the flow is in the null space of the payoff
matrix. Similar statement holds for player 2. Instead, in decision process
theory we argue that the mix of strategies is determined dynamically. We
can pick any mixed strategy for our initial point.
In decision process theory, whether or not the rate of change of the
flow dV a d changes in time depends on the flow equation that
replaces Eq. (1.13). In the normal-form coordinate basis, the
replacement flow equation is determined by Eqs. (2.49) and (4.42):

DVa
qa  q Ea 

. (7.5)
dV b
gab   abcV bV c  Vk Fabk V b  1 2V jVk  a jk
 qa
d
We see that the flow may be in the null space of each of the payoff
matrices and yet the rate of change of the flow is not zero.
Conversely, even if the flow is not in the null space of the payoffs,
the other terms, including the covariant acceleration term qa , can
conspire to make the flow steady state, dV a d  0 . We thus
acknowledge the importance of self-interest, to the extent we apply that
loose concept to the forces derived directly from the payoffs. However
we must also acknowledge the potential importance of other effects. In
order to estimate the size of such other effects we must appeal to model
calculations of the various conspiracy effects in Eq. (7.5).
An important example of such effects consists of resonant
scenarios, which arise when only the acceleration and payoff terms
contribute:

dV b
g ab  Vk Fabk V b  0 . (7.6)
d

This special case arises when the metric and payoffs are constant along
the streamline, the cooperative and gradient terms are negligible and the
orientation components are zero. Such harmonic behaviors were
described in Vol. 1. It is worth emphasizing that we expect these
272 Geometry, Language and Strategy—Vol. 2

behaviors in the active space and characterize the driving force in terms
of the active payoffs. Typically these harmonic behaviors oscillate
around the Nash equilibrium. We analyze the competitive behavior
further in Sec. 15.1. In physics, the analogous situation is a charge
particle in a magnetic field oscillating around the direction of the field.
Thus in moving away from game theory to decision process theory
with the steady state hypothesis, we expect a much richer geometry or
landscape. In particular, in the centrally co-moving frame, all of the
terms in Eq. (7.5) will be independent of the central time  . It will be in
this frame and the closely related co-moving orthonormal frame that we
investigate the steady state scenarios. Consider the acceleration values
q in the co-moving orthonormal frame. An important attribute of non-
trivial steady state behavior is that there will be acceleration. For the
player fixed frame model, this is found for the active components q .
We see that in the game theory approximation, with the initial
condition that the acceleration is zero, the acceleration stays zero at all
other values of time and space. We break this approximation in the
player fixed frame model, even when the orientation potentials are static.
This is because the variation of the acceleration with distance is
determined by the divergence Eq. (5.61):

  p
 q      p    2        q  q q . (7.7)
    

 n 1 

Whatever the origin of the conspiracy effects, they are not caused
directly by the payoffs. The results are governed by the inertial, player
interest and player bonding effects.
This suggests that we may learn a significant amount by considering
even the simplest single strategy model in which the payoffs (for the
effective players of the model) are greatly simplified. We then focus our
study of the effects of Eq. (7.7), further simplifying it since with a single
active strategy there are no tidal magnetic effects   0 .
Prisoner’s Dilemma: A Code of Conduct Application 273

7.4 Equivalent Formulations of the Initial Conditions

In game theory, there are no strategic distinctions between games whose


payoff values are scaled by a common factor, or whose values are
increased or decreased by a common amount. They lead to the same
Nash equilibrium value, which expresses the strategic content. Moreover,
the strategic content depends on scaled strategies, which are by
convention summed to unity. Again, the scale does not represent any
strategic content in game theory.
We now look into various formulations of the prisoner’s dilemma that
are equivalent in this sense to the game theory normal form, Sec. 7.3. In
decision process theory, it is useful to identify these equivalent models,
as they may lead to different dynamic consequences. For example, the
payoffs for each prisoner could be scaled by arbitrary factors  k for each
player representing the player stakes in the decision process. We could
add a constant value to every element of the payoff. Or we could rescale
the strategy values. In decision process theory, the dynamic behavior is
influenced by these possibilities, as we show in our numerical sensitivity
analysis.
We start by rescaling the strategy values. An important extension to
game theory is our introduction of the time component  0 of the flow,
which leads to the invariant definition of decision mass:

m        . (7.8)

Nothing in game theory depends on this parameter. If we scale the active


and inactive strategies by this mass, we obtain the normalized flow
components:
a
Va 
m
. (7.9)
 j
V 
j

This changes our thinking of the strategies as being summed to unity for
each player, which focused only on the relative strategic values and
274 Geometry, Language and Strategy—Vol. 2

assumed that nothing of strategic value depended on the hedge strategy


.
Now however, we ascribe meaning to a different set of ratios. The
features that remain when we scale the flow to unit length, Eq. (7.9), are
characterized by the beta  of the initial values:

1

 0       
a
a 2

j
j 2
. (7.10)

The value of beta depends on the active and inactive flow components.
For the prisoner’s dilemma, the active strategy space has four
dimensions. Since there are two players, the inactive strategy space has
two dimensions. We choose the initial inactive flows to be zero:
 j  0 0 .
The total strategic space dimensionality is n  6 . We next set the
active flow values. For numerical work, if we provide a value for  ,
we obtain  0 from Eq. (7.10), defined by the initial strategy flow. The
decision mass is then obtained from Eq. (7.8). The Nash equilibrium
for the prisoner’s dilemmas is that both players confess:
 Nash b  {0,1,0,1, m} , where m can be chosen arbitrarily at this point.
This determines beta.
Nothing in game theory depends on the game value, since it does not
influence the choice of Nash equilibria. Typically, one can add to each
game a constant amount leaving these strategic possibilities invariant. In
the symmetric form this constant is the matrix:

0 0 v v v
0
 
0 0 v v v
0 
 
c v   v v 0 0  v 0  . (7.11)
v v 0 0  v 0 
 
 v  v v v 0 
 0 0 0 0 

We are free to choose any game value. We note, however, that no


constant game value amount added to the player 1 payoff in Eq. (7.3)
Prisoner’s Dilemma: A Code of Conduct Application 275

will eliminate all of the payoff time components (elements in the row or
column labeled by time t ). We change this for our example.
Our argument is that in decision process theory, the time components
of the payoff reflect the biases that we believe accurately represents the
notion of the game value from game theory. Not every game needs to be
fair, nor do we require the time components for each player to be equal.
We see that is true for the constant Eq. (7.11) but not for the assumed
forms we have chosen for the prisoner’s dilemma. This suggests another
equivalent form that does not change the game theory strategic content,
but may represent different dynamics in decision process theory.
It should also be possible to frame the prisoner’s dilemma in a way
that it too could be put into the form of a fair game where the time
components for each player are equal. We achieve this by adding
internal payoffs to F 1ab so that the product F 1ab Nash b  0 of the payoff
and the starting strategy  Nash b are zero:

0 9
10  45  9 10 0  0   0 
 9    
  10 0 0 0 1   0 
1
10

 45  10 0
1 1
10 0  0    0  . (7.12)
9    
 10 0  110 0 0 1   0 
0 0  0  
 0 0 0     0 

This payoff plus the constant matrix c   9 10  from Eq. (7.11) reproduces
the same payoffs as the submatrix Eq. (7.1). There is an internal payoff
or utility for player 1 to choose to not confess of 110 . For example,
without this bias, the return for player 1 for the equilibrium strategy
would be negative instead of zero. We thus obtain the same strategic
behavior as the symmetrized game Eq. (7.3).
If both players choose to cooperate and not confess ( N1 N 2 ), then
player 1 loses 110 . If player 1 chooses to cooperate and not confess and
player 2 chooses to confess ( N1C2 ), then player 1 loses 1 unit. If player 1
chooses to confess and player 2 chooses to cooperate and not confess
( C1N2 ), player 1 sees maximum benefit and receives 0. Finally if player
1 and player 2 both choose to confess ( C1C2 ), then player 1 loses 9 10 .
276 Geometry, Language and Strategy—Vol. 2

The Nash equilibrium point is the strategy whose product with the payoff
is zero.
In a similar fashion, we choose the payoff for player 2, so that the
product F 2 ab Nash b  0 is also zero:

0 1
10
4
5  110 0  0   0 
 1    
  10 0 0 0  1   0 
9
10

  45  9 10 0 9
10 0  0    0  . (7.13)
1    
 10 0  9 10 0 0  1   0 
0 0  0  
 0 0 0     0 

This payoff plus the constant matrix c   9 10  from Eq. (7.11) reproduces
the same payoffs as the submatrix Eq. (7.2). We thus obtain the same
strategic behavior as the symmetrized game Eq. (7.4).
We have defined some of the initial conditions that are needed to
specify a complete solution to the field equations. Before continuing, it is
worth reiterating the multiplicity of payoffs that are possible, which
seemingly represent the same strategic situation. We look at the
elementary aspects of game theory for example and see that the strategic
content of a game in normal form does not change if the payoffs are all
multiplied by a common factor. Similarly the strategic content doesn’t
change if a constant is added to all of the payoff elements. Finally,
internal payoffs offset by the appropriate time components don’t change
the strategic result. Adjusting the time components also allows scale
changes for the equilibrium strategies of each player.
The max-min rule discussed in Sec. 1.7 yields the same equilibrium
strategies. If the game is fair, there is no difference between these games.
If the game is not fair, the only difference is the size of the payoff, which
will be modified by the same multiplicative factor or additive constant
that modifies the individual payoffs. The important question is whether
this behavior squares with our common sense. It is reasonable to imagine
that the strategic choices will in fact be impacted significantly if a game
with low stakes is made into a high stakes game. In decision process
theory, we in fact expect there to be differences.
Prisoner’s Dilemma: A Code of Conduct Application 277

The reason for this expectation in decision process theory is that the
payoffs fields carry energy. The larger the field values the more energy
these fields carry. We noted the acceleration flow depends on the inertial
and orientation flux fields, Eq. (7.7). The acceleration flow is a direct
measure of changes to the strategic values. The magnetic components in
the co-moving basis play no direct role, whereas the inertial and player
interest terms do. We thus have at least one difference between decision
process theory and game theory models that can be put to the test. As we
develop the player fixed frame model for the prisoner’s dilemma, we will
study the consequences of different game values by changing the initial
conditions with additions of constant matrices Eq. (7.11). In the next
section we introduce codes of conduct and their relationship to inactive
strategies.

7.5 Persistency, Variability and Code of Conduct

The most general approach to the prisoner’s dilemma (Sec. 7.3) is to start
with the four active strategies associated with our normal form payoffs
Eqs. (7.12) and (7.13). To these strategies we add two inactive strategies,
one for each of the two prisoners. We use the normal form payoffs plus
the appropriate constant matrix c  v  , Eq. (7.11) as the initial payoff
field value. We use an initial flow value that represents the strategic
choice we want to study. For numerical solutions, we specialize to the
player fixed frame model that has streamline solutions that can be found
using the equations from Chap. 6 based on the steady state hypothesis.
We expect solutions that are both steady state and non-steady state.
To break out of the prisoner’s paradox, we impose a player code of
conduct in which we go from two players to five players, three of which
collectively choose a single common active strategy. In these solutions,
there are no active payoffs; nevertheless we can maintain the same initial
conditions for the payoffs and strategies. We still have solutions that are
every bit as stable and correct as the Nash equilibrium solutions when all
strategies are active. Moreover, our approach is not without precedent.
The code of conduct is similar in concept to the standard of behavior of
Von Neumann & Morgenstern [1944, p. 31ff]. They noted that such
278 Geometry, Language and Strategy—Vol. 2

restrictions might lead to a variety of unexpected behaviors in the theory


such as a game theory explanation for discrimination, [Von Neumann &
Morgenstern, 1944, p. 288ff]. The approach here is quite different; it
rests on a dynamic process that incorporates isometries.
The isometries are persistent attributes of the dynamics. The method
we use to find solutions when there are isometries is to rotate our frame
of reference by a constant amount (i.e. an amount independent of space
and time). Moreover, we suggested from a preliminary look at the
behavior of the acceleration gradient Eq. (7.7) that some, though not all
essential features might be seen in these solutions even though they are
less general and simpler to study.
Therefore, from our point of view, every possible player strategic
choice need not be active. For each inactive strategy x , the associated
payoff for each player F j xb has a form simpler than Eq. (2.47), and
provides the payoff in terms of the underlying potentials:

F j xb   b Axj . (7.14)

This is because none of the metric or orientation potentials can depend


on this strategy.
From the standpoint of setting initial conditions, we don’t see any
difference between a single strategy being inactive and none of the
strategies being inactive. There is nothing that prevents us from choosing
any one of the strategies as inactive. However, we would expect a
difference in the behavior along the streamlines.
For an inactive strategy, the streamline is conserved, Eq. (2.50)
extended to the case of a non-zero source (also Cf. Table 5.1):

dVk
 q E j  q E j  0 . (7.15)
d

In particular, for the player fixed frame model, the flow is conserved.
This is distinct from the behavior expected for active strategies that will
be influenced by the rotational effects of the payoff matrix seen in
Eq. (7.5). Multiple numerical examples are provided of this behavior in
Vol. 1.
Prisoner’s Dilemma: A Code of Conduct Application 279

The situation is different for two inactive strategies and a given set of
initial conditions. If two strategies  x y are inactive, in addition to the
conservation behavior Eq. (7.15) of these strategies along the streamline,
there can be no payoff between them:

F j xy  0 . (7.16)

Looking at the payoffs for the prisoner’s dilemma, this is not true
simultaneously for our initial payoff values Eqs. (7.12) and (7.13). There
may however be a different basis in which this is true.
We start with a frame in which for each player one direction is
specified by confess (C) and the other not confess (N). An equally
complete choice would be the sum (C+N) and difference (C-N) of the
strategies. We label the coordinates

t 1 2 N 2 C2 N1 C1  ,

in our initial normal-form coordinate basis. We label 1  2  the player


inactive strategies and label t time. We make two successive linear and
orthogonal transformations starting with:

s1  1
 N1  C1 
2

s2  1 2  N 2  C 2 
. (7.17)
r1  1 2  N1  C1 
r2  1 2  N 2  C2 

The sum of the strategies for each player, r1 and r2 , represent how often
each player plays. They represent the player preference scale. The
differences s1 and s2 represent the player relative preference to not
confess. In this basis, we still can’t choose more than one strategy to be
inactive.
Another equivalent set of strategies is the sum r and difference u ,
leaving all other variables unchanged:
280 Geometry, Language and Strategy—Vol. 2

r 1
2 r1  r2 
. (7.18)
2 2
u 1 r  r1 

In this case the resultant payoffs now have a form that is consistent with
taking three of the strategies inactive.
With the coordinate basis order u r s1 s2 t , we have for
player 1 the initial payoff matrix:

0  45 1
10 2  9 10 2 0
4 
 5 0 0 0 0
F 1
 cab   9 10     110 2 0 0 0

0 . (7.19)
ab 
9 2 0 0 0 0
 10 
0 0 0 0 0 

The transformed payoff field for player 2 has a similar structure:

0 4
5
9
10 2  110 2 0
 4 
 5 0 0 0 0
2
F ab  cab  9 10     9 10 2 0 0 0

0 . (7.20)

1 2 0 0 0 0
 10 
0 0 0 0 0 

To recover the original form of the payoffs for the prisoner’s dilemma
we transform the constant value matrix Eq. (7.11):

0 2v 0 0 2v
0 
 
 2v 0 0 0 0 
c v   0 0 0 0 0 . (7.21)
 
0 0 0 0 0 
  2v 0 0 0 0 
 0 
Prisoner’s Dilemma: A Code of Conduct Application 281

Furthermore, the transformation is not specific to the exact form we


chose for the prisoner’s dilemma. We pick the following payoff based on
arbitrary values x y z w :

0 z w  12  x  y  z  w  12  x  y  z  w
0 
 
 z 0 w  12  x  y  z  w  12  x  y  z  w
 0 
 
 w  2x  y  z w  12  x  y  z  0 y  w 0  . (7.22)
1

 w  1 x  y  z w  12  x  y  z   y 0  w 0 
 2

 w 0  w 0 w w 0 
  0 0 

The transformed payoff is the most general form in which the three
strategies r s1 s2  can be inactive:

0 2 w  x y 2 z 2 2w
0

 
 2w  x 0 0 0 0 
 
y 2 0 0 0 0 . (7.23)
 
z 2 0 0 0 0 
  2w 0 0 0 0 0 
  

We see that the Nash equilibrium is satisfied with


 Nash  0 1  1 2
 1 2  0  for Eq. (7.23) if we impose an
additional condition. We see that only the first row and first column are
non-zero, indicating that we could take some or all of the strategies
r s1 s2  to be inactive. However to satisfy the same Nash
equilibrium we further require both  y z to be positive and z  x  y .
We see the possibility that an equilibrium value at a point does not imply
an equilibrium value at all points.
To understand inactive strategies better, we elaborate on the meaning
of the payoffs that we have found above. They are not active payoffs.
The solutions we consider for the prisoner’s dilemma reflect a symmetric
decision process and could be a just decision process in which the
relative player effort u of the prisoners generates a payoff whatever the
other strategy is from Eq. (7.22). There are no payoffs that depend purely
on the player relative preferences s1 s2  or total player effort r . The
game theory model assumes that this is true for all time and for all
282 Geometry, Language and Strategy—Vol. 2

strategic positions. Thus it is as if these strategies play no role. Even if


we adopt this stance, we assume only that they are inactive: they still
may influence the outcome of the decision indirectly. In this sense we
look at the symmetric prisoner’s dilemma as being equivalent to a
decision process in which there are three additional players controlling
the influence of the inactive strategies r s1 s2  . We note that the
general case would diverge from the usual prisoner’s dilemma in that it
would assume the zero payoffs found above hold only at a single point of
space and time.
Our symmetric prisoner’s dilemma may still represent a utilitarian
solution or a just solution depending on the initial flows along the player
preferences s1 s2  . In either case we achieve the effect using a code of
conduct. A code of conduct can just as easily deny rights based on self-
interest as proffer rights that uphold the public-interest. We could also
frame a model whose only solutions would be just (Ex. 4). Our choice
here was based on wanting a model where we could move between these
opposite choices as part of our sensitivity analysis of the numerical
solutions. Based on these analyses we are able to identify the
mechanisms that are needed to support the just solutions (Cf. Sec. 7.5).
With these caveats, we develop the consequences of the prisoner’s
dilemma considered as a player fixed frame model in which there is a
single active strategy u  1 2  r2  r1  measuring the relative player effort
of the two prisoners. The steady state single strategy solutions we obtain
are consistent with Vol. 1, though we extend these solutions to dynamic
steady state solutions. Moreover, the choice of the three inactive
strategies is equivalent to a code of conduct (Sec. 11.2) in which each
player will fix the rate at which they will confess. The code of conduct is
that each prisoner agrees to treat his strategy difference sk as inactive.
Although each prisoner controls his own difference, we assume that there
are various levels of enforcement that hold these strategies inactive.
Furthermore, we identify an additional code of conduct by taking the
overall player scale, which is the sum of the two active summed
strategies r1  r2 , to be inactive. This strategy is controlled jointly by the
two prisoners. Depending on the specifics, we may also envision all three
inactive strategies as being controlled jointly. The sole active strategy u
that remains reflects the relative player effort of the two prisoners. This
Prisoner’s Dilemma: A Code of Conduct Application 283

introduces an effective agent reflecting joint control, which is


accountable for the relative player effort choice. We suggest the
terminology that when this variable is positive, player 2 is the aggressor
and when negative player 1 is the aggressor.
In the next section, we start with the initial conditions based on the
initial payoff and strategy flows. We demonstrate that we need not
choose the initial strategy to be the Nash equilibrium in order to obtain
stable solutions to the field equations.

7.6 Known Behaviors

Despite the fact that there are no active payoffs, there remain non-trivial
dynamics that are a consequence of the cooperative and inertial forces. A
consequence of the expanded scope of decision process theory is that we
specify more behaviors than just payoffs and active strategies. We have
introduced inactive player strategies as the mechanism that underlies the
notion of a player and payoffs. In addition there are orientation potentials
whose existence and behaviors follow from decision process theory. We
set these additional initial behaviors of the prisoner’s dilemma for the
player fixed frame model with a single active strategy.
We keep in mind the following as we proceed. First, we are looking
for solutions to the prisoner’s dilemma that have three additional
isometries. We are thus justified in calling such solutions symmetric.
These isometries are associated with a code of conduct the players agree
to adhere to. We are also justified in thinking of such solutions as just.
General solutions to the prisoner’s dilemma have much in common with
the utilitarianism school in ethics [Tavani, 2011], whereas symmetric
solutions have more in common with just utilitarianism.
Second, we must be mindful of what we mean by initial conditions.
Solutions to partial differential equations have unique solutions when the
form of the solution is specified on some initial boundary or bounding
surface. It may be at an initial point in time or space. If our intent was to
predict the future based on a known state of the prisoner’s knowledge
about each other and the situation, then the initial conditions would be
based on time. Our goal here is a little different in that we want to
284 Geometry, Language and Strategy—Vol. 2

provide an illustrative set of numerical examples. For this reason we


consider specifying the system in a state with certain simple properties.
For example we “start” the system at a point with zero acceleration,
which facilitates our study of whether it stays at that point or not. We
shall refer to the behaviors at our “starting” point as the known
behaviors.
The third point to keep in mind as we proceed is that the general
description of the decision process Eq. (4.5) depends on the strategies we
identify as active and inactive in the normal-form coordinate basis:

d 2   jk  d j  Aaj dxa  d k  Abk dxb   gab dxa dxb . (7.24)

The invariant distance  we term the proper-time. We formulate each


decision process at the outset assuming that the only inactive strategies
are set by player inactive strategies. However we also note that we put all
strategies on an equal footing when we use the holonomic basis for the
strategies:

d 2  ˆ dx  dx . (7.25)

We can transform from one basis to another, Eq. (2.43) by comparing


terms. We now have a third basis that proves insight, which is the
symmetric normal-form coordinate basis whose form is identical to Eq.
(7.24) but the inactive strategies now include all the isometries, with a
corresponding reduction in the active strategies. We will use the
holonomic form Eq. (7.25) as the common ground to relate values from
one basis to another.
Next, we need to specify the known flows, set the values for the
orientation potentials for the symmetric prisoner’s dilemma, transform
these values from the symmetric normal-form coordinate basis to the
symmetric co-moving orthonormal coordinate basis and deal with the
specification of inertial effects that have no analogy in game theory. We
deal with each of these topics in the following sections.
Prisoner’s Dilemma: A Code of Conduct Application 285

7.7 Known Strategic Flows

In the symmetric normal-form coordinate basis, the initial flow V  is


specified by the active components V a determined by the strategies
chosen and the inactive components V j that specify the coupling of the
payoffs fields to the flow. Along a streamline, these flow components are
the rates of change:

dx 
V  . (7.26)
d
We have normalized the flow to unit length, Eq. (7.9) and Eq. (7.25):

dx  dx
ˆ 1. (7.27)
d d
Based on an initial flow direction, there is a natural set of orthonormal
vectors that can be constructed. The set aligns proper-time along the flow
direction and constructs vectors transverse to the flow using the Gram-
Schmidt orthogonalization process, starting with the initial strategy
directions.
We sketch this process. The first step is to take E a   V a and
E j  V j . We can then choose the remaining vectors in any order, which
we take to be along 1 2 s1 s2 r u . These are independent of the
flow, though not in general transverse to the flow. We take a unit
direction along the first strategy 1 , which defines the direction U1 . We
construct a new orthonormal vector that is both orthogonal to the flow
and of unit length:

U 1  U  1 E E 
E 1   . (7.28)
ˆ U 1  U  1E E   U 1  U 1 E E  

We go through each vector and use the same algorithm. Each new vector
must be orthogonal to all previous vectors and have unit length. This
process can be done in any frame of reference.
The last strategy will be the active relative player effort strategy, u .
In the symmetric normal-form coordinate frame, this strategy is
286 Geometry, Language and Strategy—Vol. 2

orthogonal to each inactive strategy. In picking the corresponding


orthonormal direction  associated with the relative player effort, we
can therefore set E j  0 for each inactive strategy j .
Since the initial active strategy flows are known, to complete the
specification of the flow vector, we need only the magnitude of the time
component of the flow m and the player inactive strategy flows. We
start the inactive flows at zero. The players are uncoupled from the
decision process. This is not unreasonable if we assume that we start at
an equilibrium position q  0 . The inactive flows are determined from
Eq. (4.64):

 E j  q E j  2 E j . (7.29)

If there is a non-zero player interest  , the player will develop a non-
zero coupling away from equilibrium. We think this provides support for
the streamline solution choice that the electric field is zero, Eq. (6.15).
Given the initial flow directions and the values of the complete
symmetric co-moving orthonormal set at some initial position, we use
Eq. (6.46) to determine the symmetric co-moving orthonormal frame at
all other positions. These equations provide coupled differential
equations in the proper relative player effort  and can be uniquely
solved once the coefficients in the equations are determined. The
coefficients are the orientation potentials, which we deal with next.

7.8 Known Payoff Fields and Orientation Potentials

Given the initial conditions for the symmetric co-moving orthonormal


vectors, we have the differential Eq. (6.46) that determines the inactive
components at all other points of space assuming we know the scalar
functions for the acceleration q , player interests  , player bonding
components  and charges e . The latter however is not independent
based on Eq. (4.69). We solve for the unknown scalars using the field
Eqs. (6.47), (6.48), (6.49), (6.50), (4.69) and (6.17):
Prisoner’s Dilemma: A Code of Conduct Application 287

    2 q          ,


      q       2  
 p 
  p  ph  h  ,
 n 1 
 p    q     2     1 2     ,
  1  

  p
 q  q q    q  2       p  ,
 n 1 
 e   q e  2   e  ,
p  0 . (7.30)

We make model assumptions about the inertial components p and the
relationship between energy density  and pressure p (Sec. 7.10). To
complete our specification of the numerical problem for the field
equations, we determine the initial values of the player interest and
player bonding components. These we show are related to the payoff
fields.
To relate the initial values of these orientation potentials to payoffs,
we start with the normalized flow Eq. (7.27). We note that this holds in
any frame of reference and provides the basis for discussing the
prisoner’s dilemma in the normal-form coordinate basis, the symmetric
normal-form coordinate basis, the holonomic basis or the symmetric co-
moving orthonormal coordinate basis:

ˆ jk   jk
ˆ ja   jk Aak . (7.31)
ˆab  g ab   jk Aaj Abk

The metric potentials ˆ in the holonomic basis tie these different
descriptions together. So for example, we have on the left the metric
potentials in the holonomic basis, whereas on the right we have the
potentials that describe behaviors in the normal-norm coordinate basis.
In addition to the strategic flows V a , we know the initial behaviors of
the payoffs, which are related to the holonomic metric potentials.
288 Geometry, Language and Strategy—Vol. 2

Assuming the transformed basis from Sec. 7.5 and further assuming that
the metric potentials are independent of all coordinates except u t ,
we relate the metric potentials to the normal-form coordinate basis
potentials:

j , k  1 ,  2
m  r , s1 , s2
a, b  u , t . (7.32)
 aˆ jm   a  jk A   jk  a A   a jk A   jk F
k
m
k
m
k
m
k
am

 aˆ jb   bˆ ja   a jk Abk   b jk Aak   jk F kab

No result of the theory depends on the initial value of the vector


potentials Aaj  Amj  0 , so there is no loss in generality setting them to
zero; this is a consequence of gauge invariance. Similarly there is no loss
in generality in taking the initial metric potentials equal to the
Minkowski metric m (defined in Sec. 2.6).
We look only for the streamline solutions of Chap. 6, which have an
initial time derivative of zero. We thus obtain the following:

u ˆ jm  m jk F um
k

t ˆ jm  m jk F tm
k
0
. (7.33)
t ˆ ju  0
u ˆ jt  m jk F utk

Implicit in this description is that the normal-form coordinate basis and


co-moving basis be aligned using the Gram-Schmidt process described
above, Eq. (7.28). The symmetrized normal-form basis and the normal-
form basis are aligned, which proves the result. So, in this case initially
t     , time is along the flow (proper-time) direction and u   , the
relative player effort is along the proper active relative player effort
strategy direction. There will be small changes however since the initial
flow is not identically along the time direction. For orientation we ignore
these small effects but take them into account in our numerical
calculations. What is striking is that there are many gradients that are not
Prisoner’s Dilemma: A Code of Conduct Application 289

specified at all, such as  uˆ jk . They represent influences that impact the
dynamic behaviors but are absent from game theory models. We must
get better insight into such behaviors.
For the symmetric normal-form coordinate basis, we have the same
holonomic metric potentials ˆ , which are expressed in a form similar
to Eq. (7.31), with the noted differences in what is active and what is
inactive:
a, b  u , t
j , k  1 ,  2 , s1 , s2 , r
ˆ jk   jk . (7.34)
ˆ ja   jk Ak a
ˆab  g ab   jk A j a Ak b

Of specific interest will be the gradients whose payoff values we know


from Eq. (7.33):

j , k  1 ,  2
m  r , s1 , s2
 u  jm  m jk F um
k

(7.35)
 t  jm  m jk F tm
k
0
 t ˆ ju   jk  t Ak u  0
 u ˆ jt   jk F k ut   jm F m ut  m jk F utk  F k ut  F utk

We know the initial behaviors of the gradients  u jm for the indicated


restricted set of indices and we know the initial behaviors of F k ut for
each prisoner. We don’t know the corresponding electric field for the
symmetric inactive directions s1 s2 r .

7.9 Symmetric Co-Moving Frame Values

We take the initial behaviors that we have determined and compute the
consequences in the symmetric co-moving orthonormal coordinate basis.
290 Geometry, Language and Strategy—Vol. 2

We start with the electric field f j components from Eq. (6.46), which
are determined by the game values:

f j
   E t E u  E u E t  F tuj
. (7.36)
f j
  
 2 q E j    E j  e E j    e E j 1  E k  Ek  

To this we add the behavior for   jm that can be derived from Table 5.1
(Cf. Sec. 4.11, Ex. 23, Eq. (4.85) specialized to the streamline solution),
which are set by the initial payoffs for the prisoner’s dilemma:

 jm  m jk E u F um
k

. (7.37)
 jm  2q E j Em  2  E j Em  E j Em   2 E j Em 

These two behaviors depend on the acceleration q along the relative


player effort direction. To focus on the possibility of other equilibrium
behaviors, we assume that the initial acceleration is zero. To get a rough
idea of what is determined, we recall that we align the co-moving frame
so that prisoner j is along co-moving direction  j and assume the initial
prisoner flows are zero E j  0 . The latter determine the initial charges,
which we are taking as neutral or zero.
With these approximations, we see more clearly what is determined.
From the game value of the prisoner’s dilemma, Eq. (7.36) we determine
the player interest:

j  1 ,  2
j
. (7.38)
  1
2 F tuj

We determine the player interest because we have assumed the electric


field components   0 are zero in order to meet the conditions for the
streamline solutions. Given the player interest, from the payoffs of the
prisoner’s dilemma, Eq. (7.37), we determine the player bonding
components that lie along the mixed axes of the inactive player directions
and the equivalent player directions:
Prisoner’s Dilemma: A Code of Conduct Application 291

j , k  1 , 2
m  r , s1 , s2 . (7.39)
    1 2 F Em  1 2 m jk F
u j m
j
tu
k
um

In performing the numerical analysis, we use the initial transformation


matrix values and Eqs. (7.36) and (7.37). There are additional
components that we have not determined.
Among the many components not determined are the player
components of the player bonding matrix:

j, k  1 , 2
. (7.40)
u  jk  2u j k

The mixed term is an example of influence of one prisoner on the other


that is not part of any payoff matrix. Our strategy in general is to set
these new terms to zero unless there are clear reasons for not doing so.
At a later time the strategy is to do a sensitivity analysis on these
assumptions.
For these mixed terms, there may in fact be reasons not to set them all
to zero. One such reason that comes into play is the requirement that the
pressure, Eq. (6.49), be positive:

 p    q     12      1 2    . (7.41)

We take the initial acceleration to be zero. We must impose constraints


on the unknown player bonding and player interest components in order
to insure that this pressure component be positive. Our initial approach
will be to use the diagonal components of Eq. (7.40) to accomplish this.
We intend to investigate similarly the effects of all undetermined
components on the dynamic behavior as they provide valuable insight
into the dynamic mechanisms.
292 Geometry, Language and Strategy—Vol. 2

7.10 Numerical Values and Inertial Effects

The numerical results in this chapter will be based on the following


choice for the unit flow with components:

 t 1 2 N2 C2 N1 C1 
V   . (7.42)
1.061 0. 0. 0.2500 0. 0.2500 0. 

In the symmetric normal-form transformed frame:

 t 1 2 s1 s2 r u 
V  . (7.43)
1.061 0. 0. 0.1768 0.1768 0.2500 0. 

With these values, for the inactive space 1  2 s1 s2 r the player
interest and player bonding components are chosen to agree with the
game value Eq. (7.36) and player payoffs Eq. (7.37) using a   1 3 ,
which leads to a decision mass of m  4 and hence  0  4.243 :

  0.1871 0.1871 0. 0. 0.


 x1 0. 0.03706 0.6607 0.4660 
 
 0. x2 0.6593 0.05706 0.4660 
. (7.44)
   0.03706 0.6593 x3 0. 0. 
 
 0.6607 0.05706 0. x3 0. 
 0.4660 0.4660 0. 0. x3 
 

We arbitrarily set all the remaining player interest and player bonding
components to zero with the exception of the diagonal components
labeled x1 x2 x3  . We distinguish between prisoner 1, prisoner 2 and
the remaining inactive dimension. The pressure component Eq. (7.41),
with units   1 , is a function of three unknowns (Fig. 7.1):

p  1.2400  x1 x2  3x1 x3  3x2 x3  3x32 . (7.45)

It is clear that we must choose values of these unknowns that are large
enough or the pressure will be negative.
Prisoner’s Dilemma: A Code of Conduct Application 293

There are two disconnected surfaces, which are illustrated when all
three parameters are equal, x1  x2  x3 (Fig. 7.1). For our numerical
work, we pick a pressure component p   1 , which occurs when all
three parameters are equal to x1  0.4732 .
We can pick a value in which the pressure is uniformly zero.
However the gradient of the acceleration  q does not vanish since it is
equal to the vorticity squared, Eq. (7.30). A non-zero inertial acceleration
implies the existence of an inertial stress or pressure. The pressure
changes value at other strategic points.

Figure 7.1. Pressure contours Figure 7.2. Pressure p  .

There are important implications to our need for a non-zero pressure.


For a Nash-equilibrium, this would not be necessary. Our choice of a
non-Nash equilibrium solution requires the existence of inertial pressure
to support the stability of the solution with a non-trivial code of conduct.
Thus we obtain new types of stabilities at the cost of requiring sufficient
forces to support them. The code of conduct requires additional support
in the form of inertial contributions. We shall see that some of these
inertial contributions appear as gravitational effects pulling behaviors
towards areas where there is already a concentration of inertia.
There are a number of parameters associated with the organizational
system components of pressure p , stress tensors  p p  , and
energy density  that we have not yet specified. These represent
significant effects that are not addressed by the concepts of payoffs and
strategies. Following the organizational dynamics ideas from Vol. 1, and
taking into account Eqs. (6.20) and (6.49), we make the provisional
choice that the pressure and energy density are analogous to a perfect
294 Geometry, Language and Strategy—Vol. 2

homogeneous fluid for p . In the co-moving orthonormal frame we


make the following assumption:

 p  2        ph


 p     q     1 2      1 2       p . (7.46)
p  ph  0

The homogeneity of the fluid is expressed as:

 p. (7.47)

The resiliency  is a measure of internal energy, which in general


should be larger than unity. For our model calculations we take   5 .
We have specified a complete set of initial behaviors from which a
unique solution to the scalar field equations, Tables 5.4, 5.5 and 5.6, can
be obtained. Given the scalar fields, we can then obtain the inactive and
active transformations Table 5.1. In practice we carry out the solution for
these quantities simultaneously as coupled differential equations. We
present the stresses, strains and vorticity that result from decision process
theory in the next sections using Mathematica [Wolfram, 1992].
Dynamic behaviors will be introduced in Sec. 8.9 and subsequent
sections. They rely on the same scalar solutions, extended to include
non-steady state active strategies, Eq. (6.51).

7.11 Outcomes

The student will have learned that the prisoner’s dilemma demonstrates
that a player code of conduct can be imposed in decision process theory
and such hypothesis leads to a “stable” set of solutions. The theory thus
extends the notion of solutions based on Nash equilibrium in game
theory, which provide another class of “stable” solutions. By adding
three codes of conduct to the prisoner’s dilemma, the player fixed frame
model for the prisoner’s behaviors reduces to a single active strategy.
Complete numerical solutions to these models are obtained that represent
both steady state and non-steady state dynamic behaviors. These
Prisoner’s Dilemma: A Code of Conduct Application 295

solutions focus on the code of conduct, cooperative and inertial aspects


of the prisoner’s dilemma as opposed to the competitive aspects.
In this chapter, the student will have learned a behaviorist (rather than
a psychological) view of the prisoner’s dilemma and a formulation of
this view inside a dynamic theory. The view was limited to the player
fixed frame model and a single active strategy, which served to show that
applying a code of conduct leads to stable solutions.
The attainment of the outcomes of this chapter is facilitated by doing
the exercises in the following section. Based on this investment, the
student should achieve the more detailed outcomes below based on
section.
 In Sec. 7.2, the student will have learned the relevance of altruism and
egotism from Thomas & Kane [2008], where it is suggested that
empirical investigations of human participants presented with
prisoner-dilemma game situations yield results that contradict
standard game theory analysis (see Sally [1995] for a review). Some
of these results have motivated so-called “psychological” game
theories (e.g., [Dufwenberg & Kirchsteiger, 1998]; [Rabin, 1993]).
The hallmark of these psychological frameworks is that they attempt
to model players’ “fairness” or “kindness,” as well as each player’s
beliefs about whether his or her own actions will be reciprocated with
(un)kind or (un)fair actions. As Rabin [1993, p. 1281] notes in his
psychological model, the notion of altruism can bear on the notion of
fairness: “the same people who are altruistic to other altruistic people
are also motivated to hurt those who hurt them” (emphasis in
original).
 In Sec. 7.3, the student recalls the normal form of the prisoner’s
dilemma.
 This is followed in Sec. 7.4 with ways of writing the payoffs and
strategy vectors that are equivalent in game theory but whose
consequences may be dynamically different in decision process
theory.
 The concept of the code of conduct is introduced in Sec. 7.5. By
identifying certain active strategies as dynamically inactive, an
internal symmetry is introduced that is stable in the sense that small
changes away from that choice lead to small changes in behaviors.
296 Geometry, Language and Strategy—Vol. 2

This is a general property of the hyperbolic partial differential


equations of the theory.
 Hyperbolic partial differential equations have stable solutions given
an appropriate specification of scalars on a hyperplane transverse to a
timelike vector. The flow provides the timelike vector and the
hyperplane is a surface transverse to the flow. The appropriate scalars
are the transformations from the symmetric normal-form coordinate
basis to the symmetric co-moving coordinate basis. The student will
be able to set these appropriate values based on Secs. 7.6-7.10: the
student will be able to specify the known flows, set the values for the
orientation potentials for the symmetric prisoner’s dilemma,
transform these values from the symmetric normal-form coordinate
basis to the symmetric co-moving orthonormal coordinate basis and
deal with the specification of the inertial stresses.

7.12 Exercises

(1) Show that the orthogonal transformation S converts the


coordinates in the seven dimensions of time and space for the
prisoner’s dilemma to Eq. (7.17), where SS T  I :

 t 1  2 s1 s2 r1 r2 
 t 1 0 0 0 0 0 0 

 1 0 1 0 0 0 0 0 
 
2 0 0 1 0 0 0 0 
S . (7.48)
 N2 0 0 0 0 1
2
0 1
2

 
 C2 0 0 0 0  1
2
0 1
2
N 0 0 0 1 0 1 0 
 1 2 2

C 0 0 0  1 0 1 0 
 1 2 2

(2) Show that the new basis coordinates x a and the new payoff fields
F abj are determined by the transformation S in Ex. 1, where for the
payoff fields we restrict the transformation to exclude the player
inactive strategies:
Prisoner’s Dilemma: A Code of Conduct Application 297

x  S -1x
. (7.49)
F j  S -1F jS

(3) Show that the transformation from the basis of Eq. (7.17) to that of
Eq. (7.18) is accomplished through the orthogonal transformation
T below, TT T  I , with the transformations of the coordinates and
payoffs being expressed analogously to Eq. (7.49):

 t 1  2 s1 s2 r u 
 t 1 0 0 0 0 0 0 

 1 0 1 0 0 0 0 0 
 
2 0 0 1 0 0 0 0 
T . (7.50)
 s1 0 0 0 0 0 0 0 
 
 s2 0 0 0 0 0 0 0 
 r1 0 0 0 0 0 1  1 2
 2

r 0 0 0 0 0 1 1 
 2 2 2 

(4) In this section, we have chosen the code of conduct based on the
differences between the strategy to confess and the strategy to not
confess. Rewrite the model and find the appropriate transformations
if the code of conduct is to pick as inactive the strategy to confess
for player 1 and the corresponding strategy for player 2.
(5) Show that a model equivalent to the prisoner’s dilemma is that of
two players who have the choice to honor a contract or break the
contract. This scenario is significantly more general than the
prisoner’s dilemma and underlies the invisible hand of Smith
[1776].
Chapter 8

Prisoner’s Dilemma—Steady
State Geometry

Our view of decision making is that it is an active process with an


underlying set of steady state codes of conduct processes specified by the
existence of players and agreed codes with persistent behaviors. In this
chapter we describe the geometry of those steady state processes and
identify that geometry with the inactive metric components  jk , the
inactive flows V j and the inertial accelerations q . Differential
geometries help articulate our assumption that there is benefit to looking
at decisions in multiple ways. An analogy relevant to this discussion is
looking at wave phenomena on a lake.
One perspective is from shore, where the advantage is that one can
see the waves. In decision process theory this corresponds to the normal
form coordinate basis (Sec. 4.1). A second perspective is as a leaf
traveling with the waves. This corresponds both to the co-moving
orthonormal basis and the central holonomic frame (Sec. 4.6).
Differential geometry provides tools to move from one perspective to
another, as well as providing a formalism that allows general principles
or laws to be expressed in a common frame independent way. The
advantage of the co-moving frames is that we focus not on the speed of
the waves, an important attribute for sure, but focus instead on the steady
state geometry. Each perspective has advantages and no perspective has
all of them.
The steady state geometry consists of the rules governing the codes of
conduct and how players engage while making decisions. The geometry
provides the utilities for making such decisions, which are formulated in
the central holonomic frame by the inactive metric components  jk . The

299
300 Geometry, Language and Strategy—Vol. 2

strength of engagement is provided by the inactive flows V j . We also


have a measure of the value of the decision, which is provided by the
time component of the player payoffs, F j . Finally there are inertial
aspects of the geometry described by the acceleration or rate of change in
which decisions of a given type are made, q . We can have acceleration
even if all of the flows are zero. These aspects characterize the steady
state geometry. We can solve the decision process theory equations in
this frame and thus completely describe the steady state geometry.
As a method to gain insight into the solutions to these field equations,
we have added an additional (and equivalent) step. There is a trade off in
difficulty between having field equations in which time and active
strategies are holonomic (Sec. 4.1) with a metric that is not orthonormal,
and field equations in which the governing active and inactive metric
tensors are orthonormal but time and the active strategies are not
holonomic.
If we look at decisions from the perspective of the co-moving
orthonormal basis, we capture the essence of the steady state geometry
in terms of a number of coupled first order partial differential equations
in terms of orientation potentials that provide significant insight into the
geometry. These deconstruct the theory into its component parts (Chap.
5). In this frame of reference, the strengths of utilities between different
codes of conduct are provided by the player bonding tensors  . The
player engagement is described by the proper flows, e . The measures
of the game value for each player is provided by the player interest  .
The acceleration along the active directions is q . The transformation
between frames is worked out in detail in the exercises in Chap. 6: see
for example Ex. 19ff.
With the centrally co-moving hypothesis, the flows are zero by
definition, though in general, the acceleration along any active direction
is not zero. There are three main contributions to the central active
acceleration: inertial, payoff and bonding. In our player fixed frame
model, the central active acceleration can be expressed in the
perspective of the co-moving orthonormal frame orientation potentials,
(Sec. 6.10, Ex. 29):
Prisoner’s Dilemma—Steady State Geometry 301

q  2 e   e e 
Q  .
1  e e 

 2

This shows that the steady state geometry determines the forces that
change the active decision flow. The four fields that govern that behavior
are: the player engagement e , the inertial acceleration q , the player
interest  and the player bonding  . The centrally active
acceleration Q vanishes only if these contributions conspire to cancel.
In this chapter we project how these behaviors impact decision making,
as well as introduce a number of related concepts that provide further
depth of understanding for these concepts.
In addition to the decision theory aspects of these concepts, it is
helpful to introduce more mechanical aspects of these concepts that
stress the analogy of this theory to physical theories. In material objects,
we are familiar with the idea that the stresses and strains of matter are
related to dynamic behavior. We expect different vibration
characteristics between a very brittle object and a very ductile object;
different transient behaviors, for example.
Typically stresses are proportional to strains as a first approximation.
Stresses and strains are also familiar in electromagnetic phenomena,
though they go by different names. In this case we associate currents
with stresses and strains with electric fields. The relationship between the
two is expressed as Ohm’s Law: the current is proportional to the voltage
(related to the electric field). In decision process theory, which has many
similarities to electromagnetic theory, we also have relationships
between stresses and strains, though the steady state geometry
relationships are more complex.
We explore these aspects of steady state geometry for the prisoner’s
dilemma. In Sec. 8.1, we focus on the stresses, which manifest as inertial
pressure and provide the input for computing the accelerations. We build
off the idea of the organizational system from Vol. 1. We show model
results in Sec. 8.2, with sensitivity analysis provided in Sec. 8.3. The
geometry is further characterized by the strengths of the codes of
conduct, expressed in terms of the orientation potentials, which can also
be characterized as strains. These behaviors are described in Secs. 8.4-
302 Geometry, Language and Strategy—Vol. 2

8.6. The way in which players interact is governed by the player


engagements, Sec. 8.7. A significant advantage to working in the
orthonormal coordinate basis is the ability to show more explicitly the
interrelationships between all of these potentials, Sec. 8.8.

8.1 Inertial Stress Behaviors

In this section we show that inertial effects represented by the


organizational system stress tensor are needed to support solutions to the
prisoner’s dilemma in which the prisoner’s cooperate for their common
good; their public-interest. Such effects are shown not to be needed to
support solutions in which players operate only on their self-interest.
These results provide important insight into the decision making process.
The inertial effects are needed to resolve the paradox of the prisoner’s
dilemma. More generally they lend support to the importance of
contracts (Sec. 11.2) as the necessary ingredient for the invisible hand
[Smith, 1776] that is implicit in game theory arguments.
In decision process theory, we identify both the forces that pull the
system together as well as the forces that push the system apart. In this
section we demonstrate that there is a type of gravity, which is a result of
the time isometry that generates an attractive force that brings things
together. The organizational media generates an opposite force, which is
a consequence of energy being concentrated, generating stress (pressure).
The stress tensor represents internal forces and holonomic constraints,
Sec. 1.6. It allows us to take into account these internal forces with a
principle of least action, Eq. (1.6), which focuses on the applied forces,
incorporating the internal forces into the energy momentum stress tensor.
Decisions occur within an organizational system that is the origin of the
internal forces. We believe this way of taking constraints into
consideration for decisions is new.
We demonstrate that we can engineer structurally stable behaviors by
insuring that there is sufficient inertia to generate attractive forces to pull
the system together and overcome any destabilizing forces. What is the
source of this inertia? We suggest it is the rules of the game, the
agreement to adhere to certain standards, to a code of conduct, to
Prisoner’s Dilemma—Steady State Geometry 303

standards of behavior [Von Neumann & Morgenstern, 1944].


Organizational systems are complicated, with many positive and
negative feedback loops that make them look like an elastic system, such
as a physical fluid. It means they don’t like to change based on their rules
of the game.
We turn first to the numerical results based on the known conditions,
Sec. 7.6. We view the prisoner’s dilemma decision process as one that
repeats. It is a set of plays in which each prisoner can wager, but only the
difference in wagers u is an active strategy. Each prisoner learns from
the previous plays and so each prisoner comes to a conclusion about
what to do next. The decisions are more complex than simply having a
payoff matrix, though that is part of each prisoner’s consideration.
One important aspect of the steady state geometry is the inertial
acceleration q that is determined by the energy density and stress
tensor. These in turn are determined by the strains and their gradient
variation as a function of the active strategy as well as the vorticity: the
turns, twists, rotations that are possible without generating strains. We
think of fixed point behavior as associated with a point at which there is
zero acceleration. The question of structural stability is then whether the
system as a whole would move towards, away or around this point if the
restrictions based on the steady state hypothesis were removed.

Figure 8.1. Pressure p versus  . Figure 8.2. Acceleration q versus  .

8.2 Model Results

The general decision process theory Table 5.6 identifies in Eq. (3.42),
the change in inertial acceleration q , which is also the acceleration
304 Geometry, Language and Strategy—Vol. 2

determined by the source energy momentum tensor. We made the


assumption that the pressure was initially non-zero. Based on the known
behaviors (Sec. 7.6), the solution to the field equations leads to a
pressure that is peaked at the origin,   0 , Fig. 8.1. We emphasize that
this result is a consequence of the field equations and the known
conditions and not an assumption. For reference, we call this the
Prisoner Dilemma Model (PDM).
Our steady state solution for pressure demonstrates its tendency to
bunch. We note that the pressure is not symmetric, but drops off more
dramatically on the left (prisoner 1 is more active) than on the right
(prisoner 2 is more active). The origin of the asymmetry is that there are
two possible solutions of pressure in Fig. 7.2 for any given positive
pressure. Our treatment is symmetric in the sense that there is another
solution x1  1.497 that is the mirror of Fig. 8.1. Nevertheless each
solution demonstrates an asymmetry.
We are not surprised that the pressure has a maximum at the origin.
The assumption that the origin was a point of zero acceleration defines
the position of the peak for pressure. Moreover, since the energy density
is proportional to the pressure, it too peaks at the origin and has the same
shape. The consequence of the strong pressure peak at the origin is that
the acceleration Fig. 8.2, which is roughly proportional to minus the
gradient of the pressure, Eq. (5.56), will be positive when prisoner 2 is
more active (on the right) and negative when prisoner 1 is more active
(on the left).
If we think of the inertial acceleration as being balanced by the
gradient of an inertial potential  q that represents the attraction that
decisions tend to follow previous decisions, then the potential should
represent an attractive well, Fig. 8.3:

q   q . (8.1)

Inertia produces many effects, one of which is the one describing


acceleration. It is analogous to the physics potential in that the negative
gradient of the potential is the force, which is proportional to the
acceleration. In Sec. 8.5 we shall see that inertia determines the volume
effects. When we compute the time component of the metric potential gtt
Prisoner’s Dilemma—Steady State Geometry 305

(Sec. 9.3), we shall see that the charge density (player engagement)
determines the initial shape. The charge density is indirectly determined
by the inertial effects. There are many types of effects, clearly indicating
that the inertial effects are not scalar but tensor forces.
At this point we observe that in the co-moving orthonormal frame,
the spatial components of the flow are zero, which leads us to Eq. (8.1).
This equation and the shape of the potential well, Fig. 8.3, support our
argument that inertial effects induce stability. We believe this is a crucial
aspect of decision making when there is a non-trivial code of conduct.
Decisions will collect together; where
they collect creates an equilibrium
point. Because they collect they attract
subsequent decisions towards the same
point in a way analogous to
gravitational attraction. This
equilibrium point is determined by the
code of conduct, not by considerations
Figure 8.3. Inertial potential  q of utility, self-interest, or even public-
interest.
In decision process theory, the attribute that the pressure bunches
around the origin does not hold in general. For these symmetric solutions
of the prisoner’s dilemma, it holds for initial pressures down to a critical
pressure pc  0.02693 defined as the point at which the acceleration
gradient vanishes, which is proportional to the scalar formed from the
player interests,    .
Below this pressure the equilibrium point moves away from the
symmetric center. The cause is traced to the presence of the centrifugal
effect of  . In the next section, we investigate this in more detail, since
this provides further insight into the source of stable behaviors.

8.3 Sensitivity Analysis

We consider several cases that modify the PDM (Sec. 8.2), each with an
initial pressure of 0.01, a value that is below the critical pressure. We
vary the input flow keeping the inertia fixed at the model value of 4. We
306 Geometry, Language and Strategy—Vol. 2

call the flow in which both prisoners choose not to confess the common
good solution. The Nash equilibrium flow where both prisoners confess
we call the Nash solution. We consider these two solutions, with and
without the non-zero game value identified in Sec. 7.5.
For the Nash solution, the shape of the pressure for zero game value,
Fig. 8.4, has a peak at the origin. This demonstrates that nothing prevents
the bunching of the energy density and pressure around zero, because the
Nash solution has player interests  that are almost identically zero
when there is no game value. We find that the Nash solution is
structurally stable for zero game value in the simple sense of attraction.

Figure 8.4. Pressure for Nash solution: Figure 8.5. Pressure for common good
zero game value solution: zero game value

In contrast, the pressure for the common good solution, Eq. (7.42),
has a minimum at the origin, Fig. 8.5. It demonstrates that with the same
assumptions and no game value, the player interests lead to a stable
solution based on the pressure gradient. Again, extra energy is expended
to insure the steady state behavior. The corresponding potential well
Fig. 8.3 turns into a potential mountain. We conclude that there is not
enough inertial mass to balance out the centrifugal effect ascribed to the
game, Fig. 8.6, though there is sufficient pressure gradient. We thus find
evidence that this solution is not structurally stable.
Prisoner’s Dilemma—Steady State Geometry 307

We thus agree with the assessment that the common good solution is
possibly less structurally stable in the absence of inertial effects.
However, we showed that with sufficient initial energy density, the
instability can be overcome, Fig. 8.1. The source of the sign change is
clearly seen in Eq. (7.7). Whenever there is a player interest, the initial
pressure has to be large enough to overcome the effects or the origin will
be unstable.

Figure 8.6. Common good solution for Figure 8.7. Nash solution for
1   : zero game value 1   : non-zero game value

There are other sources to a non-zero player interest, one of which is


the existence of a non-zero game value. For the Nash solution, the player
interest Fig. 8.7 is about an order of magnitude smaller than the effect
due to the common good solution. Nevertheless it is there and the
pressure has to be large enough to overcome this. The critical pressure
here is much smaller. Although for an initial pressure of 0.01, the
pressure has a maximum at the origin, for a pressure sufficiently small,
the maximum turns into a minimum.
As the pressure increases, the shape of the potential well,  q , Eq.
(8.1), narrows. In other words, we increase the stability. As the pressure
decreases towards the critical pressure, we decrease the stability. There
are other inertial parameters, the inactive component stresses p , p
and the energy density    p , which further refine the shape of the
potential well. The critical pressure depends not only on the initial
(average) pressure p but on the resilience  . For now we don’t have
sufficient information to further refine our understanding of these effects
and so leave them to a future study.
308 Geometry, Language and Strategy—Vol. 2

In this model case, we see the consequences of   0 . With this


assumption, we see that a non-zero game value induces instability
because it leads to a non-zero value of the player interest. If we have a
decision process with a very high game value, this process will be
intrinsically more unstable than a process with a smaller value. We
associate a very high game value with high stakes and high risk. This is
less stable than a small game value.
This completes our sensitivity analysis and we return to the features
of the model defined by the parameters of Sec. 7.6. Our conclusion from
this section is that when we are sufficiently above the critical pressure,
the inertial energy density provides structural stability that counters two
potential destabilizing effects: a non-zero value for the game and any
flow that is not a Nash equilibrium solution. We investigate metrical
behaviors of the solution in the next section.

8.4 Strain Behaviors

It is not universally accepted that human behavior is subject to the same


quantitative discipline that has been applied to the physical sciences.
However, the game theory of Von Neumann & Morgenstern [1944] is
built on this premise. These authors argued that utilities determine the
payoffs and are not only ordinal (preferences can be ordered) but
cardinal (numerical values can be assigned in a meaningful way). We
agree with them that numerical measures must be part of decision
process theory. We believe that utilities can be meaningfully defined that
are more than a mere ordering of preferences but a way to distinguish
what strategies are near or far from each other. We go beyond game
theory by requiring that utilities are convertible (Sec. 13.5), which leads
us to concepts of space and time that are non-Newtonian. We use the
metric in a general sense to define near and far.
We make the differential geometry distinction, as does Einstein in his
theory of relativity, between the position of points in space and their
distance. The positions of points represent the location of strategic
choices in space and time. In a mathematical sense these points have a
fixed place in the topology of space and time. A topology provides the
Prisoner’s Dilemma—Steady State Geometry 309

concept of what it means to be neighbors but is not sufficient to


determine the concept of distance. The existence of a measure for
distance requires a physical field that provides the connection between
points in space and time. This connection provides additional structure to
the topology. It carries energy and momentum; it is affected by other
physical fields and may change in time and vary in space. The
connection is part of the physical world distinct from the topological
backdrop. The connection is part of the overall schema that allows
utilities to be convertible.
In the physical world, the connection between points is carried out
with yardsticks and clocks. In decision processes metrical measures
survive that derive from these basic yardsticks and clocks that allow us to
quantify choices. We demonstrate this in our theory with the prisoner’s
dilemma as the teaching example. In the last section we focused on the
inertial stresses that arise in the theory. We showed that inertia provides
a fundamental mechanism for endowing the system with energy density
and pressure, which is the average of the diagonal components of stress.
We use the word stress in a way analogous to physical theories to
represent the forces present. An independent concept is the idea of
strain, which highlights the effect on the geometric configurations that
measures displacements of the system in space and time. A bridge
undergoes stress when a heavy object goes across it. The stress generates
strain as evidenced by parts of the bridge that move as a consequence.
The strains are in response to the stresses. The notion of strain makes
sense for decision processes because changes in the strategy
configuration of the system are observable, distinguishable and
measurable. In this section and the next we demonstrate that the
electromagnetic field behaviors Table 5.4 and tidal behaviors Table 5.5
define not only the strains such as  associated with movement along
the active direction  , but through the coordinate transformations they
define the metric. The inactive metric components  jk will make precise
the notion of cardinality. In this section, we explore the properties of the
strains in the inactive space.
Our numerical work, again using the PDM (Sec. 8.2) is presented
based on expanding the player bonding component  into a player
bonding coefficient  and player bonding shear components   :
310 Geometry, Language and Strategy—Vol. 2

  h 
    1
ni  h . (8.2)
ni  h 

This decomposition is not only a standard way of analyzing the


properties of a symmetric matrix, but has been used to understand the
complex field equations of Einstein [Hawking & Ellis, 1973]. Any such
matrix can be thought of as an operator that transforms the shape of a
small cube; since the matrix changes as a function of the active strategy,
the shape of the small cube will also change as a function of position.
The two types of changes correspond to these distinctions.
We can use these ideas to characterize the volume of the cube. The
player bonding coefficient leads to a change in the volume:

Volume  Volume  Volume  . (8.3)

The bond shear components leave the volume unchanged but distort the
shape of the cube by sliding a face of the cube parallel to one of the other
faces in such a way that the volume does not change. It may help to think
of a cube of Jell-O as something with a fixed volume and consider
different distortions to this. These ideas give a physical picture of the
strategic strains on a small cube that is moved along an active strategy
direction in the co-moving orthonormal frame.

8.5 Player Bonding Results

For our streamline solutions of the prisoner’s dilemma, we obtain the


following equation (Cf. Ex. 2) demonstrating the close relationship
between the stress (pressure) and the strain (player bonding):

      q   2       p  . (8.4)

In the absence of player interests, such as we saw above for the Nash
solution with no game value, the player bonding depends only on the
energy density, acceleration and stress. For the numerical solution of the
Prisoner’s Dilemma—Steady State Geometry 311

prisoner’s dilemma, we start with the initial conditions, Eqs. (7.36) and
(7.37). This gives the value   0   2.366... . The shape of the player
bonding (Fig. 8.8) is modified only slightly by the player interests.
The player bonding gives direct evidence of changes in the
configuration as a consequence of the energy density and active stress
component. If we move a unit cube in the five inactive dimensions along
the active strategy, we find that the volume changes as it moves left or
right (Fig. 8.9). In particular we see that the size of the box goes to zero
as we go further left in the direction in which prisoner 1 is more active.
As we go to the right, the size of the box becomes constant. The
asymmetry of the volume corresponds to the asymmetry noted earlier
about the shape of the pressure (Sec. 8.2). For our numerical example,
we have chosen to focus on solutions in which prisoner 2 is typically
more active.
The idea that the unit of measure changes depending on where we
are, illustrates that in decision process theory, the concept of space and
its ordinal properties are distinct from the concept of distance and its
cardinal properties.

Figure 8.8. Player bonding gradient Figure 8.9. Player bonding volume as
      (PDM) a function of strategy  (PDM)

8.6 Shear Results

The initial sources of strain are the initial payoffs, Eq. (7.39). In addition
to the player bonding, these strains generate what we characterize as
player bonding shear in which the shape changes without rotation but the
312 Geometry, Language and Strategy—Vol. 2

volume remains the same. To simplify the interpretation of the numerical


results, the equation for the shear can be written in a reduced form, Ex. 5:

  e q   
. (8.5)
e q     2    2
ni h      p  1
ni h p  

The form factor e q , (Fig. 8.10), depends only on the inertial terms. For
the case of the numerical example with a perfect fluid, the reduced shear
simplifies:

e q   2    2 ni h   . (8.6)

The reduced shear depends on the inertial factors through the form factor
e q . In particular, when the player interests vanish, the reduced shear is
constant.

Figure 8.10. Form factor e q with Figure 8.11. Reduced bond shear
initial value normalized to unity (PDM) 1q2   (PDM)

For our choice of parameters of the prisoner’s dilemma, and with


pressure components for an ideal fluid, the player interests are small
and so the reduced shear is approximately constant (see for example
Fig. 8.11). Note that the addition of non-ideal stresses p and p ,
Eq. (8.5), can change this.
The initial conditions set the initial value of the bond shear
  0    0  and reduced shear at zero relative player effort,
Eq. (8.2):
Prisoner’s Dilemma—Steady State Geometry 313

 0 0 0.03705... 0.6606... 0.4659... 


 
 0 0 0.6592... 0.05705... 0.4659... 
 0.03705... 0.6592... 0 0 0  . (8.7)
 
 0.6606... 0.05705... 0 0 0 
 0.4659... 0.4659... 0 0 0 
 

The off-diagonal elements can be expected to change slowly and


represent the payoffs for the prisoner’s dilemma as viewed in the
effective process in which there are five players and a single active
strategy.
The payoff values change due to the field equations, so for example
we have   0.1 :
 0.003865... 0.006452... 0.03687... 0.6608... 0.4662... 
 
  0.006452... 0.003865... 0.6594...  0.05688...  0.4662... 
 0.03687... 0.6594... 0.002579... 0.000006525... 0.00001007...  . (8.8)
 
 0.6608... 0.05688... 0.000006525... 0.002580... 0.000009774... 
 0.4662... 0.4662... 0.00001007... 0.000009774... 0.002571... 

One measure of the shear effects is    . It has a maximum at the
origin and is roughly constant over the whole range. In addition, shear
effects lead to changes in the payoff tensor. In other words, due to
successive plays of the decision process, each prisoner adjusts his choice.
For steady state behaviors, this will manifest itself as a spatial
dependence. For the values chosen here for the prisoner’s dilemma, the
changes are primarily those due to the change in the form factor,
Fig. 8.10, as opposed to changes in the value of the reduced shear, such
as Fig. 8.11.
There are 14 independent bond
shear components, of which six we
have identified as being related to the
payoffs of the prisoner’s dilemma.
That means there are eight
components that measure something
new, something that directly relates to
the strain configurations that result Figure 8.12. Reduced bond shear
    (PDM)
from the assumed stresses. 1 2
314 Geometry, Language and Strategy—Vol. 2

For example, consider the mixed component  12 between the two
prisoners, Fig. 8.12. We think of the prisoners as interacting only by
means of their respective payoffs. However, decision process theory
provides influence through other means, such as through the bond shear
components 12   , Eq. (8.8). Though zero at zero relative player
effort, it shows a small but non-zero effect at other values of relative
player effort.

8.7 Player Interest Results

In addition to player bonding shear and player bonding, we have tidal


player interests. The initial player interests for prisoner 1 and 2 are set by
the game values, Eq. (7.38) and the remaining values are set to zero.
These initial values in particular determine the tidal player interests  ,
Fig. 8.13. The initial directions  are set by the directions
1 2 s1 s2 r after orthogonalization by the process described in
Sec. 7.7. By assumption only the prisoner’s player interests along
1 2  are non-zero at zero relative player effort. We have noted that
the non-zero player interests strongly influence the inertial properties of
the acceleration q , Eq. (7.7) as well as the bond shear Eq. (8.6). From
Fig. 8.13 below, we see the effects are particularly strong when prisoner
1 is more active,   0 .
We see that steady state behavior results from strains that, under
special conditions, exactly cancel each other. Just as in mechanical
engineering, we conclude that forces lead to dynamic behaviors. Stresses
lead to strains and hence changes in configurations. Changes in
configuration can be minor or catastrophic. We gain insight into those
stresses from an analysis of the steady state behaviors. In Sec. 8.1, we
identified that the forces based on the payoff behavior exist side by side
with inertial forces that are significant contributors to equilibrium
behaviors. In this section we see that stresses induce strains in the
configuration of the system. These strains determine the size of the
yardstick by which one measures how near or how far apart two strategic
choices are. In this way we capture the variable nature of utility and deal
with how to convert the measure of utility from one point in space time
Prisoner’s Dilemma—Steady State Geometry 315

to another. In the next section, we analyze the persistent behaviors that


result from these stresses and strains, and determine the normal form
coordinate basis metric potentials that result.

8.8 Persistent Behaviors

Persistent behaviors, Table 5.1, reflect the basic nature of a player or


agent as defined in Chap. 4. The notion of persistence comes from the
metric and the concept of isometry. A metric is the measure of distance
between neighboring points, Eq. (2.22).
In the last section we highlighted aspects of our theoretical
foundation that require the measure between points to change. In general,
any transformation of coordinates will cause such a change. However,
there may be special transformations that leave the metric components
unchanged: an isometry is a transformation of coordinates that leaves the
metric components unchanged.
Our notion of player (agent, prisoner, etc.) is based on the existence
of an associated isometry. We recall that for each player or agent, the
isometry is defined by the existence of a vector field K  satisfying the
Killing relationship Eq. (4.22). Equivalently, there is an associated
dimension  j and a holonomic frame in which all metric components are
independent of this dimension. We showed that the persistency
components E j E j E j  , Eq. (4.61), are in fact the components of
such a Killing vector in the co-moving frame. The behaviors of these
components in the co-moving orthonormal frame completely
characterize the persistency of that player including the inactive metric
that is the measure of distance in the inactive space:

 jk  E j Ek   E j h Ek  . (8.9)

There are three categories to consider: the inactive behaviors E j , the


active behaviors along the flow E j and the active behaviors along the
active strategy E j . In the player fixed frame model the latter
components are zero, E j  0 . In this section, we analyze the non-zero
components for the prisoner’s dilemma.
316 Geometry, Language and Strategy—Vol. 2

Figure 8.13. Player interest field Figure 8.14. Electric field f j  
   in the co-moving frame in the co-moving frame (PDM).
(PDM).

Figure 8.15. Player engagement e   Figure 8.16. Diagonal components


in the co-moving frame (PDM) ln  jj   in the normal form frame
(PDM)

The initial behavior of the persistency components result from the


initial flows and from the initial orthonormal set described by the Gram-
Schmidt process, Sec. 7.7. We compute the player engagement and
electric field components from Table 5.1. These values relate to
persistency. The persistency equations are determined by the inertial and
tidal behaviors, Table 5.1, which have been computed numerically in
Secs. 8.1 and 8.4. We start by checking that our solution meets the
known conditions, Secs. 7.8-7.10.
Prisoner’s Dilemma—Steady State Geometry 317

8.8.1 Electric Field Components f j




In Eq. (7.38), we argued that the game value sets the size of the player
interest. More precisely, the game value sets the size of the electric field
components f j , Eq. (7.36):

f j
   E t E u  E u E t  F tuj
.
f j
  
 2 q E j    E j  e E j    e E j  1  E k  Ek  

We check that the electric field components (4.74) in the co-moving


frame, Fig. 8.14, match their initial values:

f j
  0.4242..., 0.4242...,0.1882...,0.1882...,0.2662... . (8.10)

The order represents j in the directions 1 2 s1 s2 r . The


curves for s1 and s2 are on top of each other. The electric field is a
derived field from the player interests  , player engagements e and
player bonding  . Based on the sizes of the initial values, we can use
the initial value of the electric field F tuj to set the value of the player
interests.

8.8.2 Player Engagement Components e

We call  player interests since from Table 5.1, Eq. (4.69), they
determine the behaviors of the player engagement e , Fig. 8.15. The
player engagements determine the mixed metric attributes, Eq. (4.51)
that determines the relationship E j  e E j between persistency
components, Eq. (4.54). This relationship shows that the player
engagements are persistency attributes and demonstrates that the initial
values of the player engagements are set by the initial values of the
orthonormal set. Our choice of zero flow for the two prisoner directions
translates to their corresponding player engagements being zero at the
origin. The other player engagements are not zero, as seen in Fig. 8.15.
The player engagements also contribute to the active metric, Eq. (4.72),
which we deal with in Chap. 9.
318 Geometry, Language and Strategy—Vol. 2

8.8.3 Inactive Metric Components  jk

In Sec. 8.5 we showed that the player bonding volume decreased as the
active strategy  becomes negative in the co-moving frame. We can
show that this is also true in the normal form coordinate basis by looking
at the behavior of the diagonal components of the inactive metric
components  jj using Eq. (8.9), Fig. 8.16. As before, j has the
directions 1 2 s1 s2 r . We see that the diagonal components are
all tending towards zero as we move left into the region that prisoner 1
becomes more active. The initial conditions were set assuming that the
co-moving frame and normal form frame were aligned at the origin.
The gradient of the metric potential determines the off-diagonal initial
values of the bond strain, Eq. (7.39) in terms of the known payoff values
of the prisoner’s dilemma:
 0.9465... 0. 0.1414... 1.2727... 1. 
 
 0.  0.9465...  1.2727... 0.1414... 1. 
 jk   0.1414... 1.2727... 0.9761... 0.02958... 0.04183...  . (8.11)
 
 1.2727... 0.1414... 0.02958... 0.9761... 0.04183... 
 1. 1. 0.04183... 0.04183... 1.005733... 

The initial off-diagonal gradient components are set by the payoffs


Eq. (7.35), while the diagonal components are set by the player bonding
coefficients Eq. (7.40) and show the shrinking of the space to the left
compared to the right (Fig. 8.17). We suggest that with all other variables
constant, a change along the utility for prisoner 1 represents a distance
measure of d 2   11 d1d1 . Since the metric components go to zero on
the left, we conclude that for two neighboring preferences at the origin
compared to two neighboring preferences a distance to the left, there will
be a perceived difference in distance as given by the connection
components. These persistency components provide a mechanism for
comparing utilities.
These results follow from a detailed quantitative analysis of the
dynamic forces at play in decision making processes. In particular, they
show that the payoffs, which determine the relative strategies players
choose, vary as a function of the strategic value chosen. This mechanism
is a key ingredient of decision process theory, one not found in game
theory. The variation of the gradient of the metric (Fig. 8.17) is clearly
Prisoner’s Dilemma—Steady State Geometry 319

seen to produce a variation in the payoffs, Eq. (7.35). Despite these


variations as a function of position, the metric potential gradients are
steady state as a function of proper-time.

Figure 8.18. Flow components of the


Figure 8.17.  ln  jj   in the normal
Killing vector E j   (PDM)
form frame (PDM)

Figure 8.19. Prisoner 1 inactive Killing Figure 8.20. Prisoner 2 inactive Killing
vector components E1    (PDM) vector components E2    (PDM)

8.8.4 Inactive Flow Components E j

The behaviors of the metric potential are set by the persistent frame
transformation components in Eq. (8.9). We gain further insight by
looking in more detail at the behaviors of these persistency components.
The flow components V j  E j as seen in the normal form coordinate
320 Geometry, Language and Strategy—Vol. 2

basis are determined, Table 5.1, Eq. (4.64), by the acceleration and the
player interests.

 E j   q E j  2  E j .

Since we take the initial acceleration to be zero, the initial gradient is


determined by the player interests. Because we start with player interests
for the two prisoners that are non-zero, we obtain non-zero flows away
from zero, Fig. 8.18.
The flow components can be interpreted as the coupling or charge
with which the payoffs contribute to the streamline equations in the
normal form coordinate basis, Eq. (2.49). These flows correspond to the
player engagements, which are the conserved charges e . If we relax
the condition in this model that there are codes of conduct strategies that
are inactive, the corresponding flows will no longer be conserved. We
get the important result that the players couple strongly away from the
equilibrium point, which shows that game theory effects are strong away
from such points. The charges for the two prisoners are opposite and
equal for the model parameters chosen, though the equality is not
significant since slightly changed initial conditions makes this equality
disappear.
The flow for positive values of  corresponds to player 2 being more
aggressive: this player is more active than the other based on the active
variable u in Eq. (7.18). The model consequence is that player 2
demonstrates a greedy (negative) charge whereas player 1 demonstrates
an accommodating (positive) charge. We view this result as important as
decision process theory derives from game theory so we expect to see
areas where game theory mechanisms are important.

8.8.5 Inactive Transformation Components E j

The spatial persistency components E j that determine the


transformation of the inactive strategies from the normal form coordinate
basis to the co-moving basis are determined by the player bonding and
shear, Table 5.1, Eq. (4.67):
Prisoner’s Dilemma—Steady State Geometry 321

 E j    E j .

We get a rough idea of what is going on by assuming the player bonding


is diagonal and itself the gradient of a potential,    .
In this case the equation can be solved:

E j  E0 j e .

This provides us with a direct interpretation of the steady state geometry


metric potentials: they are determined by these frame transformations:
  
 jk  E j Ek  E0 j   E0 k  e .

If the charges are small, the second term dominates and the geometry is
determined by the diagonal components of the player bonding. If the
player bonding shear is small, the components are all equal and we
expect the diagonal components of the metric to be approximately equal
to  2 ni  , which is obtained by solving     . For the numerical
model, we provide the solution, Fig. 8.21. If we scale the curve by 2 5 ,
we see the qualitative behavior of Fig. 8.16. It is not exactly true since
we have ignored the effects of the charges E j .
We have already seen how the
player bonding effect decreases the
size of the transformation cube as we
go to the left. Some evidence of this
behavior is also seen in the
components E1 for prisoner 1 in
Fig. 8.19, where the size of the proper
direction for each player decreased in
Figure 8.21. Player bonding 
the co-moving frame. Similarly in Fig.
(PDM)
8.20, we have the corresponding
components for prisoner 2. In addition to the player bonding effects, the
shear effects impose behaviors along the other directions.
Based on Fig. 8.21, we see that whenever  gets large and negative
(the player bonding potential  is large and positive), we enter a regime
in which the diagonal components of the inactive metric  jj goes to zero.
322 Geometry, Language and Strategy—Vol. 2

These expectations are met from the full solution, Fig. 8.16. So, as
proposed in Sec. 8.8.3, if we think of these contributions as measuring
the utility between separated points, then the same utility occurs to points
with larger and larger spacing, which suggests the concept of inelasticity
from economics. Conversely, in the other direction the player bonding
changes sign and levels off, suggesting that the measure of utility has a
limiting value and depends directly on the point spacing. We think this
makes some sense if you interpret this to mean that the more players
bond, the less players will feel the urge to compete.
A second point worth noting is that the player bonding exists not only
between different players, but also for the same player as discussed
above. For the numerical example under consideration, there is not much
variation within or between different players. This is an artifact of the
initial values chosen. There are situations where these values are quite
different. Moreover, in general the player bonding will be different for
different strategic directions: in this model we have only a single active
direction. Thus decision makers may make different choices and assign
different utilities to their choices depending on where they are at in the
strategic space. We see evidence for this in Fig. 8.17.

8.8.6 Inactive Determinant  jk

Certainly a main conclusion from our analysis is the striking


configuration changes that depend on the active strategic value. This
was exhibited in the co-moving frame by the behavior of the player
bonding  , Fig. 8.21. Another measure of player bonding in the
inactive space is the determinant of the inactive metric:

  det  jk   jk . (8.12)
Prisoner’s Dilemma—Steady State Geometry 323

This shows the same qualitative behavior, Fig. 8.22. The invariant
volume element is proportional to this
determinant. Since this determinant
goes to zero as we go left, it confirms
that the yardsticks in the inactive
space shrink in that direction.
It also shows that the invariant
volume shrinks if we go sufficiently
far to the right. This is not so obvious
Figure 8.22. Determinant of the from the behaviors that we have seen.
inactive metric  jk   (PDM) A more detailed analysis shows that
one of the eigenvalues of the metric is
tending towards zero. Furthermore, the stresses are going to zero as we
go far to the right.

8.9 Outcomes

In this chapter we have examined the steady state geometry that is a


consequence of the steady state hypothesis. We are able to provide
complete numerical solutions for the player fixed frame model, which we
do in the co-moving orthonormal coordinate basis. With this perspective
we have identified the ingredients that determine steady state geometry
as inertia, stress, strain and persistence.
There are a number of ways that the steady state geometry influences
decision making: one key way is that it determines the centrally active
acceleration as seen in the central co-moving frame:

q  2 e   e e 
Q  . (8.13)
1  e e 

 2

Game theory style equilibrium is achieved if this acceleration is zero. We


have provided insight into how this can happen as well as how the field
equations conspire to make it non-zero.
324 Geometry, Language and Strategy—Vol. 2

For example with zero player engagements, the only contribution to


the acceleration is from the inertial acceleration, which we can take to be
zero at a point. Solutions that have this attribute at say   0 no longer
have this attribute away from this point. We have examined the key
concepts that influence the steady state geometry: player engagement
e , inertial acceleration q , player interest  and player bonding
 . Together these concepts describe in detail how the codes of
conduct influence active decisions.
In this chapter we have also introduced a number of related concepts
that give more depth to our conversation about the steady state geometry.
Through differential geometry, we have identified the close relationship
of these ideas to standard treatment of stresses and strains; such ways of
thinking may at times be helpful in furthering our understanding. We
have utilized to some extent different perspectives: in particular we work
in co-moving frames of reference where much of the active dynamics is
hidden. Such traveling wave views help us focus on the codes of conduct
behaviors.
The attainment of the outcomes of this chapter is facilitated by doing
the exercises in the following section. Based on this investment, the
student should achieve the more detailed outcomes below based on
section.
 From Sec. 8.1, the student should see that external constraints are as
important in decision theory as they are in physical theories. There are
a number of effects that we don’t explicitly take into account, though
such effects carry energy and momentum. We have characterized such
effects as energy flow and momentum stress flows. The latter leads to
its average as the inertial pressure.
 From Sec. 8.2 and the sensitivity analysis, the student should see that
this inertial pressure plays a significant role in decision making. The
student should be able to determine when inertial effects are needed to
support a code of conduct or non-zero game values. To obtain a
solution of the prisoner’s dilemma that supports the common good of
the prisoners, one must have inertial or stress effects to overcome
centrifugal effects due to the player interests  . To obtain fair
solutions that are based on self-interest and reflect the Nash
equilibrium, no such inertial effects are needed.
Prisoner’s Dilemma—Steady State Geometry 325

 From Sec. 8.4, the student should see that strains, which are the
consequences of the inertial stresses, determine the player bonding
 . In the steady state hypothesis, these strains are steady state and
are determined from the known conditions. Player bonding
components and player interests result from payoffs and equilibrium
strategies that are not Nash equilibriums. Such components generate
inertial attractive effects similar to gravity as well as repulsive effects
due to pressure gradients. Shear effects are present as well, which in
the symmetric prisoner’s dilemma are attributes of the initial payoffs.
 From Sec. 8.5, the student should see from the numerical results that a
given solution is not symmetric between the two players. For the case
we considered, prisoner 1 (   0 ) is more constrained than prisoner 2.
The player bonding gradient is larger in magnitude on the left as well.
 From Sec. 8.6, the student should see that the player bonding shear
effects are sensitive to the inertial stresses. For a perfect fluid, the
contributions are from the player interests.
 From Sec. 8.7, the student should see that the player interests can be
quite large. When they are large, they drive the player engagement
gradients, moving zero engagement to non-zero engagements, such as
in Fig. 8.13.
 From Sec. 8.8, the student will be able to trace the effects of the
stresses and strains in the theory to the persistency components
E j E j E j  . The student will understand the mechanism that
generates these important new effects in decision process theory that
payoff values must change as a function of strategy. The student will
also understand the mechanism for player engagement or coupling
that determines the size of game theory effects in the theory. A result
that has possibly wider applicability is that an aggressive player
displays a greedy charge whereas the other player has an
accommodating charge. From Sec. 8.8.6 the student should see that
the player bonding directly determines the inactive metric  jk in the
centrally co-moving frame, which is not changed if viewed in the
normal form coordinate frame. The student should see that the player
bonding potential determines the utility for player cooperation.
326 Geometry, Language and Strategy—Vol. 2

8.10 Exercises

(1) For numerical work, the computations go faster if we have


differential equations for E j  and E j  even if these are given in
terms of the inactive metric, E j and E j . Show that the
differential equations are:
 E j  q E j  2 E j  f j


 E j
  2 E j


  E j
  e f j
 . (8.14)
f j
 
 2 q E j       E j  e E j    e E j  1  E k
 Ek  

(2) Show that for the general case, the player bonding component  ,
Eq. (8.2), satisfies the following equation:

  p 
    q              ni p  ni 
 n 1  . (8.15)
 p       2    2         

(3) Show that for the general case, the bond shear components   ,
(8.2), satisfy the following equation:

    q             1


ni h     . (8.16)

(4) Show that for the streamline solutions of Chap. 6, the bond shear
and player bonding can be written in terms of potentials as follows:

       0,           0


      h     
. (8.17)
       1
ni  h
      q              1
ni h    

(5) Using the substitution defining the reduced bond shear potential
  below, show that the equation for the reduced bond shear is
given below for the streamline solutions of Chap. 6. In particular,
you see that when the reduced pressure is proportional to the
Prisoner’s Dilemma—Steady State Geometry 327

identity,     h for any scalar field  , then the source term on
the right is zero. What can you conclude about the reduced bond
shear? Show that this implies that the reduced bond shear matrix
  is constant for a single strategy model.

        q          1
ni h    
   e  q    . (8.18)
    e  q     1
ni h    
Chapter 9

Prisoner’s Dilemma: Determinism


and Chaos

Our study of the prisoner’s dilemma is part of an inquiry into decision


making and its connection to uncertainty. It can be thought of as a pure
form of the problem in which two players can either adhere or not adhere
to a contract, Cf. Sec. 11.2.2. As such, this example highlights aspects of
decision making, which is one of the most human acts and seems to be
the most difficult area to formalize into a theory of behaviors that are
causal and deterministic. In fact one might think that the very nature of
decision making is one of chance and uncertainty. Game theory focuses
on these uncertainty aspects of decision making and assumes that we
have access primarily to the equilibrium point that is reached based on
rules provided by the theory.
We take a quite different view, one that aligns with the view taken in
the physical sciences. In this view, there is uncertainty along the line of
the uncertainties associated with measuring distances and time; there are
also uncertainties in knowing the exact values of initial conditions.
However, once these uncertainties have been excluded from the
theoretical description, there remain decision events that evolve in time
and spread out in space according to causal and deterministic rules. We
have proposed these rules in decision process theory. Our investigation
of phenomena or case studies such as the prisoner’s dilemma is based on
these rules.
To explore the implications of the theory, we adopted the player fixed
frame model (Chap. 8) as an approximation to computing the steady state
geometry that one would expect given the steady state hypothesis. We
found that the steady state geometry was characterized by the player

329
330 Geometry, Language and Strategy—Vol. 2

engagements e , the player interests  , the inertial acceleration q


and the player bonding    b . In our analogy of waves on a lake,
these provided the perspective of the cork on the water. Specifically, we
examined these attributes in the co-moving orthonormal coordinate
basis, which provides insight into the field equations and facilitates the
numerical computations.
Our goal in this chapter is to look at the lake from the perspective of
someone standing on land. There are many possibilities that are
generated from a given steady state geometry, not all of which are
equivalent. The theory is non-linear, with different initial conditions
leading to different possible behaviors. Technically, this means that
different initial conditions lead to solutions that are not diffeomorphic.
Our approach will be to compute the frame transformations in the co-
moving orthonormal frame. We then consider a variety of distinct
solutions corresponding to different holonomic frames. All of these
frames qualify as possible normal form coordinate bases: they differ in
their specification of the initial conditions.
One of these is particularly useful: the centrally co-moving frame. In
this frame, the steady state geometry is transformed (See Sec. 6.10, Exs.
19ff.) into the player engagements V j , the player interests F j , the
inertial acceleration Q and the player bonding potentials  jk . We keep
the same names with the understanding that the appropriate
transformations will be made from the co-moving orthonormal frame to
the centrally co-moving frame. The transformations are based on the
coordinate transformations:

   eq
. (9.1)
 
We indicate the centrally co-moving frame by a “bar” over the
coordinate variable. This transformation leads to coordinates that are in
the stationary line gauge Eq. (6.98), which can be transformed
diffeomorphically into the harmonic gauge using Eq. (6.81), with
x a  ,   U 0  V a 0 and constant V a 0 . The equations determine U 0
with boundary conditions consistent with the centrally co-moving frame.
Prisoner’s Dilemma: Determinism and Chaos 331

In the centrally co-moving frame for the general case of more than
one active strategy, there will be two additional steady state geometry
tensors: the player payoffs F j  related to the orthonormal orientation
potentials   ; and the seasonal payoff f  related to the orthonormal
potentials   . These are all zero when there is only a single active
strategy.
There are three main categories of behaviors that we expect from our
shore-side perspective: travelling wave or harmonic behaviors as a
consequence of an initial harmonic force; resonant behaviors that are
harmonic and persist in the absence of an initial time dependent force;
and transient behaviors that die out in time, leaving one of the other
previous behaviors in place. Implicitly we considered harmonic
behaviors in Vol. 1, which are not expected here because there is a single
active strategy. We ignore for now the transient behaviors, so that in this
chapter we focus on the harmonic behaviors, which extend our notion of
steady state geometry1.
For a single active strategy, the behaviors are specified by Eq. (6.44)
in terms of harmonic polynomials, or equivalently by the central frame
wave equation that was derived in Sec. 6.8, Eq. (6.52) for any single
harmonic frequency  for the coordinate x a  x a ei :

  x a   2 e 2 q 1  e e  x a     q   x a  0 . (9.2)

Note that we need not assume that the harmonics are the same for each
coordinate, nor need we assume that there is a single harmonic: each
harmonic obeys an equation of the same form. For simplicity however,
we start with a single harmonic that is the same for all coordinates.
We see the steady state geometry contributions to the wave equation.
We see the effects of the player engagements, player bonding and
inertial acceleration in the resultant dynamic behaviors. In addition, we
should see that the initial conditions in effect generate a time dependent
source. It is for this reason that we argue that the resultant behaviors are

1
This is similar in concept to the way in which steady state AC behaviors extend the
concept of DC behaviors: the behaviors are independent of time, with the dynamics
captured by the specification of the driving frequency.
332 Geometry, Language and Strategy—Vol. 2

harmonic based on a time dependent driving force. Not only are the
flows dynamic in the sense that the steady state flows may exhibit
accelerations, they are dynamic in the sense that in the normal-form
coordinate basis, they explicitly depend on time.
As an introduction, we start with a general discussion of determinism
and chaos, Sec. 9.1. In Sec. 9.2 we focus on the flow of time as seen in
theory, which sets the stage for non-linear time dependences with
equations that appear linear. We provide evidence of a gravitational field
whose origin is from the inertial acceleration and from the player
engagements. We follow this with the behaviors of the strategy flows,
Sec. 9.3 showing that the harmonic time behavior generates a harmonic
spatial behavior. This provides evidence for behavior similar to other
physical systems. In Sec. 9.4, we compute the time and distance contours
from knowledge of the values of the active metric components g ab . To
explore possible sources of chaos and other interesting behaviors, we set
up a sensitivity analysis in Sec. 9.5 based on what the players consider to
be at stake in the decision process. We carry out this analysis in Sec. 9.6
for the baseline, Sec. 9.7 for low and high stakes and in Sec. 9.9 for the
normal form coordinate basis perspective. In Sec. 9.9, we examine
recurrence plots both in time and in strategy space in search of evidence
for chaotic behaviors in steady state phenomena.

9.1 Determinism and Chaos

To understand how chaos might arise in decision process theory,


consider the sensitivity of future behaviors on initial conditions in any
differential geometry theory. This has been extensively studied under the
general category of chaos and chaos theory. In pre-Newtonian physics,
chaos represents the way we perceive the world in its unordered state. In
Newtonian physics, this possibility is presumed to go away. If we had
perfect information, so the argument goes, we would have perfect
determinism.
A modern view of chaos is that even slight disturbances on what we
think we know, may lead to unknown consequences, even in a theory
that is strictly causal and deterministic. So “what is deterministic?” It
seems unreasonable to believe that because chaos behavior is possible,
we must throw out our causal theories. They work very well and explain
Prisoner’s Dilemma: Determinism and Chaos 333

a host of data. We believe that the reasonable approach is that we must


be careful about what we claim to learn from these causal theories.
The theories after all reflect our efforts to identify concepts and
attributes that don’t change with time from concepts and attributes that
do change with time in an understandable, causal and continuous way.
We are successful if we can make this separation. Physical theories start
with this separation and note that the measurement doesn’t easily fit
either category.
So for example we understand by length, an attribute that
characterizes the height, width or depth of an object. It is a great
accomplishment in understanding to separate this concept from the
mechanisms by which we perform the measurement. The mechanisms
involve a measuring stick and us as an agent; for those of you measuring
a basement, you know that multiple measurements yield multiple
answers. Yet we are confident that the basement has well-defined
dimensions. How did we come to this conclusion and how did we learn
to separate out the uncertainties? There are uncertainties associated with
us as agent using a measurement stick, which are distinct from the
invariant concept of length. Somehow we answered the question and
today, we all agree that the separation has been done and we are
comfortable with the idea of length.
Similarly, we are comfortable with the concept of time, despite our
dependence of using clocks to make time measurements. From such
simplistic considerations over many centuries, we have adopted physical
theories of the behavior of matter that we depend on. For example, we
are comfortable with a host of physics problems that relate distances
objects travel with time. We believe we understand how a pendulum
works because we can predict the behavior starting from a description in
which we describe its restoring force as being the source of the
acceleration. The behavior is the set of positions of the pendulum over
time. We start the pendulum at rest and “drive” it by a harmonic force.
We can also start the pendulum with an initial amplitude and velocity. In
either case, we predict from Newton’s theory where the pendulum will
be at any future moment. We compare where the pendulum is by
measurements against where it is predicted to be and find agreement to a
high degree of accuracy.
334 Geometry, Language and Strategy—Vol. 2

Because of the non-linear behavior of the forces, this model of the


pendulum though simple, generates unexpected structure from these two
quite different perspectives: behavior driven by an external force or
behavior driven by initial conditions. In the first case one can imagine
that for harmonic forces, a non-linear harmonic behavior settles in after
all transients have died out. In the second case, we imagine that the non-
linear behavior will be reflected as sensitivity of the long term behavior
to different initial conditions.
What do we expect for the non-linear harmonic behavior? Suppose
we “drive” the behavior by applying an external force characterized by
amplitude and frequency. As we vary the frequency and amplitude we
stress the non-linear structures of the problem. For sufficiently large
amplitudes we in fact generate chaotic structures: we may appear to jump
from one periodic structure to another. We create behaviors that appear
erratic and lack the periodic behaviors seen with smaller driver
amplitudes. In some sense, the behaviors can be viewed as both steady
state and chaotic. In studying steady state solutions we would expect that
qualitatively different solutions might result from small changes in the
initial boundary conditions.
Alternatively, suppose we start a system from a given set of initial
conditions. Chaotic behaviors are those in which subsequent behaviors
appear quasi-periodic. There may be limit cycles and actual jumps from
one cycle to another.
In engineering and business, there are also distinct ways to gain
access to a system’s non-linear characteristics. The idea is that these
properties may in fact carry over into the realm of decision making. We
expect that decision making has attributes that involve imperfect
information as well as perfect information. The challenge is to identify
each of these, separating out those attributes that have a predictable
behavior from those attributes that are inherently uncertain.
We adopted game theory [Von Neumann & Morgenstern, 1944] in
which an intrinsic view of decisions is a productive starting point where
we separate out the pure strategies as things that have permanence. A
pure strategy is the complete set of moves one would carry out in a
decision process taking into account the moves of all of the other players
or agents in the process along with any physical or chance effects that
Prisoner’s Dilemma: Determinism and Chaos 335

might occur. It is a complete accounting of what you would do, a


complete plan given every conceivable condition. It is furthermore
assumed that you can approximate this complete list with a relatively
small list of pure strategies.
Just because there are pure strategies, there is no reason to believe
that one of these pure strategies is the right choice to make. If you are in
a competitive situation, there may be a downside to your competitor
knowing that you will pick one of these strategies. The solution is to
“hide” your choice by picking the pure strategies with a specific
frequency. The theory determines for you what these frequencies are
without informing your opponent which choice you actually will make
on any given play.
Thus your decision choice is a specific frequency choice and in that
sense represents the measurement of “length” despite the fact that in a
real decision process, like a real measurement process, there are lots of
uncertainties. You would like to determine the frequency choices the
other players make and they want to understand your choices. We
emphasize that knowing these frequency choices is not the same thing as
knowing what you will actually do on a given play. We take this
knowledge in the same way we take the knowledge about the size of our
basement. We know how to get a good approximate set of
measurements. We know that out basement has a size. For each
measurement process we don’t know what size we will actually get.
In Vol. 1, we extended game theory to decision process theory,
which causally computes future frequency values based on given initial
conditions. The theory has similarities to physics and other mathematical
models of physical processes. Just as in physics, there can be external
forces that dictate how the rates of change of frequencies change in time.
There will be stationary situations in which these rates of change don’t
change, in which the forces generating such changes are zero. We equate
that scenario with the whole literature of game theory and its
consideration of static games: the frequencies are fixed numbers. Static
games provide an important foundation for our approach, though our
results diverge once dynamic effects are included.
A second scenario is one in which the fields that generate the forces
are static, but the flows, the rates of change of the frequencies, are
336 Geometry, Language and Strategy—Vol. 2

dynamic. This is the point of our analogy about observing travelling


waves on a lake: from the perspective of the cork, the forces are static;
from the perspective of an observer on the shore, the flows are dynamic.
We gain insight into these dynamic behaviors by looking at the various
steady-state harmonic solutions. We also gain insight by considering
initial conditions that stress the system, involving lots of harmonics.
Ordinarily, we are interested in solutions that are not sensitive to initial
conditions.
However, just because we have a theory that determines future
behavior based on knowledge of past behavior does not mean that we are
justified in assuming that the predications always will be insensitive to
our starting point. Chaotic behavior results when small changes in initial
conditions generate large changes over time or space, even if that
behavior ultimately stays bounded. Globally we may see significant
deviations. We may see such changes either by studying the sensitivity to
initial time conditions or by studying the sensitivity of structure of steady
state solutions to boundary conditions.
We expect that chaotic behaviors can be generated from within
decision process theory, if there are suitable parameters that can be
varied and suitable strengths chosen. We can either study the sensitivity
to initial conditions of processes as they evolve in time, or we can study
the harmonic behaviors and their sensitivity to the boundary conditions.
If we study harmonic behaviors, it is a matter of finding the
amplitudes and frequencies that excite the underlying structures:
seemingly benign behaviors as a steady state need not indicate the lack of
interesting structures. The key is to find when such structures are
sensitive to boundary or initial conditions. Our analysis in the following
sections is an attempt to find evidence for such amplitudes and
frequencies.

9.2 Time Bonding

In this section we explore the dynamic behaviors that result from


decision process theory with numerical examples for the prisoner’s
dilemma assuming the same nominal steady state geometry used in
Prisoner’s Dilemma: Determinism and Chaos 337

Chap. 8, which presumably takes us far from any chaotic behavior. In


general we will go from the co-moving orthonormal frame to the normal
form coordinate basis, though as discussed earlier, some questions are
best dealt with in the centrally co-moving frame.
To carry out our calculations, we have to expand what we mean by
known conditions, Sec. 7.6. We could specify the flows and payoffs for
all strategic values at a proper-time of zero. However, for harmonic
steady state solutions, we borrow from engineering. We create solutions
that provide insight given our expectations that such normal
specifications could be built up from harmonics starting with steady state
solutions, solutions linear in time and then solutions built from
harmonics or phasors. In mathematics, such solutions provide a Fourier
analysis from which to build any solution, including the one of interest
where we know the initial conditions. For the numerical calculations in
this chapter we have used the harmonic polynomials Eq. (6.36) for the
streamline solutions of Sec. 6.8. We only have to specify which
harmonic polynomials we will focus on. We can easily transform these
solutions to the centrally co-moving frame as needed.
Unless stated otherwise, the illustrative results in this section are
based on proper-times  in the interval  0,1 and consider solutions that
are linear on which there is a single harmonic frequency   10 . If
harmonic polynomials are used, we take the degree N  50 . We are not
looking for an engineering solution to a particular problem. Rather we
hope to survey the types of behaviors expected. We are free therefore to
pick nominal values for the harmonic polynomials:

0.05
t  t  ,   a t    bt  sin N 

 t  ,    a t  
. (9.3)
0.05
u  u  ,   a    b  
u u
sin N 

 u  ,    a u  

These equations determine the behaviors away from the origin   0 of


zero proper relative player effort. The linear terms are determined by
338 Geometry, Language and Strategy—Vol. 2

ordinary differential equations whose initial conditions are set by the


flow boundary conditions and the Gram-Schmidt orthogonalization
process described in Sec. 7.6. Obviously many other choices other than
Eq. (9.3) are possible, some of which will be discussed in Sec. 9.5 where
we do a sensitivity analysis of our results.
Our first set of challenges is to define time and distance in the active
geometry. In a gauge theory, the measure of time and distance intervals
depends on the metric potentials, Eq. (2.46). Neither time nor distances
are absolute quantities; rather we have insight into them from the metric
potential fields. We may think of these metric fields as the dynamic
strategy bonding fields. The metric potentials are determined by the flow
and frame transformations, so we start with their behaviors based on the
streamline solutions to Table 5.1 provided in Sec. 6.8 summarized in
Eq. (9.2).
We have seen that inertial acceleration and player bonding play a
large role in the steady state geometry and in maintaining structural
stability. Furthermore, we have seen that the notion of distance in
decision processes, which is intimately tied with the concept of
measuring utility, is not an absolute concept, but a relative concept that
depends on the frame of reference. In rotating frames2, where the rate of
strategic flow changes, distances can appear shorter as seen by an
observer who is not rotating. The concept that distances are relative is a
consequence of the covariant nature of decision process theory. This
disconnects the underlying topological space from the quantitative
measures of nearness. This connection is broken for our measure of time
differences as well.
We have two results in common with the theory of relativity in
physics: our conception of time differences depends on frame; and the
metric potentials with which we analyze distances can form steady state
waves. In both cases we rely on the relationship between the active
metric and the frame transformations, Eq. (4.72):

2
It may sound strange to talk of rotations. Yet payoffs, because they are anti-symmetric
tensors, specify and generate rotations. Rotations thus play an important role.
Prisoner’s Dilemma: Determinism and Chaos 339

g ab  1  e e  E a E b  E a E b
. (9.4)
gab  1  e e  Ea Eb  Ea Eb

In the centrally co-moving frame, the metric is Minkowski, with the


exception of the time component Eq. (6.73):

g   1  e e  e2 q
g   0 . (9.5)

g  1

In the centrally co-moving frame, the time contribution to the invariant


distance is provided by the covariant metric:

e 2 q
g  . (9.6)
1  e e

This is determined from the player engagements, Fig. 8.15 and the
inertial acceleration Fig. 8.3, both of which contribute potential wells
centered at zero. Thus the time bonding has a minimum where the inertia
is greatest and where the acceleration is zero, Fig. 9.1. This is where we
expect time to move more slowly.
We expect that in areas with
large inertia, any quantitative
measure of time slows down. Not all
parts of the topology can be reached
from a given starting point:
boundaries will be set by areas
where the components of the metric
field vanish. We can show that
steady state wave solutions exist,
Figure 9.1. Time Bonding g (PDM)
exhibiting curvatures determined by
these boundaries. See the exercises in Sec. 9.11.
340 Geometry, Language and Strategy—Vol. 2

9.3 Strategy Flows E a

As a recap, a symmetric prisoner’s dilemma is formulated as a player


fixed frame model using the steady state hypothesis. It is a single
strategy model, whose streamline solutions and required scalar equations
are detailed in Sec. 6.8. We have computed the inertial behaviors Table
5.6 in Sec. 5.4.4, the electromagnetic and tidal behaviors Tables 5.4 and
5.5 that generate strain in Sec. 8.4 and the persistent behaviors of Table
5.1 in Sec. 8.8. In this section we use Table 5.1 to provide the dynamic
behaviors of the flows E a and the transformations E a to the proper
active direction that depend on the proper-time. The partial differential
equations that result are Eq. (6.51), whose solutions in terms of harmonic
polynomials can be obtained from Eq. (6.96). These equations are
effective tools to provide the time dependence of the theory. We are not
aware that these techniques have been used elsewhere.

Figure 9.2. Time flow E t  ,  (PDM)

We start with the time flow frame transformation E t  e  q E t as a


function of both strategic direction  and proper-time  . The
transformation has two factors: a transformation from the co-moving
orthonormal frame to the central frame, and a transformation from the
central frame to the normal frame. The transformation to the centrally
co-moving frame depends only on the inertial acceleration Fig. 8.2,
which is independent of proper time and has a maximum at   0 . The
Prisoner’s Dilemma: Determinism and Chaos 341

central frame transformation to the normal form coordinate basis,


Fig. 9.2 indicates travelling waves. From the harmonic disturbance
generated in time, harmonic disturbances are generated in space going in
opposite directions. To get the complete picture of the proper-time frame
transformation, we multiply the two factors, obtaining Fig. 9.3. In this
case we see that the acceleration effects dominate.

Figure 9.3. Time flow E t  ,  (PDM)

Next we look at the active strategy frame transformation, the active


flow, E u  e  q E u , which again is the product of two factors. The
harmonic characteristics are captured by the frame transformation from
the centrally co-moving frame to the normal form coordinate basis, E u .
Though for a linear solution the flow would remain zero, the harmonic
polynomial contribution provides travelling waves, Fig. 9.4. We see that
the flow indeed starts at zero. The initial behavior set by the harmonic
cos N  for the distance behaves like sin N  for the flow and thus
starts at zero. Along the streamline from the origin, this harmonic
behavior generates spatial oscillations that are clearly displayed based on
our initial conditions.
342 Geometry, Language and Strategy—Vol. 2

The travelling wave behaviors are also visible in the contour plot for
the frame transformation E u , Fig. 9.5. Here we see the distortions that
reflect the difference between our wave equation, Eq. (9.2) and the usual
wave equation. We see that the null lines between the peaks and troughs
are not at right angles. Furthermore they bend at the boundaries at the
left, determined by the vanishing of the inactive metric components. The
direction of the travelling wave also bends due to the same boundary.
If we think of the peaks as being places where inertia collects, then the
motion of the inertial peak is the movement of energy density. This
motion should be less than the maximum speed possible and should form
an angle of greater than 45 degrees.

Figure 9.4. Active strategy flow Eu (PDM)


Prisoner’s Dilemma: Determinism and Chaos 343

Figure 9.5. Active strategy contours E u (PDM)

9.4 Time, Distance and g ab

We have determined the time bonding gtt , which is formed from the
inverse of the metric components g ab . The result, Fig. 9.6, shows the
inertial acceleration and player engagement effects, which are both
independent of central time. On top of these two static effects is a ripple
in time due to the single harmonic. Because of the ripple, there will be
local minima in which inertia can be trapped. Except for the fact that we
might have too few dimensions (Cf. Sec. 6.7), we interpret the effect as a
gravity wave, since it is a wave in the metric. A more accurate
interpretation might be that it is a reflection of a frame-wave, Sec. 6.6.
We suggest that the gravity-source has the travelling wave behavior that
is reflected in the metric.
344 Geometry, Language and Strategy—Vol. 2

We see structure
in addition to the big
picture behavior of
the trapping field
formed from the
inertial acceleration
and player
engagements. Not
only is our notion of
time influenced by
the frame, so is our
notion of distance
along the strategic
Figure 9.6. Time bonding gtt  ,  (PDM)
direction, Fig. 9.7,
where the main components come from the transformation E u and a
somewhat smaller contribution from E t .
As with the time flow, there is a static behavior on which there is a
small ripple effect due to the harmonic contribution. It is significant that
the wave properties one sees in physics are reflected here as well:
harmonic behavior in time leads to harmonic behavior in strategy space;
harmonic behavior in space leads to harmonic behavior in time. We
conclude that a localized event at an initial point in time and space will
radiate outwards in both time and space.

Figure 9.7. Distance-time central Figure 9.8. Contours for central


contours (PDM) frame coordinates (PDM)
Prisoner’s Dilemma: Determinism and Chaos 345

In Fig. 9.7, the horizontal lines represent constant values of central


time  , while the vertical lines represent streamlines that are constant
strategic values of central relative player effort  . With our choice of
initial values, time in the normal form coordinate basis is approximately
the same as central time. Note the bunching of the streamlines, especially
on the right, where prisoner 1 is more active.
We study these behaviors relative to the co-moving orthonormal
frame. In Fig. 9.8 we plot the central player effort  versus central time
  e q  with contours of proper player effort  and proper time  . If
we think of the curves of constant proper time as representing events
with the same “age”, we conclude that aging appears differently in the
normal form coordinate basis depending on the initial proper (central)
relative player effort, which follows from the dependence of the central
time gradient  on the inertial acceleration, Eq. (6.57), Ex. 5 Sec.
6.10. Roughly we can say that a decision process that occurs in an area
of high energy density takes longer than the same process in an area
of low energy density. We see that the effects are due to the inertial
acceleration.
The harmonic
effects for g uu  ,  ,
Fig. 9.9, suggest that
non-linear effects in
the field equations
are important. We
see a wave that
propagates to the
right, developing a
large hole or
indentation. To
understand the
Figure 9.9. Strategy bonding guu  ,  (PDM) implications of this
we analyze two
general characteristics that we glean from these calculations.
First, there is a causality principle: signals between two points can’t
go faster than a null geodesic (Ex. 1): g tt  0 . For equally spaced
surfaces of proper-time, if time intervals in the normal form basis
346 Geometry, Language and Strategy—Vol. 2

become compressed, then the corresponding size of gtt must grow to


compensate and so we expect gtt   and hence g tt  0 . This
suggests we go a fixed distance in ever decreasing amounts of time and
so suggests the velocity of the null-geodesic becomes infinite. The
causality principle is that this limit is not reached and the time frame
transformation starts positive so we maintain E t  0 and g tt  0 .
More generally, causality for decision processes is the common sense
view that future events can’t be influenced instantaneously by past
events: sufficient time, no matter how short, must pass for a future event
to become aware of the past action. In decision process theory, the
mechanism for communication stems from the fact that there is finite
maximum speed for signals to propagate set by the null-geodesic. Such
propagation effects will be observed both in the metric potential fields as
well as the payoff fields. They are analogous to gravitational waves and
electromagnetic waves respectively.
Secondly, there is the seemingly obvious principle of possible
change that at any point in space time, change is always possible. In
other words, active strategies can change in time. What makes this not so
obvious is that strategies can’t change faster in time than a signal along
the null geodesic. If the velocity along that null geodesic goes to zero,
change becomes impossible. This occurs whenever streamlines begin to
bunch together. We see evidence of this in Fig. 9.7.
The speed of propagation follows from assuming for the active
components:

d 2  g ab dx a dx b  0
g tt  2 gtu c  g uu c 2  0 . (9.7)
2
g  g  g
c  tu   tu   tt
 g uu   g uu   g uu
Prisoner’s Dilemma: Determinism and Chaos 347

We expect signals to propagate in the forward light cone formed by these


two null-geodesics. The effects are particularly striking when we do the
sensitivity analysis Fig. 9.10 for the high stakes low  model, Sec. 9.7.
We see that the streamlines bunch together quite dramatically. If the
strategy bonding  guu   , then the two velocities go to zero and the
possible streamlines become trapped or bunched. We do see evidence for
this bunching. As we
get more bunching
(space shrinks), we
require that guu
becomes large.
Therefore we must
have g uu  0 as a
requirement to
enforce the principle
of possible change
(Ex. 2).
With these
Figure 9.10. Null-geodesics constraints for high example behaviors
stakes low  using just a single
harmonic contribution, we see dramatic changes in the character of the
steady state solutions. We go from a solution with a single potential well
to solutions with multiple wells. We might think of this as analogous to
galaxy formation. These additional pockets of structural stability are
consequence of a single frequency dominating. We have not really
proved this. However, it is not difficult to argue that a pulse of events
localized in time might be similar to a pulse located initially in space.
Such a pulse would correspond to a number of frequencies around some
common average frequency. In other words, we would argue that a single
frequency might occur corresponding to a localized energy density
distribution. Our figures are not too dissimilar to that set of assumptions.
348 Geometry, Language and Strategy—Vol. 2

9.5 Sensitivity Analysis for Dynamic Behaviors

We have provided an engineering view of the large scale structure of the


solutions, not unlike the view taken by Hawking & Ellis [1973] towards
Einstein’s general theory of relativity. The insights are quite different
from those obtained using weak field approximations to these theories.
Therefore, despite the simplicity of the model for the prisoner’s dilemma
and the relative simplicity of the common good solution, we obtain non-
simple structures in variations around this base.
To investigate this further we again do a sensitivity analysis. As part
of our sensitivity analysis of steady state solutions in Sec. 8.3, we
introduced an alternative Nash solution, which illustrated additional
characteristics. The major difference between the common good and
Nash solutions was set by the initial strategies. We discussed the effect
of setting the expected game value to zero in each case.
In this section we investigate the sensitivity based on time dependent
dynamic behaviors. To articulate the major possibilities, it will be
sufficient to consider primarily the Nash solutions with a small initial
pressure p  0.01 and zero game value. By using a small initial pressure
and zero game value, we suppress the effects we have already studied
that generate player interests. We use the harmonic parameter values
used in Sec. 8.3. They provide a baseline that facilitates our
understanding of the effects.

Table 9.1. Player Stakes  and Decision  Choices for Sensitivity Analysis

Low  Baseline  High 


Low stakes   110   110   110   1 3   110   9 10
Baseline stakes   1   110   1   1 3   1   9 10
High stakes   10   110   10   1 3   10   9 10

We consider different values for the game beta  and (equal) player
stakes  1   2   , which we choose from Table 9.1. These choices
have a major impact on the time component gtt of the metric that
determines the behavior of time intervals, Eq. (9.6). In a qualitative way
Prisoner’s Dilemma: Determinism and Chaos 349

we anticipate the effects from the numerical calculations by taking the


first term of Eq. (9.4):

g tt  1  e e  E 
1 t 2
 . (9.8)

The first factor is determined by the player engagements, Table 5.1. We


recall from our previous numerical results, Fig. 8.15 that the initial
player engagements are set by the initial flow; for those values of flow
that are zero, the corresponding player engagements are zero. Thus we
expect this factor to be governed by  .
The second factor initially depends on the inertial acceleration. In
particular the payoffs determine the (reduced) shear, Eq. (8.7). If we
increase or decrease the payoffs by an order of magnitude, we expect
these shear values to vary proportionately.

9.6 Baseline

So our expectation is that  , an attribute of the inertial field, governs the


first factor and the player stakes  , an attribute of the payoff field,
governs the second factor. The second factor gives rise to the harmonic
behaviors. We examine the baseline form for the time bonding contour
g tt  ,  , Fig. 9.12, for the Nash solution with small pressure p  0.01
and zero game value.
The baseline contour plot for the
flow E t , Fig. 9.11, displays a distinct
lattice structure without the focus
contribution from the player
engagements. The lattice structure is by
no means an obvious consequence of the
partial differential equations Eq. (6.44).
The result is a consequence of the
acceleration being small so that the terms
in the partial differential equation that
depend explicitly on proper-time can be
ignored. This is not exact nor is the
Figure 9.11. Baseline contours
lattice dependence uniform.
E   , 
t
350 Geometry, Language and Strategy—Vol. 2

Figure 9.12. Baseline contours gtt  , 

Figure 9.13. Low stakes and low  contour plot for the
metric gtt  , 
Prisoner’s Dilemma: Determinism and Chaos 351

The choice region for the decision process, the allowed space of
strategy values  , is bounded below at approximately   0.48... .
Much below this the equations become singular. We can go much higher
that   1 , though it requires progressively more harmonic polynomial
terms to get reliable results, or we must use the alternate approach with
Eq. (9.2). We obtain a lattice picture that is truncated by the choice
region. As in our previous dynamic calculations, we see the effects of the
wave equation.

9.7 Low Stakes, High Stakes

To widen the choice region, we go to the case with low stakes and low
 , where we expect little contributions from the strain, little
contributions from the player engagements (based on small  ) and large
contributions from the harmonic contribution. We get the expected result
that the player interests are almost zero, the player engagements are
small and the first factor in Eq. (9.8) is almost unity. The potential well
comes almost entirely from the second factor E t  ,  , the harmonic
contribution; note that there is a negligible contribution from the initial
shape of E t  ,0  . This is clearly displayed in the contour plot, Fig.
9.13. The choice region for the decision
process is expanded from the baseline
case.
The contour plot shows peaks (and
valleys) that move along lines of   1 ,
such as the two that start at the origin.
The light color indicates peaks and the
dark color indicates valleys. For this case
we replicate the harmonic wave equation
and phasor solutions. The underlying
lattice structure is seen in the flow E t ,
Figure 9.14. Low stakes and low Fig. 9.14. We see a hexagonal packing
 contour plot for the flow pattern. The light color areas indicate the
E t   ,  path taken along the minimum of the
potential.
352 Geometry, Language and Strategy—Vol. 2

An alternative but
not equivalent way to
narrow the choice
region is to consider
high stakes and low
 . In this case we
modify the shear
components, which
influence the time
structure, the lattice
pattern and narrow
the choice region,
Fig. 9.15. As we
Figure 9.15. High stakes and low  time bonding move away from the
gtt  , 
origin along the
strategic axis, there
are progressively deeper potential well pockets, indicating a degree of
instability. The energy generating these pockets comes from the payoffs
and the assumption of high stakes. The higher stakes along with
harmonics generates the effect. If we remove the harmonic contribution
we don’t expect the same structure. The striking feature of the contour
plot is the clover leaf pattern, an unexpected consequence of the partial
differential equation. This clover leaf pattern is not as clear in the
contour for E t .
The final case we consider is high stakes and high  , which has
solutions for a slightly narrower range, Fig. 9.16. For high stakes and
high  , the choice region is a little smaller with the same structure as
high stakes and low  , Fig. 9.15.
These are significant consequences and they are based on the size of
the payoffs. They are in addition to the effects based on the relative
values and the benefit to the players on the expected game values. In
addition to the inertia effects associated with the energy density and
pressure, there are consequences based on the  characteristic or speed
of the flow, which is a type of inertia.
One reflects a tendency each player may have to favor certain
strategies. This we associate with the energy density and pressure. A
Prisoner’s Dilemma: Determinism and Chaos 353

high inertia suggests that the players are reluctant to change their style of
play. For example, we might consider a version of chess in which there is
a timer. This changes the speed at which the game is played. A very high
speed version of chess would be a high stakes game, which introduces a
very different aspect of the inertial fields.
We speculate that
the latter definition of
inertia is significant in
physical theories of
cosmology. Early eras
of the universe might
be characterized
primarily as radiation,
so that this would
correspond to a high 
era. Things don’t clump
together. In contrast,
the current era would
be a low  era.
Galaxies have formed;
Figure 9.16. High stakes and high  time bonding there is a lot of
gtt  ,  clumping. These
characteristics hold in
decision processes. Thus we believe the effects we have identified in this
section are essential. Both the speed  and stakes  play a role that has
not been accounted for in the standard game theory approaches. They
determine the choice region and structure of the decision process.
In this and the last section, we have provided the time dependent
dynamic behaviors we expect for the prisoner’s dilemma in the co-
moving coordinate basis. In the next section we return to the normal-
form coordinate basis and examine the dynamic behaviors we expect for
the payoffs.
354 Geometry, Language and Strategy—Vol. 2

9.8 Normal Form Behaviors

There are many ways of investigating the dynamic behaviors in decision


process theory. If we are given the behaviors of the system at an initial
point in time for all possible active strategy values, we can then compute
the behavior of the system as a function of proper-time. We equally well
gain insight into the system by considering the behavior of the system as
known at a single strategic point and all times, which for us is the
presumed equilibrium point   0 . The behavior of the system is
predicted by the equations for all other strategic values  . The latter
approach leads to the former approach by superposing the harmonic
solutions to achieve any given strategic behavior  of the system at
initial proper-time   0 . This approach corresponds in concept to the
phasor analysis of circuits in electrical engineering. Examining the
response of a stable system to a known perturbation is also common in
system dynamics, for example, [Senge, 1990].

Figure 9.17. Baseline normal form basis flows

Recall from Sec. 7.6 that initially, the flow represents the strategy
choice of each prisoner, Fig. 9.17. One possibility is that this flow is
constant. We have explored the dynamic possibilities in our baseline
Nash solutions by superposing a harmonic onto this constant behavior.
We see the assumption that each player starts with the strategy to
confess. The time strategy reflects the baseline value for   1 3 . We see
the code of conduct imposed that the flow to not confess is zero and so
these flows oscillate around zero.
The harmonic disturbance propagates outward in space, Fig. 9.18, in
all appearance like an electromagnetic wave, as a reflection of the frame-
Prisoner’s Dilemma: Determinism and Chaos 355

wave behavior. In this case we see that the choice region narrows
causing the effects of
the harmonics to
damp out.
Furthermore we see
that at both extremes,
the prisoner chooses
some amount of the
confess strategy. We
expect to see these
effects in the payoffs
and other strain
parameters. It is
Figure 9.18. Baseline confess choice for prisoner 1 insightful if we
consider the transformed version of the space, Eq. (7.17), with Eq. (7.18)
for the behaviors that are internal to the decision process theory.
In this basis we have the player interest contributions F j tu , which
can be determined from the co-moving frame field f j , Table 5.1,
Eq. (4.30) using the frame transformations:
j
f
F j tu  
. (9.9)
E t E u  E u E t

As expected, the steady state behavior of the numerator for the prisoner’s
dilemma model parameters, Fig. 8.14, is modified by the time

Figure 9.19. Baseline player interests in normal form basis


356 Geometry, Language and Strategy—Vol. 2

dependence of the denominator. The behavior has by assumption the


single harmonic time
dependence at   0 ,
Fig. 9.19. We see the
harmonic wave in the
figure for the player
interests. Notice that
the player interest
field exhibits a
travelling wave
behavior. We explore
such behaviors
Figure 9.20. Baseline player interest for s1 code of conduct further in the next
chapter.
The rather small harmonic in s1 for example is seen to be damped out
quickly. Most of the behavior of this player interest component is set by
the restriction in size of the choice region. As we saw in Sec. 9.5 with the
sensitivity analysis, higher stakes and a lower  along with a lower
initial pressure at the origin all contribute to making the results more
sensitive to the harmonic.
If we use the
same analysis as Sec.
9.5, Table 9.1, for
high stakes and low
beta, we get a vastly
different structure for
the player interest
field component for
s1 , Fig. 9.21. In both
this and the previous
figure we start
with the same
Figure 9.21. High stakes-low  player interest for s1
presumption of the
code of conduct
time dependence as
being steady state with a small harmonic. With high stakes and low  ,
the system resonates or “rings” more dramatically and over longer
Prisoner’s Dilemma: Determinism and Chaos 357

(proper) distances than in the former case. However, the physical size of
the system is smaller, Fig. 9.22. In some sense the system is more brittle
when the stakes are high.
Since the player interest components reflect and replace the concept
of game value or expected payoffs, the brittleness reflects the risk
inherent in such high stakes decision processes.

Figure 9.22. High stakes-low 


distance time contours

9.9 Recurrence Plots and Chaos

Our focus in this chapter up until now has been on steady state
harmonics with a single harmonic. This focus is in keeping with our
study of steady state behaviors and the implications of the steady state
geometry to standing wave behaviors. Of course we have anticipated that
there will be non-linear effects and that they would manifest themselves
as changes of the magnitudes and phases of the flow as functions of
position and time. We saw evidence of this in Sec. 9.6 with the time
bonding contours, Fig. 9.12.
In this section we return to the concepts of determinism and chaos,
Sec. 9.1 and look for evidence in our steady state behaviors. Chaos in
deterministic systems is usually studied by starting a system in a known
position and following the causal behavior that results. For decision
358 Geometry, Language and Strategy—Vol. 2

process theory with a single strategy, that means using Eq. (9.2) and
imposing boundary conditions at an initial time   t  0 , specifying
such things as the energy flow for all values of the active strategy   u .
We use the differential equation to generate positions and flows at all
later values of time and at all strategy values. A simple explanation of
chaotic behavior is that small changes in initial values lead to
qualitatively different behaviors later. This sensitivity to the initial
conditions, suggestive of chaotic behavior, results from specific non-
linear behaviors of the physical processes.
The equations are complex and we have not completed an analysis of
the partial differential equations viewed as the initial value problem
suggested above. We suggest that we may indeed have such specific non-
linear behaviors and that they may be investigated in an alternative
manner. We can focus on a single harmonic and look for sensitivity of
structures in time as we change positions in space.
To identify structures and potential chaotic behaviors in both theory
and experiment, we use recurrence plots. A number of authors see
encouraging results using such plots, such as the study of chaotic
behaviors for heart beats by Sabelli [2005] and chaotic effects in
quantum wave functions and the early universe by Thomas et al. [2010].
The idea is that for any quantity of interest that depends on time, f  t  ,
you compare the quantity at different times and look at the two-
dimensional plot subject to the condition: f  t   f  t    . For
stochastic processes, there should be no pattern at all, since the
assumption is that future events are unaffected by past events. For
periodic behaviors, the two-dimensional plot will also be periodic. For
chaotic structures however, in contrast to stochastic processes, the
two-dimensional plot is expected to have non-repeating structures. For
the strategy and time flows we create such plots, Figs. 9.23 and 9.24.
Prisoner’s Dilemma: Determinism and Chaos 359

Figure 9.23. Strategy flow recurrence plot two-frequencies PDM

Figure 9.24. Time flow recurrence plot two frequencies PDM

Note that the same concept applies to distances: we claim that


decisions are connected not only in time but in strategic space. We
expect that neighbors in strategic space are connected. Therefore in the
centrally co-moving frame, there should be non-trivial recurrence plot
360 Geometry, Language and Strategy—Vol. 2

structures between streamlines f    f      at a fixed central time


   0 . In Ex. 14, Fig. 9.37, we identify such structures, which reflect
the non-linear character of the equations.
Thus if we start with our single harmonic steady state solutions,
recurrence plots of the energy flows E u E t  are expected to be
periodic. We find that to be the case, Ex. 13. Now consider two
harmonics  t  10 15 for normal-form time and two harmonics
 u  7 10.5 for normal-form strategic distance. We get a variety of
shapes, examples of which are the recurrences Figs. 9.23 and 9.24.
We see the start of non-periodic structures, which suggests we
generate the type of chaos envisioned by these authors in the limit of
adding up the Fourier series representing the initial conditions. The
shapes change as we change the strategic distance  0 . One still needs to
verify that this also corresponds to other definitions of chaotic behaviors.
We believe this is an interesting project to take on. We have done some
preliminary work on this already, Ex. 15.

9.10 Outcomes

The prisoner’s dilemma model illustrates that dynamic behaviors reflect


a topological ordering of time and strategic-space events that is common
to game theory formulations as well as a measurement or gauge
structure within this topology that provides a fabric on which decision
events occur. The fabric construct, or connection, has both energy and
momentum that influences, and can be influenced by, decision processes.
The student will see from the numerical calculations, analytic evidence
for inertial mechanisms associated with the stresses on the fabric and the
speed  of the energy flow. Non game-theoretic results are obtained
such as evidence for effects due to the player stakes of the process as
well as to the expected player interest or game value.
Because systems that balance forces share energy from one part of the
system to another, the student is led to the view that energy is the key
transferable quantity that replaces the classical notion of game value (see
the principle of opportunity, Sec. 13.6). All physical systems can
transfer energy; the student will note that one form of that energy is the
Prisoner’s Dilemma: Determinism and Chaos 361

subject’s potential energy per unit charge and that in the prisoner’s
dilemma model; it is closely related to the prisoner’s expected game
value payoff. These inertial and payoff mechanisms are new and specific
to the gauge theory framework adopted. We see a new mechanism, for
example, that suggests that when game values are large, the value itself
influences the game behavior.
By studying the numerical models, the student learns that the results
are based on a causal evolution of behavior starting from an initial
surface at zero proper-time or equivalently, an initial surface at zero
strategic distance (from a presumed equilibrium for example). This
demonstrates that payoffs are functions of strategies that are in general
different from that initial surface. The allowed ranges are determined by
the causality principle and the principle of possible change.
The student should appreciate that with the steady state hypothesis of
Chap. 6, all strategies can be active. The methods for solving these
equations may require different numerical techniques [Courant &
Hilbert, 1962]. Mechanical and electrical engineering disciplines solve
problems of a similar nature [Bhatti, 2005] and their techniques may be
applied to the partial differential equations Eq. (6.33). Multiple active
strategies introduce new attributes a from the commutation rules Eq.
(6.14). Nevertheless, the same techniques of analyzing the solutions in
terms of harmonic polynomials, Eq. (6.36) is possible using Mathematica
and the numerical method of lines [Wolfram, 1992].
The attainment of the outcomes of this chapter is facilitated by doing
the exercises in the following section. Based on this investment, the
student should achieve the more detailed outcomes below based on
section.
 From the discussion in Sec. 9.1, the student will have seen that for
small amplitude behaviors, systems are often linear. This means that
the amplitude of the oscillation plays no role in the outcome.
However, for non-linear systems and sufficiently large amplitudes,
the amplitudes do play a role, leading us to speculate that
deterministic behaviors have chaotic consequences. The student will
have seen some evidence of this in the numerical examples.
 The student will have seen the effects of time bonding in Sec. 9.2.
This concept follows from differential geometry; it identifies the time
362 Geometry, Language and Strategy—Vol. 2

component of the metric as leading to gradient forces analogous to


gravity in physical systems.
 In Sec. 9.3 we consider the strategy bonding forces that influence
strategic flow.
 From Sec. 9.4, the student learns from a simple example that
harmonics, generate substantially new behaviors that require two
principles: causality and change. The behaviors also demonstrate that
in decision process theory, the measurement of time is relative to the
frame of reference as is the measurement of space. Dynamic behavior
exhibits two related principles: the principle of causality and the
principle of possible change. Both refer to the characteristic of
decision process theory that the mechanism of communication is by
sending signals that must travel at a non-zero but finite speed, which
is set by the transmission speed of the wave equations.
 In Sec. 9.5, from a sensitivity analysis the student learns that there are
significant dynamic effects related to the stakes associated with a
decision. Qualitative new behaviors are exhibited when the stakes are
sufficiently high. This is in marked contrast to game theory where the
stakes have no strategic effect. The player stakes are amplitudes that
begin to play a role when their values become sufficiently large.
 In Secs. 9.6 and 9.7 we consider baseline solutions as well as low and
high player stake examples.
 In Sec. 9.8 the student learns the behaviors expected for the prisoner’s
dilemma in the original normal-form coordinate basis. The code of
conduct imposed on the solution is now clearly visible. The student
learns that the concept of game value is effectively replaced in
decision process theory by player interest.
 Finally, in Sec. 9.9 we return to the question of chaos and
determinism by considering field solutions that have more than a
single frequency.

9.11 Exercises

(1) The fastest signal between two points is along the active null
geodesic direction g ab dx a dx b  0 . For equally spaced surfaces of
Prisoner’s Dilemma: Determinism and Chaos 363

proper-time, if time intervals in the normal form basis become


compressed, then show that the corresponding size of gtt must grow
to compensate. As a consequence, show that one expects that
g tt  0 . The limit corresponds to the velocity along the null
geodesic going to infinity. The causality principle is that this limit
is not reached: g tt  0 . For a single active dimension, show that the
velocity along the null direction is:
2
g tu  g tu  g uu
c    tt   tt
g tt g  g
g g g
g tu  tu g tt  uu g uu  tt . (9.10)
g g g
g  guu g tt  g tu g tu

(2) Use Eq. (9.10) in the previous exercise to show that the requirement
of a non-zero velocity along the null geodesic requires  g uu  0 .
This is the principle of possible change that means any physical
point must be able to move in an open region, even if that region is
very small.
(3) Show that for streamlines that have speeds close to the null-
geodesic speed, the causality principle narrows the size of the
space: one gets a picture not unlike the physics picture of the big
bang with a singularity at some finite point in the past in which
space has zero size.
(4) Identify numerically the speed of the null geodesics and draw it in
the proper distance—proper-time plane. In the physical space, this
corresponds to going along the diagonals of the contour mesh.
364 Geometry, Language and Strategy—Vol. 2

(5) As an example of Ex. 4, take the model of Sec. 9.6 having high
stakes and low  and deduce that the vector direction   u along
du transformed to the co-
moving frame is given by
E    u  E  u , Fig. 9.25.

Figure 9.25. Proper-strategy vector Figure 9.26. High stakes, low  ,


E u E u  high stakes, low  null geodesic c
There are clearly three areas
where the magnitude of the vector vanishes, corresponding to where
the streamlines bunch in Fig. 9.22. Also show that the speed of null
direction c Eq. (9.10), is given by Fig. 9.26. How does the
corresponding speed c behave? Show that the speeds can exceed
unity, though they are not infinite.
(6) Show that the active metric components Eq. (9.4) can be effectively
written as the product of three matrices and so the determinant is
given as stated below. Since for reasonable values of the variables,
the determinant should not vanish or change sign, conclude that the
coordinate vector along u can’t vanish. What does this imply for
Ex. 5?
T
 Eu E u   1 0   E u E u 
g ab
  t    
E E t   0 1  e e   E t E t  . (9.11)
det g ab   1  e e  E u E t  E t E u 
2

(7) Show that another way to examine the behavior of high stakes and
low  in Ex. 5 is to look at the timeline and spaceline contours,
Fig. 9.27, associated with the time vector E t E t  , (“vertical”
Prisoner’s Dilemma: Determinism and Chaos 365

timelines) and the space vector E u E u  , (“horizontal”


spacelines).
(8) For high stakes and low  in
Ex. 5, show that the pressure,
which is a function only of 
is not static, but has the shape
given by Fig. 9.28. Note that
the pressure is in fact constant
along each of the streamlines.

Figure 9.27. Coordinate timelines


and spacelines for high stakes and
low 
(9) The proper-time   u, t  and the
proper distance     u, t   y
are each scalar functions. As an
example, show that for high stakes
and low  the scalar functions Figure 9.28. Pressure for high stakes
are given by Figs. 9.29 and 9.30, and low 
respectively.

Figure 9.29. Proper-time   u, t  for Figure 9.30. Proper-distance


high stakes and low    u, t  for high stakes and low 
366 Geometry, Language and Strategy—Vol. 2

(10) For the examples given in the


text, the flow E u along u
need not be zero. Show that for
high stakes and low  , the
behavior remains harmonic as
seen in the normal coordinate
basis, Fig. 9.31. For high
stakes and high  show that
the behavior is Fig. 9.32.

(11) As seen in Ex. 10, for high


Figure 9.31. Flow along E u for
stakes and high  , the scalar
high stakes and low 
functions for proper-time
  u, t  and proper distance   u, t  take on strikingly new
behaviors. Show that they are given
by Fig. 9.33 and Fig. 9.34
respectively. What are the allowed
regions? Argue that certain regions
are excluded and indicate why.

Figure 9.32. Flow along E u for


high stakes and high 

(12) Proper-time as defined in


Sec. 6.2 is defined in terms
of a specific path, the
Figure 9.33. Proper-time   u, t  for streamline. For the case of
high stakes and high  high stakes and small  ,
Fig. 9.22, we can arrive at
approximately the same point by two distinct streamlines from
approximately the same start. In this case we expect that the proper-
times might be different since the coordinate V  Eο is not exact
Prisoner’s Dilemma: Determinism and Chaos 367

(Sec. 5.2, Eq. (5.8)). Show


that the figure shows evidence
for this behavior. Pay
particular attention to where
the streamlines bunch
together. Why is the flow not
an exact differential?
(13) Show that the recurrence plot
for the prisoner’s dilemma
model with parameters set by
Sec. 8.2 is provided in Figs.
Figure 9.34. Proper-distance   u, t  9.35 and 9.36. Note that these
for high stakes and high  plots can be computed using
the parametric representation
for a three dimensional plot: for every pair of values    , one
can compute the three dimensional value below. One can then
restrict the region of interest by  . With a small third dimension,
one essentially has the recurrence plots normally used with only
implicit expressions of the functions in terms of normal-form time:

t  ,  , t  ,  , 
E a  ,   E a  ,  .

Figure 9.35. Strategy flow recurrences Figure 9.36. Time flow recurrences for
for single harmonic PDM single harmonic PDM
(14) Show that with the two frequency
PDM, not only are there non-trivial recurrence relations in time
along a streamline, we can have non-trivial recurrence relations in
368 Geometry, Language and Strategy—Vol. 2

(strategy) space for fixed time. In particular discuss Fig. 9.37 in


which we fix the normal form time t0 and plot the strategic flow
recurrences E u  u, t0   E u  u, t0    .

Figure 9.37. Strategy flow recurrences for two-frequency PDM

(15) A simple model of a pendulum and of decision process model has


been presented on the web [Thomas G. H., 2013] with detailed
calculations of using the differential equations to look for chaotic
behavior. There are Mathematica models available to do these
computations. Use this approach to better understand the distinction
between the steady state approach and the initial value approach.
(16) We have employed two methods of computations: one based on
harmonic polynomial solutions to Eq. (6.44) and the other based on
harmonic solutions ei to Eq. (9.2). Note that the first method
generates unbounded solutions as proper time goes to infinity and
the second does not. Given that we can make either solution meet
our initial conditions at time zero, can you use this fact to argue that
chaotic solutions must be possible?
Chapter 10

Robinson Crusoe Economics

In the last chapter we developed the prisoner’s dilemma scenario with a


single active strategy and three codes of conduct. We showed the
importance of codes of conduct in addition to choices of active strategies.
The codes of conduct provide hidden degrees of freedom and allow a
payoff matrix when viewed in the normal frame.
In this chapter we consider the Robinson Crusoe economics, which is
a simpler time dependent decision process with one active strategy, one
player with one inactive strategy and no additional codes of conduct. The
only active strategy available for Crusoe is based on foraging on his
island. His personal utility and engagement with this activity may also
change depending on how deeply involved he is in foraging.
We use this example to identify a systematic set of steps for
discussing decision processes. Part of this systematic approach is to have
a decision process taxonomy, not unlike the one that exists for game
theory, which played a central role in its development [Von Neumann &
Morgenstern, 1944]. Games can be zero-sum or non-zero sum; they can
have multiple players starting from one to any number. The players can
cooperate or choose not to cooperate. These distinctions have led to a
focus in the literature on specific aspects of games based on this detailed
taxonomy.
Decision process theory shares some of the same concepts of game
theory, while introducing a set of additional concepts suggesting that the
game theory taxonomy needs to be modified. In this chapter we start
with a discussion of the types of changes needed (Sec. 10.1). This leads
naturally to a discussion of how to categorize the different possible
decision processes as part of a systematic approach for analyzing

369
370 Geometry, Language and Strategy—Vol. 2

decision processes. We illustrate that approach with Robinson Crusoe


economics.
We start by identifying the story, which in this case is the Robinson
Crusoe economics, Sec. 10.2, noting both the choice aspects as well as
the constraint aspects of the story. We then put the story into normal
form, Sec. 10.3. We identify the natural units to be used in Sec. 10.4. We
proceed to the co-moving coordinate basis for the steady state behaviors,
Sec. 10.5. For numerical work, we identify the boundary conditions
called the “known behaviors,” Sec. 10.6, from which we compute the
steady state behaviors, Sec. 10.7 and generate from these the harmonic
steady state behaviors.
The harmonic behaviors are characterized by frame-waves, which are
detailed in Sec. 10.8. There are limits to choice based on the non-game
theoretic constraints, Sec. 10.9. The time dependence of choices for
frame-waves can be travelling waves, Sec. 10.10 or standing waves, Sec.
10.11. Such waves propagate with attenuation, illustrated in more detail
in Sec. 10.12. There are implications to the decision process values, Sec.
10.13, which is our extension of the game theory concept of value.

10.1 Decision Process Taxonomy

As an example of aspects of decision process theory that are extensions


of game theory, we have established that decision processes exhibit
stresses and strains, which help characterize dynamic behaviors. Another
extension is the significance and incorporation into the theory of a code
of conduct. We have illustrated the importance of setting the code of
conduct, noting it has profound influence over the dynamic behaviors of
decision processes.
There are many possibilities of how these concepts inter-relate in
realistic games, so it is useful to have a systematic program of
identification for all decision processes and a plan for applying decision
process theory to each. We start by suggesting the components for such a
systematic approach to and taxonomy of, decision processes. This puts
into context the prisoner’s dilemma from the last chapter, the Robinson
Robinson Crusoe Economics 371

Crusoe model in this chapter, and a general set of two-person games that
we will deal with in Chaps. 14-18.
We start with the practical steps needed to solve decision process
problems at every level, at least for steady state solutions (those
analogous to AC circuits). The steps are:

 Identify the story behind the decision process.


 Reframe the story so that it is in normal form.
 Identify the natural units. How do we measure time, utility etc.
 Transform to the co-moving orthonormal coordinate basis, since this
helps isolate the steady state behaviors.
 Use the player fixed frame model (or a model that satisfies the steady
state hypothesis) as an initial model and find the co-moving frame
solutions, which are steady state. Apply numerical methods to find
these solutions. We have carried out this program in Chaps. 14-18
with more than one active strategy using Mathematica [Wolfram,
1992] along with standard techniques that have been used in other
areas of applied mathematics and engineering, [Courant & Hilbert,
1962].
 Apply decision process theory to identify stable or steady state
behaviors that serve as known (boundary) conditions. Use these
behaviors to create the frame-wave harmonic travelling wave and
standing wave solutions from Chap. 6.

Based on this process, a natural taxonomy of decision process


extends the usual game theory taxonomy by focusing on the number d a
of active decisions: one active, two active, etc. Behind this
decomposition will be the number d i of inactive decisions that reflects
both the number of players and the number of codes of conduct. In this
decomposition, time is assumed to be present as part of the geometry but
not listed in the decomposition. Processes are classified in terms of a
strategic decomposition based on the ordered pair a, i :  d a , d i  . So
for example in the prisoner’s dilemma of the previous chapter, there are
two players each with two strategies. We treated this as a single active
strategy model with five inactive strategies including the two associated
372 Geometry, Language and Strategy—Vol. 2

with each player: the strategic decomposition is the ordered pair


a, i : 1, 5 .
We expect each player to have (at least) one inactive strategy.
In addition, there may be additional inactive strategies corresponding to
the codes of conduct. Thus we may indicate this as the irreducible
decision process

a, i  p  c :  d ,
a 
di  d p  dc  .

Thus for the prisoner’s dilemma with a single active strategy, we have
a, i  p  c : 1, 5  2  3 . If we try to enumerate the different
possible cases, we rely on the separation between active and inactive
strategies. We realize that a given decision process may change
depending on how the participants identify the codes of conduct. We also
recognize that there is a difference between the codes of conduct and the
inactive strategies associated with a given player: a player has sole
control of his or her inactive strategy, whereas a code of conduct is
controlled by multiple players. We suggest this notation as being a useful
way to decompose the decision process.
Ordinarily, we imagine that each player controls their own inactive
strategy. They have a role in controlling the various codes of conduct.
They may also control one or more of the active strategies. If a player
controls no active strategy, we call that player dependent. To make this
concept more precise, we may need to establish how that concept enters
the theoretical description. This is currently an open question.
We use the prisoner’s dilemma as an example of how this general
decomposition provides useful insight. In Sec. 6.8 we introduced the
code of conduct as a1 symmetric decision process in which each inactive
strategy that is part of the code of conduct is associated with a new
distinct player. Thus we start with the most general solution of the
prisoner’s dilemma in the normal-form coordinate basis and consider
four active and two inactive strategies, a, i  p  c :  4, 2  2  0  .

1
We use the word symmetric to indicate a technical aspect of the theory, which is the
existence of a frame of reference in the differential geometry that treats inactive and
active dimensions on the same footing. See for example the discussion in Sec. 2.11.
Robinson Crusoe Economics 373

This general decomposition can manifest in a number of ways


however. We can partition the 4 active strategies in five distinct ways,
depending on which player controls the strategies:

4  0  4, 1  3, 2  2, 3  1, 4  0 .

In addition there are different ways to pick which strategy is controlled


by which player. We have to identify these distinctions when we
characterize the known behaviors. For the general case of the prisoner’s
dilemma, we can pick 2 strategies for each player and specify which
player controls which strategy.
When expressing the known behaviors, we connect the strategy flows
and payoffs with the corresponding flows and metric (spatial and time
bonding) potentials in the natural holonomic basis. We can then
transform these values to the symmetric normal-form coordinate basis in
which the code of conduct attributes are exhibited as new players.
It is not uncommon when discussing taxonomies [Von Neumann &
Morgenstern, 1944] to propose a general program for solving each
general class and in that way build up support for the view that the theory
has general applicability. In the last chapter we made a start and have
identified some general principles.
In this chapter, we take a further step and work out the details of the
simplest single strategy decision process that has a single inactive
strategy corresponding to the strategic decomposition a, i : 1, 1 ,
which illustrates Robinson Crusoe economics. This has also been
discussed by Von Neumann & Morgenstern, [1944, pp. 9, 15, 31, 87,
555]. The model illustrates the systematic approach required for the
general case. In Chaps. 14-18, we extend this approach to two-person
games, and so suggest that the theory applies to the general case.

10.2 Identify the Story—Robinson Crusoe

The Robinson Crusoe scenario is based on the strategic decomposition of


a single active and a single inactive strategy, a, i : 1, 1 . In game
theory, we imagine a single player with a single choice to be made. The
solution to this situation is thought to be one in which the player simply
374 Geometry, Language and Strategy—Vol. 2

optimizes his or her choice. It might be thought that the game has little
strategic interest.
We shall keep an open mind. As part of keeping an open mind, it is
valuable and instructive to revisit the story on which this example rests.
Note that we might not be as sympathetic today towards Robinson
Crusoe as earlier readers.
The usual economic interpretation is based on the story by Daniel
Defoe in which Robinson Crusoe is stranded on an island. Before he
meets and rescues Friday, he survives through his “own efforts” of
hunting and farming. Here’s the plot summary from Wikipedia, [2008]:

“Crusoe leaves England setting sail from the Queens Dock in Hull on a sea voyage
in September, 1651, against the wishes of his parents. After a tumultuous journey that
sees his ship wrecked by a vicious storm, his lust for the sea remains so strong that he
sets out to sea again. This journey too ends in disaster as the ship is taken over by Salé
pirates and Crusoe becomes the slave of a Moor. He manages to escape with a boat and is
befriended by the Captain of a Portuguese ship off the western coast of Africa. The ship
is en route to Brazil. There with the help of the captain, Crusoe becomes owner of a
plantation.
“He joins an expedition to bring slaves from Africa, but he is shipwrecked in a storm
about forty miles out to sea on an island near the mouth of the Orinoco River on
September 30, 1659. His companions all die; he fetches arms, tools and other supplies
from the ship before it breaks apart and sinks. He then gets battered by huge waves as he
struggles to make it to an unknown island. He proceeds to build a fenced-in habitation
and cave. He keeps a calendar by making marks in a wooden cross he builds. He hunts,
grows corn, learns to make pottery, raises goats, etc. He reads the Bible and suddenly
becomes religious, thanking God for his fate in which nothing is missing but society.
“He discovers native cannibals occasionally visit the island to kill and eat prisoners.
At first he plans to kill the savages for their abomination, but then he realizes that he has
no right to do so as the cannibals have not attacked him and do not knowingly commit a
crime. He dreams of capturing one or two servants by freeing some prisoners and indeed,
when a prisoner manages to escape, Crusoe helps him, naming his new companion
"Friday" after the day of the week he appeared and teaches him English and converts him
to Christianity.
“After another party of natives arrives to partake in a grisly feast, Crusoe and Friday
manage to kill most of the natives and save two of the prisoners. One is Friday's father
and the other is a Spaniard, who informs Crusoe that there are other Spaniards
shipwrecked on the mainland. A plan is devised where the Spaniard would return with
Friday's father to the mainland and bring back the others, build a ship and sail to a
Spanish port.
“Before the Spaniards return, an English ship appears; mutineers have taken control
of the ship and intend to maroon their former captain on the island. The captain and
Crusoe manage to retake the ship. They leave for England, leaving behind three of the
mutineers to fend for themselves and inform the Spaniards what happened. Crusoe leaves
the island on December 19, 1686.”
Robinson Crusoe Economics 375

In this story, Robinson Crusoe is evidently not alone since he has the
help of Friday and others. Moreover, the outcome for him is dictated as
much by the morality of the times as the economic situation. His
worldview and in particular his self or independence, plays a key role and
must be considered along with the payoff aspects of his decision
processes. There are also external and historical constraints that come
with the story, which are normally ignored in an economic discussion.
This simplest economic situation is therefore one in which there is a
single agent and two strategies: a persistent internal inactive self-strategy
associated with the player’s interest flow and an active external active
strategy that through the player’s actions, produces outcomes. To this we
need to add the constraints that go beyond the statement of the number of
strategic choices. We will do this in decision process theory by a
specification of the stresses (pressure) and strains (bonds) that form the
boundary conditions of the time dependent problem.
We characterize the Robinson Crusoe story as one in which a subject
forages (hunts and farms), whose sole strategy is to determine how much
or how little of this he or she should do. The assumption is that there are
sufficiently abundant resources so that no choice needs to be made on
which approach to take (hunt or farm).
It makes sense that anybody stranded on a desert island starts out
with significant prior experiences. Just as in the Robinson Crusoe case, a
subject brings with them a set of fundamental beliefs, which is their
worldview (Cf. Sec. 11.2). Someone brought up to be interdependent
might be more likely to conserve even with super abundant resources;
they might save for a rainy day or a day in which someone else might be
stranded. They might treat Friday better. Someone brought up to be
independent and a little self-centered might strive only to optimize his or
her own pleasure. Thus there are behaviors of strategic interest even in
this simple example. We show how decision process theory handles such
behaviors and constraints, starting in the next section where we put this
story into normal form.
376 Geometry, Language and Strategy—Vol. 2

10.3 Put Story into Normal Form

Somebody stranded on an island will make many decisions during the


course of the day. Our ordinary way of thinking about decisions is a
series of sequential actions in time, called the extensive form. We
normally don’t think through all of the possible consequences that flow
from each single decision in time. It is like a chess game in which we
consider each move based on what the opponent has done; on how the
board appears.
However, we can make our plans based on our first move,
considering each possible move by the opponent and what our counter
move would be and so on. This is called the intensive form. We would
create a very long list of possibilities, which we have called pure
strategies in normal form. Our opponent, if there is one, can also make
such a list and would thus create the opposing team’s set of pure
strategies. The play of the game of chess would be a single choice by us
and a single choice by our opponent. This is called putting the process
into normal form. For convenience we label each string of choices with
an informative name, which should not be confused with the myriad
individual choices that need to be made to actually carry out the decision
process.
With that in mind we characterize the person stranded on the island as
having a single pure strategic choice to make, which we characterize by
foraging, which we take to be an informative name. It represents
possibly hundreds of individual decisions during the day, all of which are
characterized by this name. The degree to which foraging occurs is
indicated by the value of a single active strategy variable, u . Our
convention is that positive values are indicative of actively foraging and
negative values are indicative of restraint or resistance to foraging
(conserving or possibly just being lazy).
In addition to the activity of foraging, there is the mindset or
worldview of the person stranded on the island, which we characterize by
a single inactive strategy variable j . We begin with the idea that a
positive value indicates an internal belief in the value of getting as much
off the island as possible, whereas a negative value indicates a desire to
Robinson Crusoe Economics 377

conserve what is there and take only the minimal necessary to insure
personal survival.
This strategy however is inactive; nothing material changes based on
the value of this strategy. What gives Crusoe pleasure or pain does not in
fact change the decisions that Crusoe makes, which are assumed to be
solely associated with foraging. However, because of these internal
desires, there exists a value or player interest  that Crusoe can assign
to the outcome of the decision process. This value exists because of the
hidden variable of the inactive strategy, but does not depend on the value
of that inactive strategy, only the active strategy of foraging.
The dynamics will depend on this decision value along with the net
result of the constraints of being on the island. There will be stresses or
forces that may move Crusoe to take actions (invasions by other
islanders), that are not strictly part of the problem at hand. These will be
captured in terms of the pressure of the system, p . There are bonding
forces  as well as the strength of engagement e that relate how
quickly or not Crusoe learns from his experiences. Most of these effects
are not important from a static view but dictate how strategies are forced
to change in time.
The normal form has a single active and single inactive strategy,
 i : 1, 1 , corresponding to a 2  1 -dimensional space-time.
a ,
j
There is
a single player interest field with non-zero component F tu ; there are no
purely spatial components of the payoff and so no payoffs that
correspond to ordinary game theory. There is only the game value
represented by this time component. We use the single strategy model
and frame-wave solutions from Sec. 6.8.

10.4 Identify the Natural Units

In a static theory, there is no need to discuss units. In a dynamic theory


however, we want to know the duration of time between a cause and an
effect. For the Robinson Crusoe economics, we imagine time as being
marked off in days, with each day a complete play of the strategy. This
incorporates a complete play. The natural velocity is the effort Robinson
Crusoe can do in one day. Utility is measured in the same one-day
378 Geometry, Language and Strategy—Vol. 2

equivalences. In the next section we analyze this model in the co-moving


coordinate basis.

10.5 Co-Moving Coordinate Basis

Our goal is to separate the aspects of the behavior that are steady state
from those that are time dependent. Our method is to imagine that we
move in a frame in which the system appears to be at rest: the co-moving
coordinate basis. For the models we have been considering, the various
attributes of the theory will be independent of time in that frame.
The flow equations in the co-moving frame, Table 5.1, are expressed
in terms of the flow direction  , the proper inactive strategy direction
 and the proper active direction  . The flow equations depend on the
acceleration q   q , a single vorticity or player interest    , the
strategy  , the player bonding coefficient      and no shear
components, Eq. (6.46):

 E j   E j
 e  2   q    e
 E j   q E j  2 E j  E  E j e . (10.1)
f j
  2 jj E j  q e   1  e e    e  1   jj E j E j 
 jj  E j E j  E j E j   E j E j 1  e e 

The acceleration, player bonding, vorticity and pressure are determined


from the tidal, electromagnetic and inertial equations from Chap. 6, Eqs.
(6.50), (6.48), (6.47) and (6.49), which we summarize here:

 q  q 2   q  2  p  p 
   2 q
   3 2   2   . (10.2)
p 
   q  
 2

p  0
Robinson Crusoe Economics 379

The inactive pressure p and energy density are undetermined. For this
investigation we make the simplifying assumption:

p   p    p
. (10.3)
  r p

As before (Sec. 7.10), we call the parameter  r the resiliency.


We observe the following about pressure, which is the average stress.
Though it represents an external constraint that is often ignored, it is not
something that should be ignored to the extent that pressure represents
energy and momentum that may influence behavior. We live in an
atmosphere of air, which we require to breathe, which generates our
weather, and on occasion generates storms and hurricanes. It can be
important. We find that it directly influences the limits to choice,
Sec. 10.9.
In decision process, the pressure represents an average of all of the
effects on Robinson Crusoe that are not relevant to the details of the
decision making; yet these effects carry energy and momentum and as
such can’t be ignored. Crusoe’s history from England, his sea
experiences, his morality that dictates how he sees the islanders, all are
effects that carry energy and momentum but are not strictly speaking part
of the active strategy of foraging.
By making assumptions about the pressure we make assumptions
about the forces or acceleration that occur in the model. This represents a
“network” force. In particular, based on the assumptions Eq. (10.3), we
determine the form of the acceleration from Eq. (10.2):

 q  q 2   q . (10.4)

We see that if the boundary condition on the force or acceleration is zero,


then the acceleration stays zero at all other strategy values.
Foraging is assumed to range between conserving and consuming. If
one assumes that at one point between these extremes there is a force to
change (either to consume more or conserve more), then the flow of
foraging will change between these extremes as well. If there is no force
to change at one point, then the strategy flow will be constant between
380 Geometry, Language and Strategy—Vol. 2

the extremes. These arguments are steady state arguments, not time
dependent arguments.
In addition to the steady state equations, we have the frame-wave
phasors, Eq. (6.52) after we eliminate the time dependence:

  x a   2 e 2 q 1  e 2  x a     q   x a  0 . (10.5)

Solutions to these equations can be travelling waves, standing waves or


transient solutions depending on the boundary conditions.
For the known behaviors, we start with zero flow along the active
strategy and zero flow along the inactive strategy. We look to the
equations to see whether this is maintained. As in the prisoner’s dilemma
model, we solve these equations numerically using Mathematica
[Wolfram, 1992]. We treat the known behaviors in more detail in the
next section.

10.6 Known Behaviors

By known behaviors or known values we mean the behaviors on the


boundary   0 . The essential aspects of the solution depend on Eqs.
(10.1), (10.2) and (10.5), with appropriate boundary conditions. The
active strategy variable is a measure of how much or how little foraging
will be done. There is freedom to choose the co-moving frame for the
flows as mentioned above. This is the frame, in which the flow is along
the time direction so the space components vanish, V j  V   0 . In this
frame, the time component of flow V   e  defines a gravitational
potential  .
The vorticity parameter or player interest field  is related to the
decision value field, Eq. (10.1):

f j
  2 E j  q e   1  e e    e  1   jj E j E j  . (10.6)

We shall use the two concepts interchangeably since on the starting


boundary, one determines the other. The vorticity is a rotation in the
   plane, so we can use this equation to understand the relationship
Robinson Crusoe Economics 381

between the player’s active strategy choice and their inactive strategy
choice.
If we take the boundary decision value field to be positive, then
Eq. (10.6) determines that we pick the player interest to be negative.
Note that there are two solutions to player interest so that we do have to
arbitrarily set what we mean as a positive for one of the fields.
We think it reasonable to assume that at the known position, there is
no player engagement. Based on the player interest being negative, this
implies that the player engagement is positive towards less foraging and
negative for more foraging.
The known values of the transformations at the origin are taken so
that the co-moving basis and normal-form coordinate basis are aligned at
the origin   0 :

E u  E j  1
. (10.7)
E u  e E u  E j  0

If we take other values for the known flow at the origin, then we use the
orthogonalization process described in Sec. 7.6 to determine these
transformations.
If we look at Eq. (10.2), we note that we have not said much yet
about the compression or strategy bonding parameter  . We have
found that large values of strategy bonding determine how quickly
energy transfers from motion to fields as we move along the strategy
direction. Thus this will also impact the relative strength of foraging
between the extremes. We have again two possible sign choices for the
bonding: we have chosen it to be positive, which yields a large available
space for what we term to be conserving. The equations are symmetric in
that there is another solution with bonding negative in which the larger
space is for foraging.
The remaining scalar acceleration, stress (pressure) and strain (player
bonding) q p   are set to represent a solution that has a small
force towards more foraging, a small stress and a strain  that yields a
negative vorticity from Eq. (10.2):
382 Geometry, Language and Strategy—Vol. 2

q p    0.01 0.01 2.0


. (10.8)
  0.1732...
The units are   1 and the resilience  r  5 . It may be noted that these
values are not dissimilar to those used in the prisoner’s dilemma for the
sensitivity analysis. Our investigation here is a thought-experiments that
suggests that these nominal values are reasonable since they provide
insight into the theory. In the next section we consider the consequences
of these known behaviors on the scalar behaviors that are steady state.

10.7 Steady State Behaviors

For the prisoner’s dilemma, our known behavior was for zero
acceleration and a maximum pressure. However we found in our
sensitivity analysis, other solutions that had zero acceleration and a
minimum pressure. We are reminded
of weather in which we have both
high pressure and low pressure points.
Based on the known behaviors
from Sec. 10.6, we find solutions in
which the pressure is a minimum, Fig.
10.1 and for the acceleration, which
goes to zero, Fig. 10.2. Recall that if
Figure 10.1. Pressure p   we set the known value of the
acceleration to zero, it will be zero for all strategic foraging values.
A possible rationale for a small
starting pressure is that Robinson
Crusoe has been on the island for
some time and that there may be a
foraging choice that he has
determined where he has no particular
recollection of any historical or island
constraints (such as invasions) to
contend with. In other words, there is Figure 10.2. Acceleration q  
very little constraint at this point. He is not engaged: e  0 . There may
Robinson Crusoe Economics 383

be constraints elsewhere, as we see in the figure, which does grow as you


move towards more conservation.
Note that it is easy in the analysis
to dramatically increase the pressure
and determine the consequences of
that assumption. The advantage of a
deterministic model is that it allows
a number of “what-if” questions to
be asked and answered.
Figure 10.3. Player bonding   

The solutions are different in that


there are limits to choice: the space is
finite and the choice region gets
progressively smaller as we increase
foraging, Fig. 10.3. Because the
player bonding rises so precipitously,
the volume element of space, the
Figure 10.4. Player vorticity   
choice region, goes to zero. Here as in
the figures for the acceleration, the
value never quite gets to zero. The
acceleration change is balanced by
the vorticity, Fig. 10.4. The net effect
seems to be consistent with a steady
state flow that goes to the right but
progressively slows down and never
passes a certain boundary or size.
Figure 10.5. Player engagement
A way to appreciate what is
e  
happening is to consider the player
engagement, Fig. 10.5. The engagement is negative as we move to the
right and positive as we move to the left; it is set by the negative value of
the player interest through Eq. (10.1). Based on our sign conventions for
the decision value and hence for the player interest, we are associating
the large positive value for conserving on the left, and a negative value
for foraging on the right.
384 Geometry, Language and Strategy—Vol. 2

The frame transformation2 vector


E j along the inactive direction gives
another measure. It has the known
value of unity at the origin and
decreases to zero as we go towards
more foraging, Fig. 10.6. From this
we obtain the charge E j   e E j ,
Eq. (10.1).
Figure 10.6. Frame transformation
E j  

The charge and the player


engagement are zero at the origin,
Fig. 10.7. We associate the value of
the decision process with the time
component of the field, which in the
co-moving frame is f j , Fig. 10.8.
Figure 10.7. Player charge E j As the charge increases, we get an
increasing decision value. This appears to reflect the same phenomena as
in game theory for the one person game: maximizing the utility is the
purpose of the single strategy. The more foraging that goes on, the more
the activity is independent and the higher the game value. This somewhat
justifies our sign conventions.
There appears to be no value
towards acting interdependently; as
we move to the left, the value of the
decision process stays and gets
progressively smaller. It is important
to note however that in decision
process theory, this situation
represents only one of two possible
solutions to Eq. (10.2). We can Figure 10.8. Player interest f j
change the sign of both the
acceleration and the strain and again obtain a solution to the equations
keeping the same value of the vorticity. The decision value will again be
2
This also represents the “Killing” vector in the orthonormal co-moving frame: it is the
direction along which there is a conserved charge, Sec. 4.5.
Robinson Crusoe Economics 385

positive, but will display its maximum on the left corresponding to a


restraint on foraging and a tendency towards interdependence not
independence.
There are solutions in which the decision value is negative,
corresponding to picking the positive solution for player interest. In this
case the player engagement to the right of the origin becomes positive
and to the left becomes negative. We have no interpretation for this at the
moment. A tentative interpretation is that to the right, the environment is
inhospitable. The more one interacts with it, the more likely it is that one
will be hurt. In this case foraging is not such a good activity. The safest
approach is to be more of a conservationist and hope to come out at zero
value. But again there will be the opposite of this solution in which
interacting with the hostile environment can come out to be the best
approach and being interdependent is less successful.
Which of these choices is the best? What we have demonstrated is
that decision process theory provides a common ground for discussion.
In fact we have only scratched the surface. We could go into much more
detail using large values of pressure and acceleration, which widens the
set of possibilities and behaviors. To determine what governs a given
situation, a measure of the acceleration, the strengths of the stresses and
strains are expected to provide the determination. One way to see the
effects of the stresses and strains is to consider the harmonic effects. In
the next section we consider the frame-waves that result from a harmonic
stimulus.

10.8 Frame-Waves

A key insight from frame-wave solutions to Eq. (10.5) comes from


following the disturbance at the boundary over time. The disturbance can
be a pulse disturbance, a travelling wave to the right or to the left, or a
standing wave, to name just a few. If we do a complete analysis of
travelling waves, we can then construct pulse waves or standing waves
using these travelling waves.
Given the non-linear nature of the equation, we don’t necessarily
know that the solutions will follow this pattern, which is common in
386 Geometry, Language and Strategy—Vol. 2

other areas such as engineering and physics. However, in this section we


will verify that this expectation is met, with some caveats about the size
of the space; the choice region. We find travelling waves that propagate
from the boundary with attenuation due to the non-linear aspects of the
equation.
Though we use the word attenuation, we don’t imply that the energy
is unrecoverable, like it is in a resistor; it is more like an inductor where
the energy moves to a magnetic field and can be recovered. The energy
has to go somewhere and the two possible fields in this model are either
the player interest field or the metric bonding fields. We find that there
will be waves in each of these fields.
In the prisoner’s dilemma, Sec. 9.6, we saw hints of such wave-like
behaviors not only in the flow vectors but in the space and time bonding
fields and the value fields. We explore this behavior in more detail and
conclude that there are indeed travelling gravity waves, which will be
discussed in Sec. 10.9. Such waves are a consequence of the harmonic
source and propagate as a consequence of the inertial media of the space
time.
What we find is in fact evidence in the solutions of the theory for
steady state wave phenomena analogous to AC circuit behaviors. If these
waves propagate, there is a sense in which they are related to waves in
the theory, Sec. 6.6.
We recall Ex. 37 from Sec. 6.10, which suggests that we look at the
power series solution of the exact solution of Eq. (10.5) and a travelling
 
wave approximation x aTWA  E e p :

x aTWA    E e
 p

E  x a  0  . (10.9)

 p   p  i  p  1 2   0  q 0   i  2 e 2 q  1 4   0  q 0 
2
0

For a sufficiently large frequency  , there will be a damped travelling


frame-wave approximation:

x a    x a   sin    p   p  .
 TWA  p
 E  e (10.10)
Robinson Crusoe Economics 387

Here  p is a constant phase. We can take it to be, for example 0 or 1 2 


corresponding to a sine or cosine wave dependence, or someplace in
between. The travelling frame-wave approximation matches the frame-
wave solution to first order in the strategy distance.
Though we use the word damped, the amplitude may in fact grow
indicating that energy is being given and not lost. Ordinarily, we think of
a wave as propagating with some attenuation, meaning that if  p is
positive, the travelling wave moves along the negative axis (to the left),
corresponding to the lower sign for  p . If  p is negative, we reverse the
sign of  p and identify travelling waves that move to the right. In
particular, we can have movement in both directions for  p independent
of its sign.
We will see that this approximation is pretty good around the origin,
allowing us to associate the attenuation with the boundary condition
parameters for the bonding and acceleration:

 p  1 2   0  q 0  . (10.11)

We also see that the travelling wave moves with a velocity


cp  . (10.12)
p

In general this depends on the frequency of the initial disturbance.


As an example, we provide the travelling wave solutions based on our
known conditions and assumed values for the strengths of a left-
travelling frame-wave for time,
Fig. 10.9, moving wave at a
frequency of   10 and strength
E    1 200 .
The approximation matches
the travelling frame-wave at the
origin and its vicinity. There is
attenuation, but much less
strongly in the full solution than
Figure 10.9. Left-travelling time wave and
its approximation
388 Geometry, Language and Strategy—Vol. 2

in the approximation. This difference is important.


First of all, the attenuation is not a loss in the sense of energy
disappearing from the system, but of energy being transferred from
motion to one of the bonding or value fields. Thus we expect, and will
find below, that there is a corresponding left-travelling wave in the
metric and player value fields.
Second, the fact that the attenuation is not as strong as anticipated is
an indication that there may be reflections at the boundary of the choice
region, and we are seeing the effect of a returning wave from the right.
Our wave solution is the steady state of the wave that moves to the left.
We can also investigate the
behavior of the left-travelling
frame-wave for the foraging
strategy, Fig. 10.10. The
frequency choice and strength are
independent; for this example we
have chosen the frequency to be
  10 and a stronger strength
E    1 50 .
We recall that for both frame-
wave solutions, the superposition Figure 10.10. Left-travelling strategy wave
and its approximation
of two solutions will again be a
solution. We can add solutions with different frequencies and strengths
as well as the same frequencies and strengths. In particular we can add
two solutions with equal and opposite frequencies, which can then be
viewed as a left moving wave added to a right moving wave.
We call such solutions standing frame-waves. For example with the
travelling waves from Eq. (10.10) and phases  p   0 , we have:

x a    x a    x a  
SW  

x a   sin    p  cos  p


SWA  p
 2 E  e . (10.13)
x a    x a    2 E  sin    p  cos  p
SW SWA
Robinson Crusoe Economics 389

Thus we can ignore the second term. We expect that this wave does not
travel: each part moves up and down together, albeit with different
amplitudes.
In the following sections, we will investigate the behavior of both
travelling frame-waves and standing frame-waves. We start with a
discussion in the next section on the limits to choice.

10.9 Limits to Choice

In a continuous and connected system, one that is rather complex, it is


expected on general grounds that there will be limits on the values that
attributes of the system are allowed to take. This was the basis of the
study using systems dynamics by the Club of Rome [Meadows,
Meadows, Randers, & Behrens, 1972] on the time behavior of
populations and resources. They focused on the limits to growth. We
believe their study applies equally well as a general principle for any
continuous and connected system, particularly the differential geometry
systems we have been investigating. Hence we think it appropriate to
investigate the limits to choice that exist in decision process theory.
There are certainly many sources for these limits; the advantage of
picking a simple model is that we can focus more clearly on some of
these sources. By picking a very small starting pressure to start with in
Eq. (10.8) as well as a very small starting acceleration for our boundary
conditions, we may see that there are other sources which limit choice.
One such limitation is seen in the numerical work on where solutions
become singular. One of the limitations of the solution is determined by
where the inactive metric component  jj   , Eq. (10.1) becomes zero.
The metric component is proportional to the boundary difference
determined by the player engagement:

  1  e 2 . (10.14)

We term this the boundary difference since we see that it tends to zero at
the boundaries of the space, Fig. 10.11.
390 Geometry, Language and Strategy—Vol. 2

This highlights some of our


underlying assumptions in choosing
boundary conditions at   0 . We
assume that the player engagement is
zero, which is a type of equilibrium.
The connectedness of space however
changes the player engagement
value, leading to a decrease in the
Figure 10.11. Boundary difference
boundary difference. The boundary
  
difference can’t become negative,
which we show below. Therefore there is a maximum allowed value for
the player engagement. This provides one of the limits to choice.
Player engagement is not a game theoretic concept. It relates to how a
decision maker approaches the game and how involved they are in
making decisions. Game theory assumes that this engagement can be
very large or very small; it makes no difference on how the game is
played. We see this however as an external constraint that in fact may
carry energy and momentum, may communicate as a field with different
parts of the system, and as such, may be constrained.
In this context, we note that the boundary difference is related to the
central frame metric, Eq. (6.73):

g   1  e 2  e2 q . (10.15)

This metric component is directly related to what we could legitimately


call the gravitational potential.
We note that in general, this metric, which is the time bonding of the
system, is determined both by the boundary difference above as well as
the acceleration. There are limits because we don’t want the time
bonding to be negative or infinite. See for example Sec. 9.4. This is the
source of the limits to choice. For the numerical choices made in Sec.
10.6, the choice region is approximately the interval  4.5, 0.46 .
The travelling wave characteristics are closely related to the choice
region. If it is small, then there will be waves only for very high
frequencies. Thus limitations in space also imply limits in time.
Robinson Crusoe Economics 391

An additional limit is on the formation of what might be loosely


called gravity waves: waves in the bonding or value fields. We expect
that all the fields will exhibit travelling frame wave behavior. This is in
fact what we see. But are these radiation waves? Are there any limits to
choice?
One expects that there will be gravitational radiation waves in
differential geometries. The number of independent components
2   of these gravity waves depends on the dimension of space-
1 d d 3

time d . The wave components are those components that are left over
from the total number of field variables (based on the number of
independent metric components in a holonomic frame), 1 2 d  d  1
minus the number of gauge conditions d and minus the number of field
equations that are independent of time, which can be shown to be d
based on the Codacci equations, Sec. 5.3.4.
On general grounds we see that for our specific case of a, i : 1, 1
active and inactive components, the dimension of space-time is d  3 so
there should be no gravitational radiation wave components in free
space. This is not the end of the story, however. Gauge theories generate
two distinct types of wave phenomena: what we may call near zone and
far zone waves. To pick an example, take the case of electromagnetic
waves.
There are Coulomb forces in the near zone that follow an inverse
square law. A charged fluid source that moves as a travelling wave will
generate an electromagnetic field based on Coulomb’s law in response to
that motion. That wave moves as a travelling wave with the same
characteristics. It will respond based on Coulomb’s law with a delay
equal to the speed of light, yet the wave will appear to move with a speed
characterized by the charged media. It will have additional components
as well, based on the Coulomb field.
There will be radiation fields generated anytime a charge accelerates
defined as fields that drop off as the inverse distance, more slowly than
the Coulomb field. They are present in the far zone, long after the near
zone waves have died out. These waves will appear as travelling waves
that are moving at the speed of light, not at the speed of the source
waves.
392 Geometry, Language and Strategy—Vol. 2

For the models that we have been considering in decision process


theory, when will we expect such radiation waves? There are d a  1
active dimensions (including time) and d i inactive dimensions, so the
gravitational radiation waves manifest in one of three ways: they are
part of the active space of which there are 1 2  d a  1 d a  2 
components; they are part of the inactive space and so consist of the
inactive metric components of which there are 1 2 d i  d i  1 components;
or they are part of the d i payoff potential fields Eq. (1.21), of which
there are d a  1 components for each inactive strategy. The total number
of components adds up to the same:
1
 d a  1 d a  2   1 2 d i  di  1  di  d a  1
2

 1 2  d a  d i  1 d a  d i  2   1 2 d  d  3 . (10.16)

In particular, for the player fixed frame model in the co-moving frame,
we have eliminated the time behavior of the inactive metric, which in the
norm-form coordinate basis is independent of proper-time. We expect no
radiation waves in this model for these inactive components.
There are different possible wave behaviors for the payoff and player
value fields. There are d a  1 independent wave components expected,
after we account for gauge invariance and identify the wave equations
that depend on time. We expect these waves to be like electromagnetic
field waves and be present when there are two or more active strategies.
Similarly, we expect gravitational radiation waves when there are at
least three or more active strategies. Such waves will be in the far zone,
so the choice region has to be sufficiently large.
To recap, non-game theoretic aspects of the decision process dictate
the limits to choice for both the stationary aspects of the decision as well
as the travelling wave aspects. We have taken values so that the
travelling wave behaviors can be seen clearly. We investigate such
travelling wave solutions in the next section.
Robinson Crusoe Economics 393

10.10 Travelling Frame-Waves

We have learned to expect travelling frame-wave solutions as long as we


are in the choice region. This allows a natural interpretation of the waves
as waves that propagate away from a disturbance, waves that can travel
in any direction. These waves suffer some amount of attenuation, which
stabilizes due to interactions with the fields of the system. We should see
all of these behaviors from the numerical computations.
We start by looking at the energy
flow, Fig. 10.12. We first note that
there are changes in the behavior near
the boundary. If the region on the left
were extended slightly, we would see
that the travelling wave starts to
diverge, as it does on the right. If we
stay well within the choice region, we
see that the wave propagates with
amplitude that is approximately
Figure 10.12. Strategy flows
E a  ,0 
constant.
An interpretation of this energy
flow is that if Crusoe increases foraging at his base position   0 at an
initial point in time, then this solution
shows him increasing foraging at a
neighboring strategic position at a
later point in time. This solution
shows a constant set of increases that
ripples out from the initial point. The
situation is presented in Fig. 10.13 for
the time flow, which can be thought of
roughly as the energy of the flow.
We see that there is nothing Figure 10.13. Time flow E t   , 
special about where we pick the
starting point. The travelling wave behavior is the same from each
starting point, until we get close the limits to choice.
394 Geometry, Language and Strategy—Vol. 2

We expect a similar set of


behaviors for the flow along the
foraging strategy, Fig. 10.14. There
are small differences associated with
the initial values we have set. We
assumed that for time, we would
make only small changes, assuming
that the changes would be zero to start
for the flow. The values chosen can be
Figure 10.14. Strategy flow E u  ,  seen in Fig. 10.9.
For the strategy flow, we assumed somewhat larger changes
assuming that the value of the strategy vector would be unaffected at the
origin. The values chosen can be seen
in Fig. 10.10. The amplitude is four
times larger.
The other frame transformations
follow from a similar set of
assumptions. We have no frame
transformation for time, by
assumption. We obtain Fig. 10.15 for
the strategy frame transformation.
Here we start with a unit value for the Figure 10.15. Frame transformation
frame transformation, set by the E   , 
u

known conditions. The importance of the frame transformations is that


they determine the bonding potentials g ab . The gradients of these
potentials generate forces, and the
fields carry energy and momentum.
By considering a travelling wave
for the energy flow, we have
implicitly created a travelling wave
for the bonding fields. For example,
the time bonding field, Fig. 10.16,
shows such structure.
Because of the non-linear nature
of the equations, the travelling wave
Figure 10.16. Time bonding g tt  , 
for the bonding field has more
Robinson Crusoe Economics 395

structure than each of the component frame waves. Because the bonding
field carries energy and momentum, there are constraints on its values.
We see possible singular behaviors developing near the limits to choice.

10.11 Standing Frame-Waves

The relevance of travelling waves is that they provide a mechanism for


understanding the dynamics of decision processes. We think of events
generating ripples in space and time. The structure may be complex, but
can be understood by superposing these travelling waves. As a step
towards understanding the structure that might appear, we consider in
this section standing waves formed from an equal mixture of a left
moving and a right moving wave.
We start with the time flow, Fig. 10.17, which shows a periodic
structure, though not necessarily the simplest lattice structure you might
imagine. We have narrowed the region of focus so that we are well
within the choice region and far from the limits to choice.

Figure 10.17. Time flow contour plot


396 Geometry, Language and Strategy—Vol. 2

The oscillation structure represents a standing wave. We know this by


construction since we have created it from two waves moving in equal
and opposite directions. If we look at the wave as a movie, it doesn’t
flow left or right. By Eq. (10.13), the shape at the initial time has a
periodic behavior. Thus the contour plot provides a summary of such a
standing wave pattern.

Figure 10.18. Strategy flow E u  ,  contours

We see a standing wave pattern for the strategy flow, Fig. 10.18.
With a little imagination, we make out a pattern that is not a square, but
starts to look like a hexagon. We think that such structures are not
accidental but reflect the underlying dynamics. We expect that
reflections of the non-linear dynamics will be present in each of the
frame transformations.
Robinson Crusoe Economics 397

Figure 10.19. Frame transformation E u  , 

We now look, Fig. 10.19, at the strategic space transformation, E u


in which we have imposed a smaller harmonic contribution. This contour
plot shows a different standing wave structure. The point is that the
frame transformations combine to form the bonding potentials as well as
the player value fields. These behaviors translate directly into the field
behaviors.
We highlight the behavior of the time bonding, which results from
the interplay between a left moving and a right moving travelling wave.
We argued that a real event or disturbance that was caused by Crusoe at
his base position for foraging, could propagate to the left, indicating
choices that he would make. It may be symmetric and propagate also to
the right. In each case, the energy flow that propagates in a given
direction will couple to the frame transformations by definition, and this
in turn couples to the time and space bonding potentials.
We thus expect and see a standing wave for each of the time and
space bonding potentials. We show the contours for one of them,
Fig. 10.20, which is the time bonding potential.
398 Geometry, Language and Strategy—Vol. 2

Figure 10.20. Time bonding g tt  ,  contours

The standing wave structure is more complex than what we have


seen before. There is clear indication of changes near the limits to choice
and a structure in the intermediate region that is not just a simple lattice.
These structures indicate the non-linear nature of the behavior and should
be present in actual choice data. In the next section we explore ways in
which this structure might become more pronounced.

10.12 Limits to Growth

In the last section, we saw evidence of interesting structure that is the


result of the non-linear character of the theory. The behavior is related to
two quite different mechanisms. One is the behavior near the boundary
and the other is the behavior due to the strengths of some of the
parameters in the theory, such as the bonding strength. These behaviors
set the choice region, which a limitation to choice. In this section we
show that there are also limits to growth.
We illustrate this behavior by changing the previous model to one
that is identical, except we increase the bonding by a factor of 10:
  20 . We see the result for the time bonding in Fig. 10.21.
Robinson Crusoe Economics 399

Figure 10.21. Time bonding gtt for bond   20

Thus if we increase the bonding strength, we shrink the choice region


dramatically. If we maintain the same frequency, there is no longer
enough room to see the oscillations in space and so the spatial region in

Figure 10.22. Time bonding contours for gtt and   20


400 Geometry, Language and Strategy—Vol. 2

essence is now part of the boundary. We see the “cloverleaf” structure,


which can be seen in the contour plot for the same time bonding in
Fig. 10.22.
One way to interpret the phenomena is that our frequency is
essentially too low to see travelling wave structure. We have verified this
from the numerical results. Another way to confirm this is to take the
original base line model and decrease the frequency. We have done that
and again are able to reproduce similar results to the figures above.
We draw important lessons from even this simple example with a
single active and single inactive strategy. Non-linear effects may wash
out wave phenomena but may create limitations in time as well as space.
We see these structures reflected in other attributes of the solutions.
See for example Exs. 7-8.
We conclude that the boundary behaviors are essential attributes of
the theory. These boundary values follow not from the game theory
concepts but from the new non-game theoretic concepts such as player
bonding, player engagement, stress and bonding fields. If we operate
well away from the limits to choice, these non-linear effects play less of a
role and we recover the expected behaviors. However this is not true if
the space is small, the frequency is small, the stresses and bonding are
large, etc. We will see one more example of this in the next section as
applied to the decision process values.

10.13 Decision Process Values

In decision process theory, we


replace the game value with the
decision process value, F j tu  ,  .
Based on the known behaviors, we
start with Fig. 10.23. We see that this
is a standing wave. We have provided
only the behavior that is well within
the choice region. If we were to
extend the region, we would see that
Figure 10.23. Known behavior of
the decision process value gets large
F j tu  ,0 
Robinson Crusoe Economics 401

also at the left. So in the middle region, the standing wave represents the
way in which the value of what Crusoe does propagates. In this case, we
see the result of a standing wave with equal propagation left and right.

Figure 10.24. Decision value F j tu  , 

Another way to view behavior is Fig. 10.24. Based on an


understanding of the structure, we can now envision what might happen.
If there is a major event at the base position, it will propagate outward,
possibly in both directions. There will be unanticipated consequences
because of this. Strategies that might be thought to be less effective will
be tried. If we get too close to the boundary, we will start to feel the
effect of stresses that were always there but were ignored. The long
history that we provided in Sec. 10.3 may now give insights into the
nature of these stresses and strains. We note that even in the story,
Crusoe goes home so that the stay on the island was not in reality a
closed system.

10.14 Outcomes

We have employed a general approach to decision process theory and


used it to analyze a simple model of a subject who has a single active
decision available and a single inactive strategy to study the interplay of
these decision variables to external stresses and strains. We argued that a
402 Geometry, Language and Strategy—Vol. 2

study of the response to harmonic stimulus provides insight into the


structure of these stresses and strains.
Further, we provide support for the proposal by Thomas & Kane,
[2008] that the inactive strategy is related to the worldview of the subject.
The travelling wave solutions reflect an important attribute of decisions.
It is similar to a physical system: if we collect matter or charge in a
confined space, the matter generates pressure and the charge generates an
electric field. These provide forces that inhibit more matter or like-
charges from collecting and the result will be to create new dynamic
structures. These conclusions are ones that may be applied to decision
processes.
In this chapter we make the argument that decisions at a point in time
around a specific strategic area are events. They are disturbances that
generate ripples in both space and time. As in areas of engineering, we
may study such disturbances as travelling waves. We find that such
travelling waves manifest not only in the energy propagation, but in the
bonding and value fields. We thus have a new way of looking at
decisions in a quantitative way. We can decompose events in terms of
their propagation values, just as electrical engineers decompose circuit
behavior in terms of travelling voltage or current waves.
The attainment of the outcomes of this chapter is facilitated by doing
the exercises in the following section. Based on this investment, the
student should achieve the more detailed outcomes below based on
section.
 From Sec. 10.1 understand how the game theory taxonomy extends to
decision process theory.
 From Sec. 10.2, the student should learn the importance of clearly
articulating the underlying story or premise of the decision process. It
is particularly important to isolate factors that may carry energy and
momentum and become part of the boundary behaviors.
 From Sec. 10.3, the student understands the value of reframing the
sequential story into a decision process in normal form.
 From Sec. 10.4, the student learns that the natural units set the utility,
time and action scales for the decision process.
 From Sec. 10.5, the student learns the value of transforming the
problem from the (symmetric) normal-form coordinate basis to the
Robinson Crusoe Economics 403

co-moving coordinate basis. The problem has the same content in the
theory; each frame provides valuable insight into the problem. This
basis in particular helps isolate the steady state behaviors.
 From Sec. 10.6, the student must identify all the variables that must
be initially specified and provide known values for these variables.
 From Sec. 10.7, the student isolates the steady state behaviors of the
problem.
 From Sec. 10.8, the student has learned how to decompose the
solutions to the frame-wave equation into travelling waves. Part of
this decomposition involves identifying the approximate solutions.
 From Sec. 10.9, the student learns the importance of identifying the
limits to choice, and quantifying that as the choice region. The
approximate travelling wave solutions may help to characterize the
choice region.
 From Sec. 10.10 the student will have learned the characteristics of
the travelling wave solutions.
 From Sec. 10.11 the student will have learned the structures one
might expect for the standing wave solutions.
 From Sec. 10.12 the student learns more about the boundary
behaviors, which can set time limitations as well as limitations to
choice. Long time frame events will have different dynamics than
short term events, which will be more similar to game theory
expectations.
 From Sec. 10.13 the student learns more about the decision process
field.

10.15 Exercises

(1) Because of the structure of the field equations of decision process


theory, distinctly different games are characterized by the strategic
 
decomposition a , i  p  c :  d a , d i  d p  d c  . We have a fairly
complete solution for processes of the form a, i : 1, di  , along
with two examples: the prisoner’s dilemma and the Robinson
Crusoe example. What examples can you think of that illustrate
game with more active strategies? We expect processes to reflect
404 Geometry, Language and Strategy—Vol. 2

increasing complexity as we change the number of active strategies.


So we see the progression as:

a, i  p  c : 1, 
1  d p  dc 

a, i  p  c :  2, 1  d  d 
p c

a, i  p  c :  2, 2  d  d 
p c

a, i  p  c : 3, 1  d  d  .
p c

a, i  p  c : 3, 2  d  d 
p c

a, i  p  c : 3, 3  d  d 
p c

...

(2) The prisoner’s dilemma has solutions with a strategic


decomposition of a, i  p  c :  2, 4  2  2  if we define the
code of conduct such that the two strategies to confess, C1 C2 
are inactive and dependent. Describe the effects that will be new and
distinct for such solutions. Show that in the co-moving frame there
will now be a non-zero steady state magnetic field.
(3) In decision process theory, to each player or agent there is allocated
one inactive strategy. It thus makes sense that such strategies are not
dynamically changed, but have values that are conserved. We say a
player is an inactive player if every active strategy it is accountable
for is treated inactive by all players in the decision process. This is
an extreme example of an isometry of the solution. Discuss the
equivalent decision process that has one fewer players. In particular
discuss the effects the inactive player still has on the process.
Robinson Crusoe Economics 405

Figure 10.25. Distance time contours for   2

(4) Imagine a political game in which there are three “players”: there is
a player that chooses the political party (say A or B); there is the
player that chooses a financial payoff to the party in power (say high
or low); and there is the player that chooses the popular vote (say
for A or for B). Assuming the game has been defined already in
normal form, discuss the attributes of the game and suggest one or
more possible codes of conduct and the form the equivalent game
might take.
(5) With the numerical calculation based on Sec. 10.8, describe the
effects seen in the holonomic coordinates, Fig. 10.25.
(6) How do the effects seen in Fig. 10.25 compare to those in the
prisoner’s dilemma, Sec. 9.4?
406 Geometry, Language and Strategy—Vol. 2

Figure 10.26. Distance time contours for   20

(7) With the numerical calculation based on Sec. 10.12, describe the
effects seen in the holonomic coordinates, Fig. 10.26, the case
corresponding to high player bonding.
(8) We have computed the time bonding, Fig. 10.28, using the base
model of Sec. 10.7 with a lower frequency of   3 . Discuss its
similarities to the model in Fig. 10.21.

Figure 10.27. Time bonding field with


low frequency   3
Chapter 12

Decisions and Determinism

In the last chapter we focused on the historical treatment of Game


Theory. Some issues arose that relate to the distinctions we believe are
important in a dynamic treatment: in particular we identified codes of
conduct and the tragedy of the commons as attributes we want to
incorporate in a new way.
There are other significant aspects that we approach in a way that is
different from current literature: the way we deal with uncertainty and
determinism in decision making. These aspects of game theory are
associated [Aumann, 1989] with the period 1910-1930. Games as used
today are described in two forms: extensive and intensive.
The extensive form provides the common sense description of
recreational games and realistic decision processes. For example, we
understand chess as a recreational game, which is played by two people
who take turns making their moves according to agreed rules. The game
ends after few moves or many depending on the relative skill of the
players. At the end of the game each player receives (or pays) a payoff.
An important contribution of Von Neumann & Morgenstern, [1944] is
their framing of general transactions in the economic world as being
mathematically the same as an extensive recreational game.
A necessary assumption is that in the economic world, there is a
utility function that allows us to assign a payoff to the outcome, just as in
recreational games. In foundational game theory, Sec. 1.4, we implicitly
adopt this view for decision processes. It is not an explicit adoption
because it is difficult though not impossible to frame the theory in the
extensive form in such a way that it covers all conceivable games.

429
430 Geometry, Language and Strategy—Vol. 2

It is more convenient to frame the strategic content of the theory in


the intensive form [Von Neumann & Morgenstern, 1944], also called the
strategic form [Aumann, 1989]. Using chess again as an example, to
most effectively compete, each player could lay out every conceivable
scenario on a decision tree, each pure strategy, based on possible moves
the two players might make over the duration of the game. For chess the
combinatorial possibilities are staggering but finite. Each player would
then characterize their strategic choices quite simply: they each choose
one of their pure strategies.
The game in this intensive form consists of a play, which is a single
choice by each player and the payoff is based on the utility or outcome
assigned to each player. Obviously, each play may consist of a large
number of moves. The advantage of this view is that we can treat every
game in the same way, distinguishing games only by the number of
players and the number of pure strategies associated to each player. The
intensive form captures all of the essential strategic details of the game or
decision process.
Two games with the same intensive form should behave equivalently.
In the foundational game theory, Sec. 1.4, we have adopted explicitly the
view, which we also call the normal form, that decision processes can
always be described in this way. So far we align with the work from this
early period 1910-1930.
We diverge from the historical treatment in the way we treat
uncertainty and the dynamics. There is a difference in philosophy behind
uncertainty and there is a difference in how we view the mechanism of
determinism: we view mixed strategies as frequencies, not probabilities
and we view determinism the result of mechanistic laws not Bayesian
probabilities.
We start with the early period in which the foundational aspects of
Game Theory are established. We provide our view of uncertainty in
Sec. 12.1, which we think aligns perfectly with the historical view,
though not the current views. We review the corresponding utility theory
in Sec. 12.2.
We articulate our view of dynamics in the next few sections. In Sec.
12.3, we propose the dynamic law of competition that is satisfied by the
decision process theory extension to game theory. In Sec. 12.4, we
Decisions and Determinism 431

propose that the active players in the decision process cooperate through
a dynamic law of cooperation. Such cooperation is by the active players
and involves the active strategies. In Sec. 12.5, we propose a third
important law, the dynamic law of organization. This law deals with the
societal or organizational setting, which takes into account the
constraints of the context in which competition and cooperation occur.
With these laws, dynamic competition, code of conduct (Sec. 11.2),
and dynamic cooperation (Sec. 12.4) provide the dual notions of non-
cooperative and cooperative games in decision process theory.
Everything hinges on how we deal with uncertainty, which we turn to
now.

12.1 Dealing with Uncertainty

In the early period 1910-1930, insight was obtained on different types of


uncertainty, an insight facilitated by a focus on the intensive form of
games. Uncertainty enters game theory in a couple of ways. For
recreational games, such as card games, there is a well-known element of
chance that is separate from any uncertainty the players may have in
making their pure strategy choice. This type of uncertainty can be
absorbed into the definition of the pure strategies. We consider as part of
the moves, each possible random choice and redefine the pure strategies
accordingly. Thus from a theoretical view, there are no new issues on
how to treat such games.

12.1.1 Strictly Determined Processes

A second uncertainty requires a fundamental change to a theory that


deals only with pure strategies. It was observed by Von Neumann &
Morgenstern [1944] that the outcome of certain games (e.g. Ex. 1) is
determined by an optimal choice of pure strategies whereas the outcome
for others (e.g. Exs. 3-9) is not. When the outcome is determined, every
player can analyze each of his pure strategies based on whatever the
other players might do and identify the worst outcomes, which
corresponds to the minimum payoffs for each of his pure strategies. The
432 Geometry, Language and Strategy—Vol. 2

pure strategy that has the largest payoff would then be the best choice
from this defensive stance, called the max-min solution. In the chess
game example, the max-min solution for each player has a payoff. If the
two payoffs are the same, the game is strictly determined.
This is the max-min theorem and holds for all strictly determined
games. For strictly determined games, the outcome or payoff is the same
for every player and each player plays one of their pure strategies.
Furthermore, there is no other pure strategy that will give a better
outcome irrespective of what the other players might do. In general,
games are not strictly determined, so these general games need to be
dealt with separately.
For games that are not strictly determined, the max-min solution may
not be optimal. Though the intensive form exposes only the strategic
properties of the game, it fails to take into account an important aspect of
uncertainty for games and decisions. We know most about games if we
play them repetitively. As in scientific inquiry, we have a hard time
creating a theory for phenomena that occur once. The exception is for
strictly determined games, since it makes little difference if a game is
played once or multiple times. The max-min solution is still the optimal
solution.

12.1.2 Repetitive Games and Mixed Strategies

For any repetitive game that is not strictly determined, any player that
consistently plays a single strategy can have their strategy discovered by
the other players. Other players may cease acting defensively in hopes of
improving their outcomes. Once a player discovers that their defensive
strategy choice starts leading to lower payoffs, they will be encouraged
to change to some other pure strategy. This in turn will lead the other
players to modify their choices. We discuss in more detail below whether
a player makes a choice of pure strategies based on a prediction of what
other players might do, or makes their choice based on past history of
what is no longer a good idea.
Either way, each player sees an advantage to diversify their portfolio
of choices and choose plays according to some frequency distribution of
pure strategies, called a mixed strategy. For players that pick mixed
Decisions and Determinism 433

strategies, it can be proved that there is a max-min theorem for two


person games [Von Neumann & Morgenstern, 1944]. The max-min
theorem is generalized for multiple players with idiosyncratic utilities
with an equally effective and equivalent theorem [Nash, 1951]. Nash
equilibrium is a mixed strategy choice for each player that is optimal in
the sense that there is no other choice available to any player that would
be superior. To the extent that every game has Nash equilibrium, the
steady state behavior of games is established. The underlying meaning of
mixed strategies however is still open. We address that next.
We share with Von Neumann & Morgenstern [1944, p. 19] the view
of mixed strategies as a frequency distribution:

“Probability has often been visualized as a subjective concept more or less in the
nature of estimation. Since we propose to use it in constructing an individual, numerical
estimation of utility, the above view of probability would not serve our purpose. The
simplest procedure is therefore to insist upon the alternative, perfectly well founded
interpretation of probability as frequency in long runs. This gives directly the necessary
foothold.”

We thus think of decisions being made based on knowledge of past


events, including past frequency distributions, as opposed to an estimate
of future behavior. It is our view that these frequency distributions
change in time according to deterministic equations based on past
behavior. We don’t hold that such frequency distributions predict the
future using the mathematics of probability theory. In other words, our
deterministic equations are not equivalent to Bayesian probability theory
[Bayes, 1764] in its various forms.

12.1.3 Bayesian View

We stress this point because the direction of game theory subsequent to


Von Neumann & Morgenstern [1944] has been based on probability
arguments as the mechanism for future predictions. It is of course
tempting to predict the future using a minimum of assumptions about
what is known. This is especially true when we know little about the
underlying processes. It provides a strong reason for using probability
434 Geometry, Language and Strategy—Vol. 2

arguments. Decisions are about things we care about. If we can predict


the future then our payoffs will be higher. The question is whether we
use probability or an understanding of decision processes to make that
prediction. Either way we may use statistical arguments.
The estimation process makes a prediction based directly on the
statistical evidence. We have referred to this as the Farmer’s Almanac
approach. In this case, we require little knowledge of the details of past
weather phenomena. In contrast, fluid mechanics models make
predictions based on the assumption of continuity of the physical process
starting from known positions and the observation that these processes
obey laws. We need however vast quantities of data about the past in
order to utilize such models.
For such models, statistical evidence is used along with the null
hypothesis [Fisher, 1925] to look for evidence against such laws. In other
words it is not enough to postulate a theory; we have to demonstrate that
the theory does not fail. When we identify failures we must modify the
theory to correct the fault. The difference of these two approaches was
discussed during the formative stages of the development of statistics
[Fisher, 1925].
We say more about that. Much of current game theory relies on
Bayesian [Bayes, 1764] or inverse probability as the means to predict the
future. The view is that if we observe the occurrence of an event
repeating itself in the same way many times, it will be more likely to do
so in the future. If the occurrence is a set of strategic behaviors for a
game, we may even adopt that behavior as the equilibrium behavior.
However, we adopt the opposite view along with Fisher [1925, p. 9]:

“… For many years, extending over a century and a half, attempts were made to
extend the domain of the idea of probability to the deduction of inferences respecting
populations from assumptions (or observations) respecting samples. Such inferences are
usually distinguished under the heading of Inverse Probability and have at times gained
wide acceptance. This is not the place to enter into the subtleties of a prolonged
controversy; it will be sufficient in this general outline of the scope of Statistical Science
to reaffirm my personal conviction, which I have sustained elsewhere, that the theory of
inverse probability is founded upon an error and must be wholly rejected. Inferences
respecting populations, from which known samples have been drawn, cannot by this
Decisions and Determinism 435

method be expressed in terms of probability, save in those cases in which there is an


observational basis for making exact probability statements in advance about the
population in question.”

In the context of making decisions, we thus argue against a Bayesian


approach. We hold to the notion of frequency (as opposed to lotteries or
probabilities) for mixed strategies and the use of statistics as a necessary
ingredient of the scientific method, but not a predictor for future
behavior.
In decision process theory, we treat the situation the same, whether a
person believes they are predicting the future or acting on information
from the past. We assume that however one arrives at a frequency
distribution for mixed strategies, knowledge of that distribution as a
function of strategic position and past times provides a scientific basis
for future predictions using a calculus based theoretical framework, the
decision process theory. The difference from the early game theory work
is not foundational, yet there are differences with later works. We differ
because we believe the focus for a dynamic description is on the
repetition of the game, which gives the observational data for the
frequency distributions of mixed strategies.

12.1.4 Calculus View

This confusion inherent with mixed strategies is closely tied to the


conversation about stable solutions. We have characterized this
confusion as two schools of thought: one that frames probability theory
based on set theory, combinatorial methods and Bayesian probability to
predict the future; the second, counterpoised to this, frames decision
processes based on underlying processes and utilizes differential
equations and mathematical tools from the physical sciences. There is
evidence to support belief in the former approach [Von Neumann &
Morgenstern, 1944, p. 6]:

“The importance of the social phenomena, the wealth and multiplicity of their
manifestations and the complexity of their structure, are at least equal to those of physics.
It is therefore to be expected—or feared—that mathematical discoveries of a stature
436 Geometry, Language and Strategy—Vol. 2

comparable to that of calculus will be needed in order to produce decisive success in this
field. (Incidentally, it is in this spirit that our present efforts must be discounted.) A
fortiori it is unlikely that a mere repetition of the tricks which served us so well in
physics will do for the social phenomena too. The probability is very slim indeed, since it
will be shown that we encounter in our discussions some mathematical problems which
are quite different from those which occur in physical science.
“These observations should be remembered in connection with the current
overemphasis on the use of calculus, differential equations, etc., as the main tools of
mathematical economics.”

In support of their view, we see value in the axiomatic tools that have
been applied that have provided much needed analysis and clarity of
distinctions about what constitutes decisions in general and economic
decisions in particular.
We believe the pendulum of this discussion has perhaps swung too
far. We are not in agreement that “tools of the physical sciences are
inappropriate”. We believe a number of advances have been made since
that statement was written, as well as a few setbacks. The particular
complexity, namely the existence of a stable standard of behavior,
suggested by Von Neumann & Morgenstern [1944] does not always exist
[Aumann, 1989, p. 13]:

“Von Neumann and Morgenstern were thus led to the following definition: A set K
of imputations is called stable if it is the set of all imputations not dominated by any
element K :
“This definition guarantees neither existence nor uniqueness. On the face of it, a
game may have many stable sets, or it may have none. Most games do, in fact, have
many stable sets, but the problem of existence was open for many years. It was solved by
Lucas (1969), who constructed a ten-person TU coalitional game without any stable set.
Later, Lucas and Rabie (1982) constructed a fourteen-person coalitional game without
any stable set and with an empty core to boot.”

Moreover, there is new understanding that shows that the differential


equations of physics allow for vastly more complexity [Thom, 1975]
than was previously believed, complexity of the type that might in fact
come out of the set theory and combinatorics view. See for example
Decisions and Determinism 437

[Chandrasekhar, 1961], which has particular relevance given the


similarity of the underlying mathematics of the decision process theory
to the theory of relativistic charged fluids.
In the connection to relativistic charged fluids, we note similarity to
solutions there and complexity demonstrated numerically in our solutions
to the prisoner’s dilemma such as Fig. 9.15. One of the striking
predictions of game theory that rested entirely on the algebraic and
abstract formulation was the standard of behavior. In Sec. 11.2, we
provided a view of codes of conduct, which we feel is an improved
definition of that concept. This improved definition satisfies both the set
theoretic basis in symmetry and the process basis consistent with a
calculus based approach.
The concept of standards of behavior or codes of conduct, along with
the ability to frame any decision process in intensive form provides the
starting point in decision process theory. We adopt knowledge of the
earliest period of game theory, 1910-1930. Fortunately, many of the
elementary textbooks on game theory address the question of how to
frame games in the intensive form. We benefited greatly from one such,
(Williams, 1966) and suggest the student work through their examples.
They may appear simplistic, but in reality capture some of the essence of
Game Theory. We provide exercises at the end of the chapter, some of
which are from this reference.
In this section we have raised the notion of uncertainty and discussed
our view of probability and statistics. This viewpoint changes how we
view utility, which we turn to in the next section.

12.2 Utility Theory

Utility is the value or worth we give to things we buy or sell. We assign


utility to actions we take towards others and actions others take towards
us. In commerce we believe mechanisms are in place to value goods and
services making it possible for a market to exist. More generally, all
decisions require a notion of value associated with the outcomes of that
decision. To discuss value or utility, there is no loss in generality to
discuss the decision process in normal form. In normal form, each agent
438 Geometry, Language and Strategy—Vol. 2

makes a single pure or mixed strategy choice. The resultant outcome is


that each agent will give or receive something of value dependent upon
the collective choices made (and dependent on when that choice was
made).
From a theoretical standpoint therefore, it is critical that this concept
of value or utility be made precise. The basis of a quantitative theory of
decision processes rests on having a quantitative theory for utility. This
has been done through an axiomatic approach, which provides the
necessary mathematical infrastructure for a precise and usable definition.

12.2.1 Preferences

In the axiomatic approach to game theory, utility is discussed in the


context of how decisions deal with certainty, risk and uncertainty [Luce
& Raiffa, 1957]. A simple decision expresses a preference or choice
between alternatives. If we choose between two distinct alternatives A
and B, then we obtain our choice with certainty. If we choose between
an outcome A and some mixture of B and C, then we obtain an outcome
with risk. If we pick the mixture, we know only the frequency with
which we might get B or C but we don’t know which. If there are other
players or factors that affect our decision, then we obtain an outcome
with uncertainty.
This verbiage suggests that our preference is a prediction of what we
will obtain in the future based on our choice. It includes an ordering
based on what will probably happen. Luce & Raiffa [1957] go on to
discuss utility in terms of lotteries and prizes, a context that is suited to
the modern game theory approach [Gardenfors & Sahlin, 1988] in which
equilibrium behavior follows from the rational behavior of the
participants. This context favors the idea that probabilities are like
lotteries and are estimates for future behavior and potential gain or
prizes. The context makes possible discussing future behaviors in terms
of certainty, risk and uncertainty.
Here we recontextualize utility so that it conforms to decision process
theory. In our view, we don’t envision predicting the future (Sec. 12.1)
based on probability. Nevertheless, we frame our view in a historical
context. Though an early treatment of utility goes as far back as
Decisions and Determinism 439

Bernoulli [1738], the treatment that best meets our needs starts with Von
Neumann & Morgenstern [1944], which is based on frequency of
occurrence rather than probability.
Their idea is that it is self-evident (in the sense of creating axioms)
that between two alternatives, A and B, from past behaviors a decision
maker can choose the one that has more utility. This knowledge is not
enough however to provide a basis for utility in game theory. A decision
maker needs to know more. Between the alternative A and the alternative
of some combination of B and C, again by past behaviors a decision
maker can choose the alternative that has more utility. When the concept
of mixture is made precise, a numerical utility function can be defined
that expresses a utility for any mixture of choices. They adopt the most
conservative approach, which is that the mixture is based on the
frequency of occurrences of choices B and C in the past. This approach is
consistent with the usual application of the scientific method [Fisher,
1925], as it does not invoke Bayesian or inverse probability (Sec. 12.1).
The issue of how we predict the future becomes more complex when
we deal with more complex decisions than picking preferences. Now we
have issues of uncertainty in our outcomes based on decisions made by
others or factors outside of our control. We hold to the view that despite
these complexities, a process view is possible in which the future flows
from past behaviors of all the agents involved in the decision process.
We may be able to calculate that flow but only if we make assumptions
about those underlying processes. Hence, we make no use of probability
as estimation.
We do make use of probability or statistics in the usual way it is
employed with the scientific method to discredit assumptions that appear
extremely unlikely. Thus in our dynamic theory we base our assumptions
of what is known about decisions in the present and past and use decision
process theory to determine future behaviors. We use the scientific
method to modify our theory based on observations and measurements.

12.2.2 Measurement

We also use probability and statistics to measure. For example when we


measure the distance between two points, we do the measurement several
440 Geometry, Language and Strategy—Vol. 2

times and average the values obtained. There are differences each time
based on the uncertainties of the measurement process that are unrelated
to the concept of length. In decision processes, there are uncertainties in
measuring the utilities that are unrelated to the presumed concept of
utility between A and a given combination of B and C.
Thus decision process theory allows us to use past data on
frequencies to make choices. Based on this approach, we expect the
axiomatic proof of utility [Von Neumann & Morgenstern, 1944] to
follow unchanged. As a consequence, we should be able to compare
frequency distributions of each decision maker and allocate a numerical
utility, the payoff, to each of them. In order to assure that we can carry
out this program, we envision that sufficient repetitions of the decision
process can be observed in order to arrive at reasonably accurate
measures. We thus specify what we mean by a point in time: it is an
interval that is large enough to carry out the above program, though not
so long as to open the discussion of what behaviors might be steady
state1.
We believe there are numerous examples of this process used today in
the commercial world, examples that assign time-dependent numerical
utilities and payoffs to processes that are inherently human and
subjective. Consider as an example the production of code by software
developers. The production of code results from a collection of decisions.
If the decisions were correct, the code works; otherwise the code has
parts that don’t work and has defects that will be identified in the
discovery or testing stage.
The percentage of code that works versus code that doesn’t work
represents a frequency distribution that is knowable in principle at the
time the code is produced but in practice is not known until later. One
can nevertheless model the production of code as a quantity of code with
a certain defect fraction. The defect fraction is not an equilibrium
statement. It depends on the skill and experience of the software
developers. Over time and based on feedback during the discovery stage,
the number of defects will go down.

1
Cf. Sec. 3.9, where we discuss an analogy to the breakdown of classical mechanics or
optics when the time or space intervals become too small.
Decisions and Determinism 441

This is an example of a strategic fraction changing in time as well as


a utility that increases over time as the production quality increases. To
improve profit and customer satisfaction, the software development
company provides feedback loops to decrease defects, increase skill
levels and thus lower cost and improve quality. A static view using
frequencies as estimates of future behavior would miss this insight.

12.2.3 Cardinal and Ordinal Utility

We assert that it makes sense to consider the cardinal (we know the
actual value) as well as ordinal (we only know the relative order)
properties of utility based on comparing choices that involve
combinations with frequencies [Von Neumann & Morgenstern, 1944,
p. 617]. For choices that involve certainty, only ordinal utility is needed.
The consideration of mixed strategies however brings us back to a
cardinal view (up to linear transformations).
There appears however, a belief that remains that the cardinal view is
too restrictive and only the ordinal view is needed. It is framed as the
view that only the set theoretic view is needed because the calculus view
lacks the capability to produce complex effects [Luce & Raiffa, 1957,
p. 18]:

“The problem is to find an act satisfying … [the linear programming problem]. It is


clearly a decision-making problem under certainty; however it cannot be handled by the
traditional methods of the calculus. What is known as the theory of convex bodies has
proved crucial.”

We have demonstrated in decision process theory that dynamic stable


points of games will also satisfy the equilibrium condition and hence the
theorems of convex bodies.
We now know that complex algebraic results are not excluded as
attributes of solutions to differential equations. Our adoption of a
differential geometry approach allows the possibility of local behaviors
in which each individual player can adopt their own view of the payoffs
and utilities without making it impossible to compare those utilities and
without assuming their utility measures are the same.
442 Geometry, Language and Strategy—Vol. 2

We thus see no reason to exclude calculus as incompatible with many


of the algebraic results that have been obtained, including those on
utility. We see no reason to expect that calculus is any less rich in
allowing complex structure than algebra. The modern view of calculus is
that it is a rich combination of topology, group theory and algebraic
structure.
This rich structure and its possibilities are amply supported in the
literature. For example, viewed as a gauge theory by Hawking & Ellis
[1973, p. 50], calculus is used to make the case for large scale structures
in space and time that go back to the big bang. Viewed in the
mathematical world as a fiber bundle topological structure with a
connection [Eilenberg & Steenrod, 1952], calculus is put in context of an
algebraic and topological structure. We have provided introductory
lectures on the subject of both these physical and mathematical views,
with a more complete bibliography [Thomas G. H., 1980].

12.2.4 Local Utility

We restate two caveats on utility that relate to the decision process


theory introduction in Chap. 1, which echo and generalize a similar
discussion by Luce & Raiffa, [1957, p. 33]. First, the utility function is
determined up to a linear transformation and is specific to each decision
maker. Second, we make the further assertion that this utility function is
local (the gauge theory property), so it is determined at each point in time
and space. We allow the individual to pick any one of the possible utility
functions: the theory will provide results that are not dependent on their
choice. Furthermore, the theory will allow utilities for different
individuals to be meaningfully compared.
We see utilities as reflecting the underlying measure of distance
between strategic points in space and time. This underlying distance
measurement or metric bonding may vary and evolve over time and
space according to dynamic rules, rules that we have specified in
decision process theory. In this theoretical framework, the metric
determines the payoff tensor, Cf. Sec. 2.11, Eq. (2.47). We relate the
payoff tensor to the utility of Von Neumann & Morgenstern [1944] using
the following argument that helps articulate their general framework.
Decisions and Determinism 443

To set the stage, as an example, we assume a utility measure for some


business transaction. Suppose we have invested in three companies. Our
experience is that our investment in company A yielded $25K, company
B yielded $30K and company C yielded $10K. Based on this experience
we prefer A to C and B to A. Because we are using a numerical measure,
we see that we also prefer A to a 50% investment in B plus a 50%
investment in C. This is because a ½ interest in each of B and C would
have yielded $20K, less than the return from A.
The insight of Von Neumann & Morgenstern [1944] is that this
argument can be turned around. First if we prefer A to B and assign a
numerical utility u  A to A and u  B  to B, then because of our
preference, the numerical utility of A should be larger than the numerical
utility of B: u  A  u  B  . The standard economic argument concludes
only that it is reasonable, therefore, for there to be a utility function u
but it is unique only up to transformations that preserve order (ordinal).
The second and crucial step (cardinal) is the assumption that if we
prefer A to a fractional investment f in B and 1  f in C, then the
numerical utility of A should be larger than the weighted sum of B and C
using the fraction f :

u  A   fu  B   1  f  u  C  . (12.1)

Before drawing their conclusion, we note that the word investment was
used here because of our starting example.
We don’t however require a concept of money, but a concept that we
are willing to split our effort between B and C by some fraction. In the
context of decision process theory, the simplest way to envision splitting
our effort is by our actions. We thus envision that we prefer putting our
effort in A to splitting our effort between B and C based on the fraction
f .
If this is the case, then the numerical utilities will be related in the
above way, Eq. (12.1). These ideas are made mathematically precise by
Von Neumann & Morgenstern [1944], who show that under reasonable
mathematical assumptions, the utility function is determined up to a
linear transformation. In other words if there is any other utility
444 Geometry, Language and Strategy—Vol. 2

function u that also satisfies Eq. (12.1), then there will be numerical
constants a , b that relate the two utility functions for any choice A:

u  A  au  A  b . (12.2)

They thus arrive at the converse of our starting investment example.


From our theoretical view, this is an allowed transformation of the
reference frame: significantly, the equations in the theory are
independent of such frame transformations.

12.2.5 Payoffs

We make connection to our payoff fields as follows. We apply this utility


function to decisions in their normal form. Each agent makes a decision
by choosing a single pure strategy. We make the assumption that each
agent cannot only assign a preference to her pure choice versus any set of
pure choices of the other players, but can assign a choice based on
fractional combinations of choices as defined above.
Making the same assumption as before, it is clear that each agent has
a utility function with which to compare any fractional choice of her
strategies against any possible fractional choices made by each of the
other players. This utility function is determined up to a linear
transformation. It is clearly determined at some specific time and relative
to some specific context of choices that have been made. In other words
it is determined locally.
This utility function provides for each agent, the elements of what we
have called the payoff tensor. The values will be for each pure strategy of
that agent against each pure strategy of each other agent. It is also clear
that we obtain a new feature, an internal preference or internal payoff
for pairs of strategies of the same agent. We see that the payoff tensor
captures and extends the essence of the utility argument. In decision
process theory we in fact don’t claim nor require it to be unique. In the
theory there is the concept that at each moment in time as well as at each
strategic point, the theory is non-unique up to a set of general linear
transformations.
Decisions and Determinism 445

We extend the concept of such transformations to make them local.


We also insist that the theory be causal in the sense of physical theories.
We don’t predict the future but base our calculations on past behaviors
that obey laws dictated by a principle of least action. We return to this
discussion in Sec. 12.5.
Thus from game theory we have obtained significant insight into
payoffs and how agents use them to deal with competition. We take this
up next.

12.3 Dynamic Law of Competition

We view the mechanism of determinism as being the result of certain


laws as opposed to the result of Bayesian probabilities. In this section,
we propose the first of these, the dynamic law of competition that is
satisfied by the decision process theory extension to game theory. In
essence, competition emphasizes the importance of the individual and
their need to defend themselves or “win” in a given situation.
We have gained insights about competition through decision process
theory and the two simplified scenarios on the prisoner’s dilemma
(Chaps. 7-9) and Robinson Crusoe economics (Chap. 10). They show
that the known conditions focus on the game theory steady state
solutions. We learn from the literature on game theory (Chap. 11) that
numerical solutions can always be found for games using linear
programming [Luce & Raiffa, 1957].
Based on these considerations, we assert that there is an economic
equivalence principle between the linear programming solutions and
numerical solutions of decision process theory based on the known
conditions. To justify this, we start with the game theory limit in Sec.
12.3.1. We describe in Sec. 12.3.2 the game theory technique of
numerical solutions using linear programming and relate it to decision
process theory. We show the extension of such solutions to decision
process theory solutions in Sec. 12.3.3. In Sec. 12.3.4 we propose the
dynamic law of competition.
446 Geometry, Language and Strategy—Vol. 2

12.3.1 Game Theory Limit

From our numerical examples of the prisoner’s dilemma, we have


learned the known behaviors, which were based on knowledge of the
flows and payoffs. These behaviors hold along some hypothetical set of
strategic values  0 . More generally, we expect game theory to provide
valuable information on steady state solutions based on its well-
developed literature, Cf. Sec. 13.9.
To be sure, we still need to augment our knowledge from game
theory to provide a complete specification of the known behaviors. For
example, in the prisoner’s dilemma we found that there were additional
strains, Sec. 8.4, as well as additional stresses, Sec. 8.1 that must be
specified. We used this augmented set of known behaviors, along with
the partial differential equations to determine behaviors at other points of
time and space.
Nevertheless, the steady state behaviors provided our starting point.
Despite our need for additional information, a large component of the
known behaviors was based on the game theory competitive behaviors.
The nature of such behaviors depends on the nature of the game. We
summarize some of the salient points from the theory of games.
The easiest steady state behaviors to examine are two-person non-
cooperative games, i.e. competitive games. The simplest examples are
ones in which each player has exactly two pure strategies. These provide
an informative and illustrative starting point, which extends naturally to
non-cooperative M  N two-person games where the two players have
M and N strategies respectively.
The competitive aspect is seen clearly in the two-person zero-sum
2  2 non-cooperative game, which is characterized by the payoff matrix
for player 1; the payoff for player 2 is the negative of this, since we
consider a zero-sum game. The 2  2 sub-matrix Grs for player 1
contains all the information about this game.

G G12 
G   11 . (12.3)
 G21 G22 
Decisions and Determinism 447

The two players choose pure or mixed strategies in an attempt to achieve


the best outcome and prevent the worst consequences.
Two situations can arise: first, there may be one strategy that
dominates, i.e. one strategy is clearly better than the other. Suppose the
numerical values of the payoffs are such that G11  G21 G12  G22  .
There is no advantage for player 1 to ever choose the second strategy
(row 2), since no matter what player 2 does, the first row dominates each
element of row 2. If it also happens that one column dominates for player
2, then the game is said to be strictly determined. The steady state
solution is for each player to pick their dominant strategy.
We see that the definition of steady state behavior is the equality:

max min Grs  min max Grs . (12.4)


r s s r

The most defensive strategy for each player is for them to consider the
worst that can happen and then pick the best of these worst cases.
It is not always possible however to find such a steady state solution
since not all games are strictly determined. For games that are not strictly
determined, it is still possible to find a defensive strategy for two-person
zero-sum non-cooperative games using mixed strategies [Von Neumann
& Morgenstern, 1944]:

max min U r GrsW s  min max U r GrsW s . (12.5)


U W W U
rs rs

Player 1 chooses a normalized mixed strategy

U r : U r  1 ,
r

and player 2 chooses a normalized mixed strategy

W s : W s  1 .
s

In the space of all such normalized mixed strategies, player 1


considers the worst that can happen for each mixed strategy choice that
player 2 makes, and then considers the maximum of all of these. Player 2
makes the corresponding defensive choice. The mathematical theorem is
448 Geometry, Language and Strategy—Vol. 2

that there will always be at least one solution to this problem. The steady
state behaviors for this class of games will be these defensive solutions.
The steady state behaviors provide a rule or association of strategic
choices to the given payoff matrix elements.
For strictly determined games, we can often determine the dominant
strategies by inspection. It is in general more complicated to determine
the mixed strategy steady state behaviors for games that are not strictly
determined. There is one case however that is particularly easy: again it
is the 2  2 game.
We consider such a game in which neither player has a dominant
strategy. We compute the following differences for each row and column
as follows:

G p2  1 p2  2 odds1
p1  1 G11 G12 G22  G21
. (12.6)
p1  2 G21 G22 G11  G12
odds 2 G22  G12 G11  G21 G22  G11  G12  G21

Since neither player has a dominant strategy, either the row and column
elements labeled odds will all be positive or all negative. If they are all
negative, we multiply the elements by minus one. It can be demonstrated
that the steady state behavior Eq. (12.5) occurs when player 1 picks a
mixed strategy based on the odds1 column and player 2 picks the mixed
strategy based on the odds2 row. To get the normalized mixed strategy,
divide the odds by their sum, given in the odds2/odds1 entry.

12.3.2 Linear Programming

The ease with which one can compute the mixed strategies for 2  2
games makes them attractive as examples (Cf. the exercises at the end of
this chapter), but not representative of the richness present in zero-sum
non-cooperative games. By increasing the number of pure strategies for
each player from the 2  2 case, the decision process can be made
significantly more realistic. It also becomes much tougher to find the
steady state solutions by hand or by calculator. In addition to the
Decisions and Determinism 449

numerical complexity, there will be combinatorial complexity since there


may be dominant strategies as well as mixed strategies. The practical
approach to obtaining solutions for such games is linear programming
[Luce & Raiffa, 1957].
We summarize the linear programming method for competitive
games, namely two-person zero-sum non-cooperative games. A
competitive game is described by a sub-matrix Grs , where the first and
second indices span the M  N pure strategies of player 1 and player 2
respectively. We look at the game from the perspective of player 1. He
expects to receive a payoff v or greater by playing a normalized mixed
strategy

U r : U r  1 :
r

U G
r
r
rs v. (12.7)

In game theory, the steady state strategies are unchanged by adding a


constant payoff to every pair of pure strategies, so in that theory there is
no loss in generality to assume that the payoff matrix elements are
positive, which implies that the expected payoff to player 1, the best he
can hope for, is also positive.
In a competitive situation, player 1 wishes the payoff to be as large as
possible. We can divide the inequality Eq. (12.7) by this positive
unknown, redefining the mixed strategy u r  U r v , which no longer is
normalized. The resultant problem is that of finding the minimum of
 u r subject to the constraints:
r

u G
r
r
rs  1. (12.8)

This problem is the linear programming problem whose solution is now


straightforward and part of standard packages of software such as
Mathematica, the one we have used for our calculations.
In like fashion we analyze the game from player 2’s perspective. She
expects to receive a payoff v or more, which translates to an
450 Geometry, Language and Strategy—Vol. 2

expectation that player 1 receives the amount v or less, by playing a


normalized mixed strategy

W s : W s  1 :
s

G W
s
rs
s
v. (12.9)

Again, because the game is competitive, Player 2 wishes this value to be


as small as possible. We can divide the inequality by this positive
unknown, redefining the mixed strategy w s  W s v , which is no longer
normalized.
The resultant problem is that of finding the maximum of  w s
subject to the constraints: s

G
s
rs ws  1 . (12.10)

This problem is the linear programming problem dual to the previous


one. The simultaneous solution of these two linear programming
problems provides the max-min steady state behaviors Eq. (12.5).
The linear programming approach emphasizes the algebraic nature of
the solution. We have a geometric problem in a convex space. This does
not look particularly like a calculus problem we are used to dealing with.
Nevertheless we can put this problem into a form suitable for description
in a differential geometry.
In decision process theory, we used the fact (Sec. 1.7) that every
game can be put into symmetric form [Luce & Raiffa, 1957]. We solve
the two linear programming problems simultaneously using the
symmetric game form. For any game we add a constant to each element
so that the symmetric form elements are all positive.
We use linear programming on the modified symmetrized game
matrix to find the strategies u r w s t  with t being the hedge
strategy. For convenience of notation we write these as u a . Given the
linear programming solution, we revert to the original symmetrized
payoff:
Decisions and Determinism 451

ua  0 ,
u F
a
a
ab  0. (12.11)

The linear programming solution of the game is equivalent to finding the


null vector of the symmetric payoff Fab . The null vector is what we call
the payoff direction, of which there may be one or more. In decision
process theory, we expect steady state behavior if it is identical to the
payoff direction of each player.
The convex set geometry constraint problem is turned into a problem
in differential geometry in which the forces are set by a structure that
generates rotations. We saw that the force generated is a Coriolis force,
one that operates in an unusual way in that the force is at right angles to
the motion, Sec. 1.7 and in particular Eq. (1.13). We get a fixed point
when the forces balance appropriately and so the algebraic result is a
consequence of dynamics. From a deep theoretical perspective, this may
not be surprising since it is known that algebraic relationships capture
topological properties of geometry [Eilenberg & Steenrod, 1952].

12.3.3 Decision Process Theory Solutions

Not all games are limited to two players, not all games are non-
cooperative and not all games are zero sum. These extensions can be
made and we find that the treatment of such games nevertheless follows
the above outline.
We argue as follows. In decision process theory, we ascribe a code of
conduct to cover persistent cooperation, cooperation that survives all
dynamic processes. There is a payoff matrix for each player (including
pseudo-players that owe their existence to the code). As argued in Sec.
12.2, we remove the constraint that the utilities of each player are zero-
sum or transferable. Still, under rather general conditions we expect that
there will be steady state behaviors, Nash equilibrium [Nash, 1951],
which are determined by the payoff directions. Such behaviors have the
attribute that they assume all players behave rationally and assume that
with rational behavior, no player can expect to benefit more than what is
prescribed to the steady state behavior by deviating from that
452 Geometry, Language and Strategy—Vol. 2

prescription. In many cases we obtain such prescriptions by inspection,


which we did for the prisoner’s dilemma, Chap. 7.
In the context of decision process theory, we see that the various
prescriptions for computing (or proving the existence of) the steady state
behaviors can be summarized as stating that for the collection of all the
players’ payoff matrices, there is a rational and compelling rule that
associates a steady state flow. In decision process theory, we specify
such a rule, which is a consequence of the field equations, which
determines the flow given each of the player’s payoffs, Eq. (2.49) in the
absence of inertial forces.
We go further however in this theory. The payoff matrices are not
arbitrary but themselves are subject to rules, Eq. (4.14), a generalization
of Ampère’s law, Eq. (1.24). In electrical engineering, this is the
statement that the incremental sum of the magnetic field around a closed
path must equal the current enclosed. If we imagine that for each player,
the payoff matrix is roughly constant in a tube surrounding the steady
state path, then Ampère’s law implies that the current that generates the
payoff field would spiral around the outside of that tube. If the tube is not
very big, then the average motion of the current and the payoff direction
are the same. We propose that our generalized Ampère’s law Eq. (4.14)
extends and replaces the max-min rule Eq. (12.5). We propose that it also
provide a correspondence to Nash equilibrium for more general games.

12.3.4 Law of Competition

We thus propose a general rule that extends the ideas of competition


based on steady state behaviors. The rule reflects two aspects. First,
competitive behaviors are based on a dynamic law of competition
relating flow vectors to the payoffs, Eq. (4.14):
1
2 g bd  b  g ac jk F k cd    T j a . (12.12)

Though this provides the law of competition, there are some caveats.
First, the law depends on the cooperation potential,  jk , Cf. Sec.
12.4. The interaction is the product of the gradient of the cooperation
potential and the payoff matrix (Cf. Eq. (4.16)).
Decisions and Determinism 453

Second, steady state behaviors must reflect, as their name implies, no


acceleration of the flow, Eq. (12.15). This closed-loop behavior provides
unity between decision process theory view and the game theory view.
The economic equivalence principle is that our view and the max-min
rule are the same for two-person zero-sum non-cooperative games.
We support our proposal with the observation that the conserved
organizational current T j a in Eq. (12.12) is determined by the
organizational system dynamics. In the simple model supported by the
organizational dynamic arguments, this conserved current associated
with player j is the product of the flow and player interest flow, Eq.
(3.43), Tab     p VaVb  pg ab .
Each player is characterized by a different interest flow density (Cf.
producers and consumers, Sec. 12.4) given by the flow component V j .
The player interest flow determines whether the flow spirals clockwise or
counterclockwise around the (null vector of the) payoff. The orientation
of the spiral indicates whether the player is a giver or taker in the
decision process.
The simplest case is that interest flow densities are all equal and there
is a single null vector common to all players. In this case we recover the
case above in which the flow of the interest produces a current that
generates each of the player payoffs. We see other options however. The
interest flow densities need not be large at the same points. Players could
in fact pursue different strategies that are nevertheless optimal. In other
words, the payoff direction for one player need not be the same as
another. So the dynamic behaviors might in fact be different tubes
corresponding to each player that might propagate in isolation, interact or
scatter, and then propagate again in isolation as a complicated
topological shape. This is steady state behavior only if both Eqs. (12.12)
and (12.15) are satisfied.
Closely related to the dynamic law of competition is therefore the
dynamic law of cooperation, which we deal with next.
454 Geometry, Language and Strategy—Vol. 2

12.4 Dynamic Law of Cooperation

To initiate our discussion, we follow Luce & Raiffa [1957, p. 114],


starting with their list of assumptions normally made about cooperative
decision processes:

(1) “All preplay messages formulated by one player are transmitted without distortion
to the other player.
(2) “All agreements are binding, and they are enforceable by the rules of the game.
(3) “A player’s evaluations of the outcomes of the game are not disturbed by these
preplay negotiations.”

The cooperative decision processes are those in which the players make
binding decisions that hold throughout the duration of play, what we
termed a code of conduct in the previous section.
Our starting point is [Luce & Raiffa, 1957, p. 118]:

“The cooperative two-person theory of von Neumann and Morgenstern (1947)


singles out the negotiation set as the ‘cooperative solution’ of the game. In words, the
players act jointly to discard all jointly dominated payoff pairs and all undominated
payoffs which fail to give each of them at least the amount he could be sure of without
cooperating. They have argued that the actual selection of an outcome from the
multiplicity of points in the negotiation set N depends on certain psychological aspects
of the players which are relevant to the bargaining context. They acknowledge that the
actual selection of a point from N is a most intriguing problem, but they contend that
further speculation in this direction is not of a mathematical nature—at least, not with the
present mathematical abstraction.”

We anticipate that cooperation requires bargaining [Harsanyi, 1989] and


arbitration [Luce & Raiffa, 1957].

12.4.1 Game Theory Limit

Their specific example helps solidify our understanding of cooperation


as described in the theory of games. We consider the set of payoffs for
the two players:
Decisions and Determinism 455

 2 1  1 1 
G1    , G2   . (12.13)
 1 1   1 2 

At the outset, we see that the two payoffs are not zero-sum. In Sec. 12.3
we show how to compute the strategies for competitive games. Using
those techniques, we compute the most defensive strategy, which is that
each player assumes no cooperation with the other player.
We assume that each player plays a competitive game assuming his
own payoff matrix. The ideal strategies for player 1 are then to play his
strategies with the odds 2:3. We would say that player 1 is playing
against an imaginary opponent, his view of how player 2 would play.
This imaginary player 2 would play the odds 2:3. The real player 2
however plays the odds based on his view of the decision process and
plays the odds 3:2. His view of the imaginary player 1 is that that player
would play the odds 3:2 as well. The real player and imaginary player
don’t align. If each player only played this defensive strategy, they
would each receive a payoff of 1 5 . We call these strategies the max-min
or defensive solutions.
The players might however alternate between the first and second
choices in synch with each other, so they would receive the payoffs
 2,1 or 1,2  . The average payoff 3 2 to each player for this behavior is
much better. To achieve this improved payoff, the players must
cooperate with each other.
The difficult issue is the nature of the cooperation and how to
compute the outcome of that cooperation while still appropriately taking
into account the competitive nature that lurks behind each player’s sense
of utility (Cf. Sec. 12.3). One solution [Von Neumann & Morgenstern,
1944] is to consider the possibilities for the two players viewed in the
two-dimensional space of their utilities. Players will cooperate if they
can achieve more than what they could do based on the max-min
solutions. This defines the sub-region of the utility diagram in which
cooperation can occur. Cooperation can be persistent, in which case we
invoke the code of conduct argument from the previous section.
Alternatively, cooperation can be dynamic.
456 Geometry, Language and Strategy—Vol. 2

12.4.2 Decision Process Theory Solutions

In this case, we find that cooperation does not replace the competitive
nature (Cf. Sec. 12.3) of the decision process but is an additional force.
We propose as fundamental, the law of cooperation, Eq. (4.21), stating
that the cooperation potential  jk between player j and player k is
determined by the payoffs of the two players and the inertial stress
between them:
1
2 g ab a  jl  b lk    T j k  1 4 F j ab Fk ab . (12.14)

For a simple but illustrative model of the stresses, Eq. (3.43), the
organizational system stress between distinct players
T k     p V Vk  p k , is proportional to the product of the player
j j j

interest flows. In words, unless the players share a common ground,


there is no cooperation. The common ground arises in Eq. (12.14)
because their interest flows (charges) overlap, their payoffs overlap or
both.
The cooperation potential depends not only on the overlap but on the
relative sign of the two player interest flows. This is different from game
theory considerations. We frame the sign of the player interest flows into
two distinct economic categories: producers and consumers. A more
prosaic description for any decision process is givers and takers,
respectively. A market choice would be sellers and buyers, respectively.
Our provisional convention is to take consumers with the positive
sign of interest flow and producers with the negative sign. A recreational
game is typically between consumers. In this sense our distinction does
not arise. Moreover, in a two-person decision process, we expect the
payoff matrices for a buyer and seller to subtract. Instead of a zero-sum
game for example, we might get a zero-payoff game as a consequence.
There will be no payoff forces, only cooperative and inertial forces. We
can see these possibilities by looking at the expression for the forces,
which we do next.
There will be different types of cooperation depending on the mix of
givers and takers in the decision process. The cooperation law leads to
the conservation law Eq. (5.57), and with sources qa , Eq. (3.42):
Decisions and Determinism 457

dV b
g ab   abcV bV c  Vk Fabk V b  1 2 V jV k  a jk  qa . (12.15)
d

This provides the closed loop (Cf. Sec. 13.2) result that the time rate of
change of the flow (the acceleration) depends on inertial (or frame)
effects (  abc term), competitive effects (that depends on the payoff) and
cooperative effects (that depend on
the cooperative potential). We discuss
the source contributions in Sec. 12.5.
Assuming only cooperative
effects, steady state behavior occurs
when at a maximum or minimum
potential of the cooperative potential.
Since the product of the interest flows
and cooperation potential involve the
Figure 12.1. Prisoner's dilemma square of each player’s interest flow,
 11   it is reasonable to expect the overall
acceleration or force to be a
maximum at equilibrium. In that case, moving away from the point
would be a restoring (negative) force. This would hold for the
autonomous cooperation potential (the self-cooperation effect, Fig. 12.1)
as well as the cooperation between distinct players.
We might further expect that the
cooperation potential for distinct
players would be a maximum when
the interest flows have the same sign
(both givers and both takers) and a
minimum when they have opposite
signs, Fig. 12.2. Based on the model
from Sec. 8.2, we see from the above
two figures that our expectations are
qualitatively borne out. The model Figure 12.2. Prisoner’s dilemma
also demonstrates that each player  12  
may exhibit both buyer and seller
attributes, depending on the strategic value, Fig. 8.15. The theory
provides a quantitative realization that incorporates all the effects.
458 Geometry, Language and Strategy—Vol. 2

12.4.3 Law of Cooperation

We provide a few practical illustrations of the law of cooperation, Eq.


(12.14). I had a long argument with a former colleague some time ago
about the distinction between consensus and coalitions. These two ideas
often come up in the context of cooperation. He suggested the following
distinctions [Laves, 1994]:

 “Consensus: ‘The process of abandoning all beliefs, principles, values and policies in
search of something in which no one believes, but to which no one objects’. Margaret
Thatcher, 1993.
 “Coalition: The art of enlisting divergent and independent interests to attain a
valuable objective that everyone believes in, if perhaps for somewhat self-interested
reasons.
 “Cognition: The process of educating people in preparation for using another
technique to attain unity of purpose.
 “Coercion: Invoking higher authority to attain a short-term goal. It is most effective
in preparation for coalition building OR when a decisive victory is possible.”

A coalition is the identification of an objective that everyone believes


in. This makes it distinct from the idea of a consensus that possibly
nobody believes in. The dynamic law of cooperation goes beyond the
idea that the outcome is just better for everyone. The proposal makes
clear that a common ground is required and we have provided a precise
definition of what this concept means. This common ground requires
knowledge of each player’s interest flow V j , (a concept analogous to
charge density in physics) as well as their payoffs. If there is no
alignment, there can be no potential for cooperation. It is not hard to
think of examples in city, state or national governments where common
ground is found between divergent groups, Cf. [Ury, 1993].
An example of this is the treaty that was signed between the United
States and the Soviet Union during the Cold War. These countries did
not share the same values, social systems or economic systems. They
each saw however the distinct possibility that their possession of nuclear
weapons could destroy all life on earth. This was their common ground.
As a consequence they were able to cooperate on treaties that reduced the
Decisions and Determinism 459

stockpiles of weapons without changing their ideologies. They did not


cease to be competitors.
A coalition necessarily requires that the interest flows of the parties
align. The consequence of no alignment is well illustrated when national
political parties become so polarized on issues that they become
incapable of being the representatives of the people they were elected to
serve. From decision process theory, we expect that under no
cooperation, the only forces will be those of competition. If there is no
code of conduct other than self-interest, the politics become dominated
by special interest groups. The system looks more like the interaction of
clans, gangs or medieval war lords than a creative cultural, economic and
social environment.
Neither decision process theory nor game theory envisions that the
existence of cooperation turns off the competitive nature of decisions.
What mediates that competitive nature is the establishment of one or
more codes of conduct. For example with the political example
mentioned above, a positive outcome between parties occurs when they
find that within their organization their views are not monolithic. A party
might find that most of its members are conservative on fiscal matters
but not on social issues. A liberal party might find that most of its
members are liberal on social issues but not on fiscal issues. The two
parties might find common ground by being fiscally conservative and
socially liberal. To get to this type of distinction in the modeling we
require the proposed decision process theory. It is a process framework
that reflects how realistic systems behave. It is clear that such a process
description needs to be dynamic since the common ground is usually in
flux.
The law of cooperation in decision process theory requires a common
ground, an overlap between the players’ interest flows or payoffs or both.
Without such common ground there can be no cooperation. Without
cooperation and without a code of conduct we conclude that behavior
reverts to purely competitive self-interest. Thus to establish cooperation
one needs both the establishment of appropriate codes of conduct and
common ground. With cooperation, the possibility exists for the creation
of a robust free market in the sense envisioned by Smith [1776].
460 Geometry, Language and Strategy—Vol. 2

12.5 Dynamic Law of Organization

We have dealt with two aspects of decision making that are covered
extensively in game theory: competition and cooperation. There is a third
aspect however. Though not always recognized, decisions occur in a
context of an organizational system.
Certainly competition and cooperation provide one such context, but
there are others that are like constraints in that they impact the outcome
but we need to know only some aspects of these constraints. If we
included every conceivable player and all conceivable environmental
factors, we would not need constraints. We can’t totally ignore all of the
other players however, any more than we can ignore constraints in
physical processes, Sec. 1.6.
Decisions are made in the context of a family between family
members, in a community, political system or in a society. Decisions
may be made in the context of a business organization. They may be
made in the context of an environmental system. Thus even the physical
environment provides a system of constraints that decisions need to live
within. In each of these cases the system provides an organizing source
that determines the context.
In decision process theory, there are both forces that pull the system
together as well as forces that push the system apart. The inertial
organizing sources generate forces that can do both as well. In physical
systems they are related to the concept that an elastic system that is
deformed returns to the same shape as before, once the deformation
forces are removed. The shape is a type of organization. In a fluid, flow
goes from high pressure to low pressure. The system likes to return to a
state of equal pressure everywhere. This is a specific type of organization
that is characteristic of most organizational systems. They don’t like
changes. They change under stress, but revert to the original form after
the stress is removed. These effects are not dissimilar to cooperative
effects.
In addition, inertial forces may generate vorticity or rotations, such as
Coriolis forces for fluids. In this regard, inertial forces act like payoffs.
When time is an isometry, this is not just an analogy but an exact
statement: there will be a payoff field associated with time as if it were a
Decisions and Determinism 461

code of conduct. We see this in the solutions when there is more than one
active strategy. Thus we believe there are organizational payoffs and
organizational cooperative effects that are equally important in making
decisions as competitive and cooperative payoffs.

12.5.1 Organization and Contextual Frame

If the competitive and cooperative forces are absent, the principle of least
action Eq. (4.17), with  Tabpayoff   Tabinactive  0 , determines the time and
space bonding, the contextual frame potential gab between active
strategies in terms of the inertial organizing source energy momentum
Tab only:

Rab  1 2 g ab R   Tab   Tabpayoff   Tabinactive . (12.16)

We have described models for the inertial organizing sources Tab in


Sec. 3.7. We need to explain in more detail what we mean by the
contextual frame. This will clarify the distinction between the contextual
frame and the organizing sources.
If we are at a point that is at rest relative to the fluid, then no
decisions are being made. If the organizing sources describe a perfect
isotropic fluid, then at this point all directions look alike. We would
expect that any decision that might be made will not favor any particular
choice. This is the ideal situation in which the contextual frame is totally
fair.
If we interpret the metric as this contextual frame, then it will be a
unit matrix. It is clear that fairness in this case is heavily influence by the
fact that we are not making decisions: the decision flow and hence
energy flow is zero. Away from this point, there are many things that
might skew what we consider fair. For example, if decisions are being
made at an accelerating rate, it is plausible that some decisions are now
favored. If the organizing sources are not described by a perfect fluid,
then we may see changes to the contextual frame. Other dynamics may
change the contextual frame, as indeed is predicted by Eq. (12.16).
462 Geometry, Language and Strategy—Vol. 2

12.5.2 Organizational Equilibrium

Thus we have organizational equilibrium condition for decisions that is


distinct from the Nash equilibrium concept based on the dynamic law of
competition, Sec. 12.3. If at every point of space we are at rest and there
are no competitive or cooperative forces, the contextual frame should be
totally fair. It might also be true in the presence of some of these forces
such as in the Fixed Frame Model of Chap. 6. We have a contextual
frame that is isotropic in nature in the holonomic central frame in which
the decision flow is zero at every point. Once we go to the normal frame
in which decision flow occurs, the contextual frame is no longer
isotropic.
However, even for a perfect fluid, the isotropic attributes of an
organizational system at rest lead in a general frame to a contextual
frame that no longer looks fair due to the non-zero energy flow, energy
density and pressure. The fluid tries to be uniform, so high pressure areas
are smoothed out as a consequence of its stresses being isotropic. Thus
high pressure causes flow to go from high pressure to low, as the system
tries to reach an organizational equilibrium. This may not be possible so
the behavior will in general be time dependent: dynamic. However, the
organizational equilibrium of the fluid may manifest as standing and
travelling wave phenomena. It is still there in the co-moving frame, but
not necessarily easily identifiable in some other frame.
The contextual frame is the local perspective in which a decision is
made. The frame reflects the orientation of all of the strategic decisions
at that point of space and time. That contextual frame is a result of the
organizational system payoffs, as well as the competitive and the
cooperative payoffs. The flow is also a result of the combination of these
payoffs. It is the energy flow. Decision process theory provides a
unification of all of these forces into a consistent framework.
Because we envision a theory in which organization, competition and
cooperation play equal roles, we argue that there must be convertibility
between utility associated with each. There must also be a global
perspective, which we turn to in the next chapter.
Decisions and Determinism 463

12.6 Outcomes

From this chapter, the student will have learned the distinctions in game
theory and decision process theory for two important concepts:
uncertainty and determinism. Decision process theory approaches
decisions as based on deterministic mechanisms as opposed to Bayesian
probabilities. Uncertainties exist but are separated between the frequency
with which a choice is made and the uncertainty associated with making
that choice, akin to the separation made in the measurement process in
science and engineering.
The attainment of the outcomes of this chapter is facilitated by doing
the exercises in the following section. Based on this investment, the
student should achieve the more detailed outcomes below based on
section.
 From Sec. 12.1, the student will have learned the advantages and
pitfalls of applying the concepts of probability and uncertainty to
decision processes. The student will have learned that in decision
process theory, each player sees an advantage to diversify their
portfolio of choices and choose plays according to some frequency
distribution of pure strategies, called a mixed strategy. It is not
required that this frequency distribution be a prediction of what is
likely to happen in the future.
 From Sec. 12.2, the student will understand how to identify a different
utility function for each agent. The theory identifies the resultant
payoff fields with the energy of the system. Energy is convertible and
comparable, though utility is not.
 From Sec. 12.3, the student will have learned the relationship between
the max-min rule of game theory and our dynamic law of competition
for obtaining the flow vectors from the payoffs, Eq. (12.12). The
economic equivalence principle is that our view and the max-min
rule are the same for two-person zero-sum non-cooperative games.
 From Sec. 12.4, the student will have learned the proposed
fundamental law of cooperation, Eq. (12.14). Players contribute
differently to this law depending on their interest flows, whether they
are givers or takers. The law of cooperation in decision process
theory requires a common ground, an overlap between the players’
464 Geometry, Language and Strategy—Vol. 2

interest flows or payoffs or both. Without such common ground there


is no cooperation. Without cooperation and without a code of conduct
we conclude that behavior reverts to purely competitive self-interest.
Thus to establish cooperation one needs both the establishment of
appropriate codes of conduct and common ground. With cooperation,
the possibility exists for the creation of a robust free market in the
sense envisioned by Smith [1776].
 This provides in Sec. 12.5, the dynamic law of organization and in
Sec. 13.5, a convertibility mechanism for exchange, which is a new
idea that extends the concept of value. It is not equivalent to the
elementary assumption that all utility functions are the same. It is
based on a specific underlying dynamic process based on the
principle of least action. In a dynamic system, we can convert utility
between players, as well as convert between interest flow, inertia and
opportunity.

12.7 Exercises

(1) Write the TIC-TAC-TOE game in extensive form.


(2) Rewrite the TIC-TAC-TOE game in intensive form. Show that it is a
strictly determined game. What is the Max-Min solution?
(3) A pair of blue bombers is on a mission: one carries the bomb and
the other carries equipment. The Blue/Red payoff matrix is
[Williams, 1966, p. 47]:

 60 100 
 100 80  .
 

Determine the mixed strategies for each. The rows and columns are
labeled: Blue 1=bomb carrier in less-favored position; Blue 2=bomb
carrier in favored position; Red 1=attack on less favored position;
Red 2=attack favored position.
(4) We are told the following river tale by Williams, [1966, p. 50]
“Steve is approached by a stranger who suggests they match coins. Steve says that
it’s too hot for violent exercise. The stranger says, “let’s just lie here and speak the
words ‘heads’ or ‘tails’—and to make it interesting I’ll give you $30 when I call
Decisions and Determinism 465

‘tails’ and you call ‘heads’, and $10 when it’s the other way around. And just to
make it fair—you give me $20 when we match.”
The payoff matrix is

 20 30 
 10 20  ;
 

determine the mixed strategies for Steve and the stranger.


(5) The attack-defense game according to Williams, [1966, p. 51]:
“Blue has two installations. He is capable of successfully defending either of them,
but not both; and Red is capable of attacking either but not both. Further one of the
installations is three times as valuable as the other.”
The attack or defense of the lesser installation is labeled “one,” the
other is “two.” The payoff matrix for blue is

4 1
 3 4 .
 

Determine the mixed strategies for each.


(6) The music hall problem according to Williams [1966, p. 52]:
“Sam and Gena agree to meet outside the Music Hall at about 6 o’clock on a winter
day. If he arrives early and she is late, he will have to drive around the block,
fighting traffic and slush, until she appears. He assigns to this prospect a net worth
of 1 . If she arrives early and he is late, she will get very cold and wet. He
estimates his joy-factor in this case as 3 .”
The payoff matrix is

 0 1
 3 0  .
 

Determine the mixed strategies for each.


(7) The huckster says Williams [1966, p. 56]:
“Merrill has a concession at the Yankee Stadium for the sale of sunglasses and
umbrellas. The business places quite a strain on him, the weather being what it is.
He has observed that he can sell about 500 umbrellas when it rains and about 100
when it shines; and in the latter case he can also dispose of 1000 sunglasses.
466 Geometry, Language and Strategy—Vol. 2

Umbrellas cost him 50 cents and sell for $1; glasses cost 15 cents and sell for 50
cents. He is willing to invest $250 in the project. Everything that isn’t sold is a total
loss (the children play with them).”
The payoff matrix is buying versus selling with positions rain and
shine:

 250 150 
 150 350  .
 

Determine the mixed strategies.


(8) Write the game paper-scissors-stone game in extensive and intensive
form. Show the game consists of two players, each of whom
chooses one of the three choices: paper, scissors or stone. The
payoffs are that scissors cuts paper, paper covers stone, and stone
dulls scissors. This game is not strictly determined. Determine the
max-min mixed strategy solution Cf. Williams [1966, p. 98], who
suggests the following payoff matrix for each player, where the
rows and columns are labeled scissors, paper and stone:

 0 1 1
 1 0 1  .
 
 1 1 0 
 

(9) Determine the optimum strategy for playing the game of Morra
[Williams, 1966, p. 163]. There are two players. Each player holds
up either one, two or three fingers and simultaneously guesses what
the other player will display. If just one person guesses correctly,
the payoff is the number of fingers held up. There are 9 pure
strategies for each player that can be labeled with two numbers, the
number of fingers held up and the guess as to what the other player
will do. The pure strategy 13 is to hold up 1 finger and guess 3.
Show that the steady state behavior involves only three of the pure
strategies, 13, 22 and 31. Show that they should be played in the
ratio 5:4:3.
Chapter 13

Global Systems View

We have noted that the idea of using physics and differential equations
for modeling economic behaviors is quite old [Edgeworth, 1881].
Though this approach has not generated a following and indeed is dated
based on what we now know about the physical and economic worlds, it
contains some important concepts. Physical laws are often determined by
observing a phenomenon in isolation, from which a behavior is extracted
that is true locally.
For example, the world we live in is observationally flat. The
observation is something we really understand and relate to. When we
make local observations about flatness at different points of our world,
the difficult issue is how to express what we have learned in a consistent
way. Just because the world is locally flat does not imply that the world
is globally flat. The insight we obtain from theoretical views such as
differential geometry is a way to stitch local views together. We are thus
led from a local view to a global view of the system we are studying.
This approach is common in the domain of the physical sciences. It
has also been applied with success in the social sciences. The concept of
using differential equations to model social phenomena was proposed by
Forrester [1961] and used by the Club of Rome [Meadows, Meadows,
Randers, & Behrens, 1972, p. 26] to describe the global changes one
might expect given initial conditions on such things as population, birth
rates and resource consumption:

467
468 Geometry, Language and Strategy—Vol. 2

“We too, have used a model. Ours is a formal, written model of the world1. It
constitutes a preliminary attempt to improve our mental models of long term, global
problems by combining the large amount of information that is already in human minds
and in the written records with the new information processing tools that mankind’s
increasing knowledge has produced—the scientific method, systems analysis and the
modern computer.”

The important insight is that differential equations provide the large


scale structure of time and space [Club of Rome, 1972, p. 24]:

“Although the perspectives of the world’s people vary in space and in time, every
human concern falls somewhere in the space-time graph. The majority of the world’s
people are concerned with the matters that affect only family or friends over a short
period of time. Others look further in time or over a larger area—a city or a nation. Only
a very few people have a global perspective that extends into the future.”

The Club of Rome’s insistence on considering both large distances


and large times matches the view of Hawking & Ellis, [1973], who
obtained insight into the large scale structure of space and time at the
cosmological level.
These views have provided both philosophical and mathematical
guidance in creating decision process theory. In this theory we are forced
to consider not only the limits to growth but the limits to choice. This is
necessitated by the continuous nature of time and strategic direction. As
in physical models, the theory closes the loops in time and space,
completing the local observations with general principles that are subject
to experimental verification.
We explore these global aspects in this chapter. We start with what
we mean by the global connections in Sec. 13.1. The most well-known
example of a theory implementing such global connections is Systems
Dynamics, Sec. 13.2. There are also other physically based models, Sec.
13.3. In Sec. 13.4, we reprise three local laws from Chap. 12 that are new

1
“[From the Club of Rome] The prototype model on which we have based our work was designed
by Professor Jay Forrester of the Massachusetts Institute of Technology. A description of that model
has been published in his book, World Dynamics (Cambridge, Mass.: Wright-Allen Press, 1971).”
Global Systems View 469

from decision process theory. From a consideration of the stresses, we


are led to a focus on consequences of this global view, starting with the
convertibility of energy-momentum in Sec. 13.5. We introduce the
concept of opportunity in Sec. 13.6. In the next section we discuss inertia
and player interest, Sec. 13.7, followed by a discussion of opportunity
costs, Sec. 13.8. We finish by closing the loop, Sec. 13.9, and review the
known conditions in the larger context of a systems based theory.

13.1 Global Connections

We see game theory as an accurate, but local view of how strategic


decisions are made over short intervals of time and short distances in
space. Using the language of the Club of Rome, game theory translates
well for family and friends, for cities and for states but not all of these
simultaneously.
Consider the following military example. Traditional armies can
envision battles with large armies and armaments. Armies study how the
last war was fought. They have trouble envisioning new wars that may
be fought under different conditions as exemplified by the urban war
currently going on in Iraq. The standard model of behavior as used in
game theory is not sufficient to help one think out of the box; one needs
courageous generals. Our view of a theory that addresses the large scale
structures of decisions in space and time is that we must change the
standard model to encompass such game-changing strategies, without
relying on the existence of such generals.
Indeed it was our search for a theory that would locally match game
theory yet allow for large scale structures that led us to theories with new
topologies. We needed theories that in some sense were locally flat such
as the shape of the earth, yet would correctly see the global curvature
characteristics. The formal mathematical structures with this property are
known [Steenrod, 1951]. They have been independently created as bases
for physical gauge theories starting with Maxwell and Einstein and used
to understand the large scale structures of space and time [Hawking &
Ellis, 1973]. Such theories combine the power of topology, algebra and
differential geometry. In other words our goal was not to replace what
470 Geometry, Language and Strategy—Vol. 2

has been learned about economic behaviors from game theory but to
extend those theories.
Before addressing the large scale structures that are possible in such
theories and the motivation for expecting such behaviors we obtain from
systems dynamics, we note that game theory has not entirely relied on
the current algebraic approach. There have been studies of continuous
and learning behaviors for repetitive games [Clemhout & Wan, 1989].
There have been approaches that use the physics metaphor to focus on
the stochastic nature of economic processes [Murphy, 1965]. Decision
Analysis also focuses on the stochastic nature [Howard, 1964], according
to the Wikipedia article on this subject [Howard, 2010].
We find these approaches interesting and useful, though we have
reservations towards this approach as expressed in Sec. 12.1 based on the
way they deal with uncertainty. In addition, though some of these
approaches are differential in character, we feel that they don’t address
the global connections suggested by the Club of Rome. In this context,
we feel that decision process theory is most closely aligned with the
approach suggested by Forrester, [1961] based on the causal and
continuous nature of time.
There are many insights to be gained. We start with a summary of the
key elements, not the least of which is the availability of improved
computational techniques. An important element of great practical
advantage is that there are multiple software packages that make
application of systems dynamics no more difficult than applying a
spreadsheet model to an accounting problem. In this regard, we have
found the iThink® software by High Performance Systems [Richmond,
2001] to be helpful.

13.2 Systems Dynamics Modeling

Systems dynamics forecasting requires more than correlations, it requires


causation and closed loop thinking. Causal thinking is typical of physical
models, including those that we may use on a daily basis. As an example,
though early models of weather forecasting relied more on statistical
models than on causal flow, current weather predictions are based
Global Systems View 471

entirely on a causal model of fluid flow [Friedman, 1989]. Systems


dynamics models have been applied to a wide variety of social and
economic problems [Meadows, 2008]. Though there is no reason why it
can’t be applied to decision making, it is not commonly used.
So is this approach at all useful? It might be argued that the maturity
of the field dictates one approach rather than another. So even if decision
making provided an excellent example of a causal forecasting problem,
you could argue that the field is not yet sufficiently mature for these
operational-thinking techniques. On the contrary, we argue with systems
dynamics proponents that system thinking is useful even at the earliest
stages of thinking about any forecasting problem.
One of the problems that system thinking tackles head-on is the need
for and use of differential calculus. Systems dynamics advocates
acknowledge that this is a big stumbling block to adoption of their ideas.
To this end they have come up with a language that helps with the
understanding of the calculus, while maintaining the calculation power.
The basic concepts that underlie systems dynamics and systems
thinking are flows, stocks and closed loops. These concepts provide an
abstraction of the underlying differential equations that are sufficiently
simple to be understood by a wide audience ranging from school children
to CEO’s [Senge, 1990]. In the software, stocks are represented as
rectangles, flows as faucets and the causal nature or cause and effect is
indicated by curves with arrows indicating the flow of time. We identify
closed loops as situations where the water flows in a closed circle.
We give an example in Fig. 13.1. We model predators, such as foxes
and prey, such as rabbits. The number of predators is collected in the
stock or reservoir labeled predator, and the number of prey is collected in
the stock labeled prey. The number of predators increases in proportion
to the number of prey, since they are its food source. Correspondingly,
the number of prey decreases in proportion to the number of predators. In
each case, the numbers of predators and prey increase due to their own
birth rate as set by parameters mu (  ) and lambda (  ), respectively.
472 Geometry, Language and Strategy—Vol. 2

Figure 13.1. Systems dynamic predator-prey model

The prey, such as rabbits might also increase because of their food
source (carrots). The bi-directional character of growth is represented by
flows that have arrows that point in both directions. We represent the
continuous nature of time and approximate the discrete nature of creation
or demise of predators and prey by a continuous amount in their
reservoir. The picture we create in terms of flow results in a quantitative
description of the populations Fig. 13.2, which in this case has unusual
structure in no small part due to the additional food source of carrots.
Global Systems View 473

Figure 13.2. Predator prey populations.

What is significant about the systems dynamics approach is not its


application to the simple problem above, whose simplicity was there in
part to provide a convenient way to explain the concepts. What is
significant is that it forces us to think beyond the initial problem, to all
effects that need to be included. It forces us to close the loop. For
example, in addition to considering foxes and rabbits, we are forced to
consider carrots. Without food, the rabbits will die and so then will the
foxes. These examples of global connectivity and causal connectivity are
the key aspects of the Club of Rome’s study of the world’s problems.
Events big and small are connected in both time and space.
The stocks and flows represent both these causal and global
connections. They can be applied to a variety of concepts from physical
aspects such as the number of predators and prey, the carrot crop as well
as to more psychological or social aspects. For example, in applying this
approach to a business, one can model not just productivity and staffing
but the burnout rate and thereby learn that burnout can play an important
role in understanding delivery times of products.
The Club of Rome dealt with many issues that have a social character,
such as the harm done to the environment by the burning of fossil fuels.
Certainly the amount of carbon dioxide is a physical attribute; what kind
of attribute is the harm to the environment?
We see clear advantages to the systems approach over statistical or
stochastic approaches. To quote from Richmond [2001, p. 19]:
474 Geometry, Language and Strategy—Vol. 2

…In short, Operational Thinking is a big deal! It’s a big deal because, like Closed-
Loop Thinking, it has to do with how you structure the relationships between the
elements you include in your mental models. Specifically, Operational Thinking says that
neither “correlation,” nor “impact,” nor “influence” is good enough for describing how
things are related. Only causation will do!”

This operational approach has been applied successfully to real-world


situations and the approach produces actionable insight. Real examples
involve not a few equations as in the predator-prey example, but
hundreds or even thousands of coupled differential equations. These can
be dealt with by modern computing techniques.
Not only is our goal to use systems thinking but extend it to the where
we consider not just causation and limits to growth, but connections and
limits to choice. Our goal is also to provide an understanding or meaning
behind the equations that result from systems thinking, which is
facilitated by basing our approach on differential geometry.

13.3 Physically Based Models

The motivation for moving to a physically based model, a model based


on a least action principle, is to obtain conceptual understanding with
systems thinking. The idea is to get a better handle on the assumptions
for the causal and global connections. We use the scientific method to
improve on the assumptions. Decision making is complicated and we
expect that the models to reflect that complexity: we find that we will
have lots of equations. The Prisoner’s Dilemma Chaps. 7-9 has hundreds
of equations. However, we believe the concepts are more organized and
easier to evaluate, change and/or extend. The example of making
accurate weather predictions is an excellent example of both attributes:
the fluid flow model involves a large number of equations yet the
concepts are orderly and the causal and global connections clear.
In decision process theory, we have emphasized both the causal and
global connections between what is known and what is predicted. We
have identified the closed loops. They have the same significant effect in
decision process theory as they would be expected to have from an
Global Systems View 475

Operations Systems viewpoint. For example, in decision process theory,


the flow of decisions creates the values of the payoff fields. The values
of the payoff fields dictate the behavior of the flows.
Understanding the consequence of systems dynamics models and
physically based models requires specific computing tools. The original
game theory models could be solved as linear programming problems:
coupled linear (algebraic) equations with constraints. This may have
been challenging in 1950 for corporations such as Rand, but is not as
challenging today. We have already noted that systems dynamics models
require stock-flow software [Richmond, 2001]. The number and
complexity of such equations is challenging for hand-calculations but has
been addressed effectively by commercial software.
Our solutions to the Prisoner’s Dilemma rely heavily on software that
solves coupled ordinary differential equations: we use the symbolic
programming Mathematica software by Wolfram [1992]. In Chaps. 14-
18, we apply more powerful techniques when there are multiple active
strategies. We solve the corresponding coupled partial differential
equations [Courant & Hilbert, 1962], again using Mathematica with the
numerical method of lines.

13.4 The Three Laws—Reprise

We have identified competition, cooperation and organization as three


activities that have their origin in historical discussions and have a place
in decision process theory. They each produce forces that we view from
the perspective of their interactions on flows assuming the sources are
essentially constants. It is by closing the loop that the theory turns these
activities into laws. For example, competitive payoffs that are considered
constant matrices become dynamic quantities.
For competitive processes, the sources that create the payoffs are the
“currents,” T j a , which are the stresses between the inactive player
strategy and the active strategic directions. In Sec. 12.3 we identified the
dynamic law of competition, Eq. (12.12), which is an example of closing
not only the causal loop but the choice loop. We might not come to this
476 Geometry, Language and Strategy—Vol. 2

law if we look purely at competitive behavior where we assume that


payoffs are constants.
For cooperative processes, the sources that create the cooperative
behaviors are the stresses T j k , which are important when there is
common ground between players. In Sec. 12.4 we identified these
sources as generating the player bonding fields, Eq. (12.14). Again we
have an example of closing the loop, which replaces a constant behavior
with a dynamic construct.
The third law, Sec. 12.5, is evoked when there is a need to take into
account the effects of organizations on strategic behaviors. Such
behaviors should be the averages of the competitive effects of the players
not included in the process, and so should mimic payoffs that they would
have. The effects also include the averages of the cooperative effects of
the players not considered. The sources for this include the competitive
and cooperative stresses as well as a new source, Tab , the inertial stresses
between active strategies, Eq. (12.16). Again this would not be
considered at all without closing the loop.
Thus we have isolated three laws that govern decision making based
on sources we collectively call stresses. They generate effects on
quantities we collectively call strains or utilities. We obtain a general rule
found also in physical systems that stresses generate strains. This is
closing the loop, since we started with the relationship that strains
generate stresses.

13.5 Energy-Momentum Convertibility

As an example of the role that stresses play, consider the law of


competition and the stress T j a that plays the role of a current. We can
express this in a way that emphasizes that connection by writing the law
of competition, Eq. (4.16) in terms of the competition potential A j a :
1
2 g bd  b d A j a   T j a  1 2  a g bd  b A j d
 1 2  b g ae g ec g bd F j cd  1 2 g bd  jl  b lk F k ad . (13.1)
Global Systems View 477

The potential is determined by the organizational currents T j a , as well


as other “displacement” currents generated by competition, cooperation
and organization.
These effects are strongly interrelated. The equation for the
competition potential is equivalent to Eq. (12.12). The currents on the
right-hand side depend on several factors. The cooperation potential  jk
is determined by the organizational stress components T jk from Eq.
(12.14). The organizational potentials g ab are determined from the
organizational stress components Tab , Eq. (12.16), as well as the other
stress components. The gradients of these potentials determine
displacement currents, which in turn determine the competition potential
A j a . There is no way to disentangle which current provides the “true”
source for the competition.
There is a similar effect for the other laws. We have three dynamic
laws in which all of the feedback loops ultimately close. The system is
tightly coupled, so the laws of competition, cooperation and organization
are interrelated. These interrelationships reflect the existence of global
connections between events. When there are closed loop behaviors, it is
not always easy to disentangle a single source as the cause of a given
effect. The exception occurs locally when we know all but one of the
effects is small.
We find support for this type of global connectivity from Systems
Dynamics [Forrester, 1961] and their causal and global perspective. In
many practical examples they have been able to identify the connections
and find that in fact certain archetypes are easy to spot (Cf. Ex. 5). Once
found, the actual behaviors may no longer reflect any particular
component.
There is an important consequence of such tight couplings. Since we
have a hard time identifying a particular component, we are forced to
look for attributes that are conserved. In decision process theory, it is the
energy-momentum tensor. The energy and momentum of the system can
flow from one type of interaction to another. Energy and momentum are
therefore convertible and comparable.
This provides the mechanism for exchange which is a new idea and
can be subjected to test. It is not equivalent to the elementary assumption
that all utility functions are the same. It is based on a specific underlying
478 Geometry, Language and Strategy—Vol. 2

dynamic process that follows from the principle of least action. The
conservation of energy and momentum is expressed as the dynamic law
of organization Eq. (4.17), which determines the contextual frame g ab
between active strategies in terms of the total energy and momentum of
the system:

Rab  1 2 g ab R   Tab   Tabpayoff   Tabinactive . (13.2)

We see each of the contributions. It is worth noting that each side is


separately conserved and that the conservation law for the right-hand
side implies Eq. (13.3) below. The left-hand side of Eq. (13.2) is the
opportunity energy momentum tensor, which is a function of the
contextual frame potentials g ab and their gradients to first and second
order.
The right-hand side depends on the stresses. The cooperation energy
momentum tensor Tab inactive , Eq. (4.20) is determined by the cooperation
potential  jk . The competition energy momentum tensor Tab payoff ,
Eq. (4.19) is determined by the payoff matrix. The contextual frame
potentials are determined by all of the sources, including the inertial
organizing sources Tab . The context reflects the fact that decisions take
place in a specific organizational setting, along with competitive and
cooperative forces. These three contributions have indistinguishable
effects on the contextual frame potentials.
Thus the specific organizing source is not important. Their
contributions are convertible. The dynamics of Eq. (13.2) provides the
law of organization that determines the contextual frame potentials and
orientation flux fields at each point of space and time. For example, any
frame effects  abc that contribute to the flow acceleration Eq. (12.15)
will not necessarily have an identifiable source:

dV b
Qa  g ab   abcV bV c  Vk Fabk V b  1 2 V jV k  a jk  qa . (13.3)
d

This is in contrast to the local source terms on the right that are
specifically related to competition, cooperation and organization.
In a more general sense, the frame effects are the consequences of the
topological connectivity assumptions about utilities. We may think of the
Global Systems View 479

contextual frame effects as a narrative about what is happening. The


narrative articulates the frame rotation effects on how the story is
described.

13.6 Opportunity

We see that the stresses play an important role in the theory. The stress
tensor components T have been added to reflect the attributes of the
decision process that are part of the context. There may be millions of
people in the society, but in a specific decision process we are focused on
the few people actually making the decision. We can’t entirely ignore the
million people however. They carry energy and momentum; as sources
they generate the forces for each of the three laws we have just
identified.
Forces in general generate acceleration and for the central frame
active acceleration, Qa in Eqs. (6.78) and (13.3), we associate the idea
of opportunity. The opportunity results from a tension between
alternatives and so depends on directions in space and time. We use the
word tension because we believe it represents the force that moves the
decision makers toward or away from choices based on their current
knowledge of past behaviors. This is the tension that moves the system
towards the future.
Opportunity depends on the competitive force, cooperative force and
the elasticity or connectivity between strategic possibilities, such as the
production and consumption required of a company to produce a product
and make a profit. We can define a distinct opportunity for each: interest
opportunity, cooperative opportunity and organizational opportunity,
respectively.
Decision processes have complex opportunity connections. The
challenge is to capture the essence of such connections to accurately
portray observed effects. We provided a decision process theory model
for the inertial organizing source in Sec. 3.7, a perfect fluid. We have a
different model that seems appropriate in the player fixed frame model,
Sec. 5.4.4. These suggest a range of possibilities.
Using Sec. 3.7 as an example and based on ideas of organizational
dynamics described in Vol. 1, we can choose an ideal elastic situation in
480 Geometry, Language and Strategy—Vol. 2

which the stress tensor is isotropic and characterized by two parameters,


an energy density  and pressure p . The energy density measures the
inertia of the system and the pressure gradient contributes to the
organizational opportunity: the force is proportional to the gradient (rate
of change) of the pressure with respect to position,  a p .
More precisely, we get the source term qa in Eq. (13.3) using Eq.
(3.45) in terms of the energy density and pressure:

b p
q a  h ab . (13.4)
p

We characterize the ratio of energy density to pressure as the resilience


 of the system in the prisoner’s dilemma, Chaps. 7-9, which we argue
is a measure of the system’s elasticity. In some model calculations, the
energy density appears to clump together, Cf. Fig. 8.1. In this case, it
suggests that the pressure gradient is the manifestation of organizational
opportunity and opposes any forces, such as the player interest flow, that
try to pull together the inertial mass.
As an example of organizational opportunity, consider a software
organization with hundreds of people assigned to a particular project2.
With any growth of the project, there will be progressively more
overhead. Overhead includes the increased number of managers required
to oversee the project. The assumption is that these managers perform no
work. They make no widgets or write no code. Because of their addition,
the number of people required grows faster than linear with the size of
the project. This overhead is a manifestation of resilience, Vol. 1; the
larger the resilience of the system  , the larger the overhead.
Ordinarily one thinks of overhead as a waste of resources. We
suggest however that the identification of  is also the measure of the
organizational opportunity. In other words organizational opportunity is
like potential energy. An organization that has zero overhead, though
more efficient, may be incapable of responding quickly to new situations.
An organization with high overhead may have trained more people, since
many may have been idle and managers may have been skillful in

2
See Ex. 4 for a systems dynamics approach.
Global Systems View 481

training. In a crisis, managers can be reassigned to be workers. When


more work shows up unexpectedly, such an organization may be better
able to handle the fluctuations. In this regard, it is an interesting fact that
organizations oscillate between high and low overhead. Perhaps this
reflects the dynamics between having too much opportunity versus not
enough.

13.7 Inertia and Interest Opportunity

Convertibility is closely tied to the game theory concept of Nash


equilibrium. From a mathematical perspective, Nash equilibrium
suggests the existence of fixed points of the dynamic equations.
However, in game theory, we are also talking about the concept of events
not changing; they are steady state. Events are steady state because of
some dynamic, which may be due to a difficulty in making the system
move or accelerate, which we call inertia. We have called this
organizational equilibrium, Sec. 12.5.2 to distinguish this from Nash
equilibrium.
Decisions are processes that will tend to continue along whatever
path they have been set; including staying at rest if that is where they
started. This also implies that things in constant motion remain in
constant motion along their initial direction. As with Newtonian and
post-Newtonian physics, change occurs only when forces are applied.
We thus envision that opportunity, which we define as active
acceleration, provides for the possibility that utility can be converted to
motion and vice versa. This possibility is a part of our dynamically
extended game theory; specifically, the conservation law, Eq. (13.3).
Concepts such as Nash equilibrium predict future behavior based on
the assumption that players should behave rationally. If players have
exhibited a behavior in the past, then they are likely to continue that
behavior in the future. The longer they have exhibited that past behavior,
the more likely they will continue to behave that way [Bayes, 1764].
We take a different view of this tendency. We say that organizational
equilibrium is a real and tangible property of decisions and their
connectivity and creates inertial mass. If the tendency is very
482 Geometry, Language and Strategy—Vol. 2

pronounced then the inertial mass is very large. Future behavior may in
fact remain stable because there are no forces sufficiently strong to move
the system. The inertial mass is not a prediction of the future but a
statement about the reality of the decision process based on present and
past observations. The inertial mass effect can be overcome if other
forces become stronger.
Opportunity also includes our notion of player’s interest flow. In the
central frame, this is the force Vk FakV  in Eq. (13.3); there are no payoff
contributions in the central frame. We associated interest flow with the
possibility of a player being a buyer or a seller. Buying and selling are
social transactions [Mill, 1947]:

“Again, trade is a social act. Whoever undertakes to sell any description of goods to
the public, does what affects the interests of other persons, and of society in general; and
thus his conduct, in principle, comes within the jurisdiction of society: accordingly, it was
once held to be the duty of government, in all cases which were considered of
importance, to fix prices, and regulate the processes of manufacture. But it is now
recognized, though not until after a long struggle, that both the cheapness and the good
quality of commodities are most effectually provided for by leaving the producers and
sellers perfectly free, under the sole check of equal freedom to the buyers for supplying
themselves elsewhere. This is the so-called doctrine of Free Trade, which rests on
grounds different from, though equally solid with, the principle of individual liberty
asserted in this Essay.”

If we don’t sell something but choose to buy, there is an opportunity


cost associated with our choice to not sell in addition to the cost of
buying [Mill, 1848]. The player interest generates the strength of the
payoff field, which contributes to the energy momentum. It is the same
whether buyer or seller. We imagine that if one person chooses to buy,
then to compensate someone else chooses to sell. There is convertibility
between the two.
From our discussion of the law of cooperation (Sec. 12.4), we expect
that buyer and seller contribute with opposite signs to the (potential)
energy, based on a physical and mathematical argument that their
gradients produce the forces that generate motion. For example, when a
company sets up to sell a product, they become both producers and
Global Systems View 483

consumers. Often they have to borrow money to create their business. In


decision process theory they create inertial mass associated with
producing and inertial mass for consuming. We imagine that these
distributions occupy different points in strategic space, as illustrated in
our prisoner’s dilemma example, Fig. 8.15.
If we think of interest flow as analogous to charge, since opposite
charges would normally attract and come together, we think the same
occurs for opposite interest flow. Opposite charges would contribute with
different signs to the potential energy, so we expect the same for
opposite interest flows. From a business perspective it makes sense that
the tension that keeps these interest flows apart is the opportunity for
making a profit. Decision process theory may provide insight that
production and consumption contribute oppositely to the potential energy
of the system. We believe that the concept of convertibility extends to
cover player’s interest flow as well as inertial mass.

13.8 Opportunity Cost

The opposite of opportunity or potential behavior is actual behavior,


which we think of as opportunity cost. As an example of opportunity
cost, we consider the choice of working at a low paying job (production)
versus taking a year off to be trained (consuming) to do a higher paying
job. If we take the year off we lose the wages of the lower paying job for
one year. That is the opportunity cost of taking the training course. At the
end of the year however we hope we will be able to find a higher paying
job. We see that the opportunity cost assumes something about risk and
the future. We certainly take a risk in not working for a year, especially if
we already have the lower paying job. The risk is that we may not get
hired for the higher paying job at the end of the year. The reward
however is that if we do get hired, over the long term we are likely to
receive more total income than we would have received had we stayed in
our current job.
There are other types of opportunity costs not associated with buying
and selling. As an example of such an opportunity, suppose we have a
car with known mechanical problems that will make the car inoperative.
484 Geometry, Language and Strategy—Vol. 2

Do we keep the car or do we junk it? We have researched the cost of the
repair of the mechanical problems and have determined the cost to be
$2K. If the car breaks down tomorrow, we will be out the $2K as well as
having to get the car towed to the junk yard. If we knew for sure the car
would break down tomorrow, we would be ahead by simply driving it to
the junk yard today. However, we don’t know for sure the car will break
down. In fact we have been driving the car for the past year with this
known problem. If the car drives for another year we save the price of
buying a new car. Is the $2K a real cost? What we can say for sure is that
it reflects the opportunity tradeoff of getting rid of the car versus keeping
the car.
In decision process theory, we analyze the problem as follows. We
have data from the mechanic who has serviced our car and many similar
cars with the same problems. Based on what he has told us we have a
current assessment of the utility of the car and whether the car will make
it to the future. This is not the same as knowledge of that future. We
make no assumption that the mechanic is in fact accurate in his
assessment. Based on our current assessment of the car’s utility we
make our decision. If we could iterate the experience, then we would
change our assessment of future utility as we gained or lost trust in our
mechanic. Our mechanic too might improve. What we know for sure
however is the future assessment based on our past experience.
We propose that there is an opportunity utility that expresses the
opinion of future utility for any given strategy, one that is neither
cooperative nor competitive. So, just as payoffs express the view of how
players will respond in a competitive situation, opportunity utility
expresses the view of how the future might unfold. This view is not a
prediction of the future however; it is simply the current view of that
future, just as the payoff matrix is the player’s view or utility measure of
how all the other players will behave. We can compute the utility over
time to estimate whether the current view will hold; even if it doesn’t, we
get a prediction for what the future utility will be.
Since we take utility to be energy, we consider keeping the car or
junking the car to reflect (potential) energy. If the potential energy is
higher to keep the car we hold on to it. If the potential is lower, we
continue to drive it. The potential reflects a reality about our experiences
Global Systems View 485

to date with respect to the car. These experiences are about the past and
present. To the extent that the process captures the essence of the
ongoing decision processes, we do in the end gain knowledge of the
future behaviors, though not as probability predictions but as systems
dynamic predictions based on causal behavior.

13.9 Closing the Loop

By looking at the global properties of the theory, our notion of the


known conditions has changed. We must consider the limits to growth
and the limits to choice. We must satisfy the three laws and so we have
additional boundary conditions to impose.
Thus any student who wishes to apply decision process theory must
do some additional work. That student must satisfy the game theory
requirements: take any economic or decision problem, put it into normal
form, use natural units and extract the payoffs, game values and Nash
equilibriums. The additional work is that the student must supplement
this information and satisfy the full set of the known boundary conditions
so that the full set of equations can be applied.
As an example, for the player fixed frame model, all of the equations
developed in Chap. 6, need to be applied and solved. We recall the main
ideas, which provide a subset of solutions to the field equations:
 Games are specified in the normal-form coordinate basis by
providing the mixed strategy articulated as flows E a for each
(active) pure strategy a .
 For each pair of pure strategies there is an outcome or payoff to player
j , which we label as F j ab .
 We identify the players as corresponding to the inactive strategies.
 We identify true players as those that are accountable for some subset
of active strategies and observer players or observers as those players
who are accountable for no active strategy.
 We identify the code of conduct by stating which of the available
active strategies are restricted to be inactive.
 The payoffs and flows are considered as known behaviors at some
point in time and some initial choice of strategies.
486 Geometry, Language and Strategy—Vol. 2

 At this point, we may have determined only a subset of the known


behaviors. Further research or assumptions may be required to
complete the set. We deal with that later.
 Information in the normal-form coordinate basis is transformed to the
co-moving coordinate basis by means of the steady state coordinate
transformations, E j E j E j  0 , of the inactive strategies and
the dynamic coordinate transformations, E a E a   e E a E a  ,
of the active strategies.
 These transformations obey the equations specified in Table 5.1. As
stated above, we identify the active flow transformation E a with the
known behavior strategy flows.
 The normal-form coordinate basis payoff fields F j ab are assumed to
be specified as part of the known behaviors. These are related to
persistent attributes  f j f j   from Table 5.1, which in turn are
determined from the scalar orientation potentials in the co-moving
coordinate basis.
 We specify the payoff fields as follows: we use the concept of
preferences to identify the spatial components F j mn . These
preferences, including internal preferences are assumed to exist
following the utility arguments in Sec. 12.2. In terms of the assumed
known behavior of the un-normalized flow  a   m m0  , the time
component F j 0m  1 m0 F j mn n is set by the requirement that the
product of the payoff field and flow is zero. The time components are
thus the normalized values of the expected payoffs assuming the
known behavior flow.
 The inertial constant m0 provides a measure for how quickly the
system responds to change. This constant is not determined initially,
though it may be determined from sensitivity analysis.
 There are different types of orientation potentials, which can be
characterized by their rough correspondence to analogous fields in
physical models. One grouping given in Table 5.4 is analogous to
electromagnetic fields. The streamline solutions correspond to the
case that the electric field components  are zero. As a
consequence, the co-moving payoff fields   are steady state. A
second consequence is that these magnetic field components are
determined by the currents analogous to Ampere’s law in physics.
Global Systems View 487

 A second grouping, given in Table 5.5, is related to bond forces. The


symmetric player bonding matrix  components are steady state.
These terms are not as well studied in the physics literature and so no
direct analogies come to mind.
 The third grouping, given in Table 5.6, provides the inertial stress
components. Whereas the orientation potentials describe the rotational
and strain aspects, the stresses describe the forces that act on the
system.
 To the extent that the stresses are steady state, we expect the
acceleration components q to be steady state. This leads to the
player interest components  being steady state as well from Table
5.5.
 It appears at this point that the tidal magnetic components   might
depend on the proper-time. However, we take the point of view for
streamline solutions that these scalar fields as well be steady state.
In writing out these steps, we impose no limitation on the number of
players or the number of strategies available to each player.
We make no assumption about the codes of conduct other than they
must be identified. For each player we assume the existence of at least
one code of conduct associated with that player’s self-interest. We recall
our concept of equivalence that any strategy that is inactive can be
considered as an additional player with no associated strategy. Thus the
code of conduct is an attribute of every game.
We turn to game theory for help in specifying some of the known
behaviors, for articulating the intensive form of the game and for an
understanding of mixed strategies. Nevertheless decision process theory
has distinct differences. The known behaviors, namely the given and
appropriate boundary conditions, are used with partial differential
equations, to obtain well-behaved solutions.
In Chaps. 14-18 we obtain numerical solutions for problems with two
or more active strategies. We use techniques that have been used in
mechanical and electrical engineering on similar problems (Courant &
Hilbert, 1962).
As part of determining known behaviors, it is helpful to think of
steady state situations, though not because they are the common place in
the real world. Rather we study them because from a study of how steady
488 Geometry, Language and Strategy—Vol. 2

state behaviors change when subjected to an impulse, we learn about the


dynamic mechanisms. We thus arrive at a practical, calculus-based
process for analyzing decisions. In Part 3, we apply this process to three
active strategy decision systems.

13.10 Outcomes

In this chapter the student will have learned that as part of closing the
loop, there are three laws that govern aspects of the decision making:
competition, cooperation and organization. The decision making process
is a system in which all these laws operate. In general it is not sufficient
to characterize a decision process as competitive, cooperative or
organizational since all three attributes may be present. It is important to
consider the closed loops of the process, which help identify the global
characteristics such as the limits to growth and the limits to choice. In
Part 3, we will examine in much greater detail the consequences of this
system view for three active strategy systems.
The attainment of the outcomes of this chapter is facilitated by doing
the exercises in the following section. Based on this investment, the
student should achieve the more detailed outcomes below.
 From Sec. 13.1, the student will have learned an appreciation of the
limits to growth and the limits to choice. They provide a global
connectivity that is not visible at the local level. The student will have
learned the importance of closed loops. Closed loops have the same
significant effect in decision process theory as they would be
expected to have from an Operations Systems viewpoint. For
example, in decision process theory, the flow of decisions creates the
values of the payoff fields. The values of the payoff fields dictate the
behavior of the flows.
 From Sec. 13.2, the student will have an exposure to systems
dynamics thinking and their approach to causality using stocks and
flows.
 From Sec. 13.3, the student should understand the basis for the
approach taken in decision process theory.
 From Sec. 13.4, the student gets a reprise of the three laws.
Global Systems View 489

 These laws provide in Sec. 13.5, a convertibility mechanism for


exchange, which is a new idea that extends the concept of value. It is
not equivalent to the elementary assumption that all utility functions
are the same. It is based on a specific underlying dynamic process
based on the principle of least action. In a dynamic system, we can
convert utility between players, as well as convert between interest
flow, inertia and opportunity.
 This leads into a definition of opportunity in Sec. 13.6, which applies
to competition, cooperation and organization.
 From Sec. 13.7, the student should be able to see how Nash
equilibrium is replaced in the theory by the concept of inertia. The
student should note that this is related to the convertibility of interest
opportunity.
 From Sec. 13.8 the student will see an example of a kinetic effect in
the theory that is related to opportunity cost.
 In Sec. 13.9 the student will see that closing the loop means we have
to extend what we mean by the known behaviors, namely the given
and appropriate boundary conditions that are used with the partial
differential equations, to obtain well-behaved solutions. This section
ties the ideas of game theory to the main body of work in the earlier
chapters of the book. It will be used for the applications in Part 3.

13.11 Exercises

(1) In Fig. 13.3, we provide a systems dynamics model using iThink©


Software on the limits to growth. Identify the stocks, the flows and
the closed loops.
(2) In Fig. 13.3, for each rate one can input an equation that represents
the local expectation of what should happen. For example, for the
BirthRate, one would expect it to be the product of the BirthFraction
and the Population. If you ignore the Goal, what would the
DeathRate formula be? What would the net Growth formula be?
Solve the differential equation implied and compute the Population.
490 Geometry, Language and Strategy—Vol. 2

Figure 13.3. Limits to growth

(3) For Fig. 13.3, assume that DeathRate is the product of the
DeathFraction and the square of the Population divided by the Goal.
Show that the growth is limited.
(4) In Fig. 13.4, we provide an example of a typical application of a
Systems Dynamics model. It models the delivery time for a large
software project. It is a global system that is made up of local
component loops, whose validity can be individually verified. For
example the FeatureBacklog is dependent on the rate
RequestForNewFeatures and the DevelopmentRate. Identify the
other local components.
(5) Building solutions to practical problems using Systems Dynamics
from small component parts as in Ex. 4, is facilitated by having well
understood archetypes. Using the World Wide Web, find and
describe each of the most common ones.
(6) Systems thinking is not a set of tools but a language [Richmond,
2001]. The nouns are the stocks, the verbs are the flows. Sentences
are formed by expressing simple processes, such as archetypes.
Paragraphs are formed by using the sentences to describe part of a
system. Global systems consist of many paragraphs. With this in
mind, look again at Fig. 13.4. What story (global system) does it
tell?
Global Systems View 491

Figure 13.4. Software Project.


Chapter 14

Vorticity and Active Strategies

In Part 3, Chaps. 14-19, we focus on numerical solutions to games with


three active strategies. This has applications to games with more than
one person, each of which can have one or more strategies. One of the
significant differences between these games and those discussed earlier
is that the games have non-trivial competitive forces. Such forces can
generate not only oscillations of each strategy in time, but produce
cyclonic shapes or vortices in space. These form natural resonances
behaviors, some of which have discrete harmonic frequencies as
discussed in Vol. 1.
The essential strategic flow behaviors of decision process theory are
harmonic in the normal frame, which we write as follows in terms of the
central holonomic frame variables:

x a  ,   U 0 a    V0 a  
  U a   cos  V a   sin   . (14.1)

The inactive flows V j   are conserved along the streamline. In a


general context, the harmonic structure of the active strategies is a
frequency-decomposition for an arbitrary initial impulse at an initial
time.
We use the gauge freedom to fix the gauge with this harmonic
behavior, which provides unique solutions once we fix the boundary
conditions. Knowledge of the frequency and phase information allows us
to recreate that arbitrary impulse. This is the phasor representation.
Resonant structures are a special case, demonstrating unusually large
responses to a given input.

495
496 Geometry, Language and Strategy—Vol. 2

In Vol. 1, we found that such resonant structures follow directly from


a consideration of the streamline behaviors, making natural and
appealing assumptions about the inactive and active metric values and
the assumption they would not change along the streamline. The discrete
resonant frequencies  d are determined by the average payoffs from Eq.
(1.13), in which the payoffs are the generators of the vortical rotations.
The effects of competition, cooperation and organizational effects (Sec.
13.4) are taken into account as orientation flux field effects
approximated by assumptions on the metric and the composite payoff.
The assumption is that the frame is fixed and the core behavior is free
fall, influenced primarily by the rotational effects generated by the
player payoff fields.
In addition to free fall, other harmonics are possible due to an initial
impulse. We have considered in detail the prisoner’s dilemma and the
Robinson Crusoe scenarios. In each of these cases there is one active
strategy. Time is included by allowing for steady state behaviors based
on Eq. (14.1). We have demonstrated the harmonics for these scenarios.
In Part 3, we further our insight into decision processes by extending
our single active strategy discussions to include any number of active
strategies and any number of players. Based on the insights from the last
chapters, we are now in a position to provide the means to solve such
models numerically and so justify and extend the previous work.
We can obtain steady state solutions by assuming that there is a
holonomic co-moving frame in which time and the inactive strategies are
mutually commuting isometries and the metric components are
independent of their values. This approach extends the player fixed frame
model. We have yet to fully work out the details of that approach so we
restrict attention to the player fixed frame model (Chap. 6).
Our approach in these next chapters is as follows. In this chapter, we
will review what we mean by the term strategy, summarize the equations
and put them in a three dimensional notation and establish a reliable
method for numerical approximations. In Chap. 15, we establish a Game
equivalency principle and explore concepts such as still point and free
fall which extend the concept of Nash equilibrium in three active
dimensions. In Chap. 16, we apply these results to the special but
interesting case of two-person decision processes in which each player
Vorticity and Active Strategies 497

has two strategies. We use as examples, a war game and a social decision
process: the attack-defense model and the work-wealth-wisdom model
respectively. In Chap. 17, we focus on the steady state geometry aspects
of the model. In Chaps. 18 and 19, we address the structural and social
distinctions, respectively, which arise from the model and suggest how
they play out in the real world.

14.1 Introduction

In this chapter we accomplish three tasks. We draw together the concepts


and aspects we will need for a thorough discussion of three strategy
decision processes, including a review of what we mean by strategies and
which strategies might be chosen as active and which inactive. We
summarize the equations from the previous chapters, which fall into the
category of arising from the time dependence aspects and those which
are orthogonal to the time evolution and hence stationary. We choose an
effective method for numerical computation when there are three spatial
dimensions.
Our discussions rest on the harmonic decomposition into phasors,
Eq. (14.1), which is based on Eq. (6.41) from Chap. 6 and is determined
from the spatial components of the boundary condition as well as the
payoffs. We determine both the orientation flux field effects and the free
fall behaviors. These orientation flux field effects reflect the dynamics of
competition, cooperation and organization. They reflect the considerable
freedom we have in setting the initial boundary conditions.
By using phasors, we compute only steady state behaviors. As noted,
in the normal frame, the axes are fixed; we assume that time is an
isometry. The harmonic decomposition with appropriate initial
conditions provides an exact solution of the field equations. Once the
initial conditions are specified on a spacelike surface, we generate all the
fields in the normal form basis.
In this normal form basis, the motion along the streamline   e  q 
will be a harmonic based on the system response harmonics  e q in the
central frame. As in engineering, we study the time evolution behavior
498 Geometry, Language and Strategy—Vol. 2

by looking at the response of the system to the expansion system


response harmonics  e q .
Our approach in this chapter will be first to address the nature of
decision process theory strategies in Sec. 14.2 based on the idea of
preferences. We consider how we might identify certain general
strategies as inactive based on strategic effort in Sec. 14.3. We then
identify the known behaviors, Sec. 14.4 as characterized in Sec. 13.9.
Our approach is to work in the orthonormal frame, which is dealt with in
Sec. 14.5. In the orthonormal frame, we consider the surface geometry
that is orthogonal to the energy flow in Sec. 14.6.
The forces or accelerations in the theory can be split between those
that are vortical forces and those that are gradient forces. The gradient
forces are dealt with in Secs. 14.6 and 14.8, while the vortical forces are
dealt with in Sec. 14.7. The resultant equations for 3  1 active strategies
are put into the form of three dimensional space in Sec. 14.9.
As a first step in solving such equations, we consider lattice solutions
in 2  1 dimensions, Sec. 14.10, which leads to the concept of locked
behaviors in Sec. 14.11 and player bonding effects, in Sec. 14.12. We
choose our numerical approach to 3  1 dimensions as one using
Mathematica and its method of lines in Sec. 14.13. We present resultant
vortex behavior in Sec. 14.14.
Thus in this chapter, we provide a general summary of the theory.
Using examples with two and three active strategies for illustration, we
show exactly how the vorticity composite payoff tensor in general is
determined by free fall behaviors and orientation flux field effects. In the
process we find that a number of distinctions introduced in decision
process theory help us understand the mechanics of decision making.
We show that new and interesting aspects of the theory appear for the
first time when there are three or more active strategies. We expect the
interplay of competitive, cooperative and organizational effects to
generate noticeable effects.
We see that decisions are made by many agents, each of which can
choose from a variety of strategies. It is a question of both the player’s
skill and the player’s efforts or resources to make the most effective
decision. We turn to that issue in the next section.
Vorticity and Active Strategies 499

14.2 Decision Process Theory Strategies

To orient the discussion, recall that for game theory solutions (Sec. 12.3),
strategic results are not dependent on position in strategy space: the
payoff matrix is constant and the metric is unity. In contrast, except at
equilibrium, we expect players to change their mixed strategies with
position and time. Furthermore, we envision that the strategies are not
normalized. We have some guidance from game theory on the dynamics
of such un-normalized mixed strategies. We also have guidance from
decision process theory, Cf. Sec. 2.1.
There is another distinction that we did not emphasize with our
example of the prisoner’s dilemma. At a point in space, we don’t
distinguish between an active strategy and a code of conduct. We can
always treat an inactive strategy or direction as active and leave it to the
dynamics of the theory to show that that direction is conserved. This is
true of the physical world as well: we can choose to ignore the fact that
motion is on a sphere, leaving it to the equations to establish that fact.
The key is that we have a way to move from one description to the other.
So with the prisoner’s dilemma, we look at the situation first from the
perspective that each prisoner has two strategies. We then look at the
situation from a different perspective in which some of those strategies
are in fact part of a code of conduct. In our analysis of strategy in this
chapter we need to keep these perspectives in mind.
In Sec. 1.4, we proposed that strategies are a measure of action: the
effort one applies to a specific choice or a mixture of choices is
determined according to frequency. We expect such a measure to be non-
negative and have suitable units. Thus any specific strategy would satisfy
0  yk   . This differs from game theory because these strategies are
not normalized for each player.
In particular we suggest there are consequences due to the overall
size. The size is measured in terms of a scale set by the unit of
measurement, which we have provisionally identified as the effort. It
makes sense to separate the behaviors due to the size and the behaviors
that are relative to that size. Any ratio of strategies represents the relative
values whereas any strategy or sum of strategies represents the size.
500 Geometry, Language and Strategy—Vol. 2

Since humans respond to actions such as sound in a logarithmic way,


we suggest a convenient method borrowed from engineering that
measures power levels. We use logarithmic units (natural logs however
as opposed to decibels) to define the preference xk  ln yk associated
with each strategy. Preferences take values over the full set of real
numbers   xk   . We used preferences in formulating our
numerical models for the prisoner’s dilemma in Chaps. 7-9. A scale
independent ratio of strategies corresponds to differences s jk  x j  xk
between two such preferences: the player relative preference.
The scale is defined for each agent K by the average of that agent’s
(logarithmic) preference

1
rK   xk  ln K .
n  K  kK

This is the player preference scale. We then measure relative


preferences against the common player effort K with units of effort by
defining yk  K e xk .
We can equally well measure the relative preferences against the
overall decision preference scale as

1
r
n
 xk  ln  ,
where the sum extends over all strategies.
The player preference scale is determined by the player effort and the
decision preference scale is determined by the decision effort
respectively:

rK  ln K   xk  0
kK
. (14.2)
1
r  ln    nK rK   xk  0
n K

The player relative preferences xk as defined are unitless and


independent of the common player effort K . There is a corresponding
Vorticity and Active Strategies 501

decision relative preference x j  rK  r  x j offset from the player


relative preference.

14.3 Strategic Effort

This formulation is different from that used in game theory but related in
that it achieves the similar effect that the normalized mixed strategies
will depend only on the relative preferences, not on the preference
scales. The sum of the un-normalized preferences for each player will be
determined by its player preference scale or the overall decision
preference scale. For decision process theory we argue that the strategy
flows yk correspond to game theory strategies. The flows transform as
vectors so we can equally use the preferences xk . Ratios of such strategy
flows will be differences of preferences and hence independent of scale.
By separating the relative preferences from the preference scale, we
distinguish between those choices that are due to the efforts put into each
strategy (resources) and those choices that are independent of such
efforts (skill).
To elaborate, we think of strategy as measuring the effort put forth for
a pure strategy or mixed set of strategies. The effort is an amount, which
is non-negative and can increase or decrease. Essentially, this effort is a
quantitative measure of the player’s strategic opinion about a given
strategy. As with populations, we think of steady state situations with
either a constant rate of increase or decrease. If the population rate is
proportional to the population itself, the steady state increase or decrease
will lead to either exponential growth or exponential decay. Preference
thus measures the slope of the exponential increase or decrease.
In this case a constant rate corresponds to a constant rate of change of
the preference. In this context, we relate such constant rates in decision
process theory to inactive behaviors. Because this desired behavior of
the preference flow is a constant, we have a useful alternative to the
normalized strategies of game theory. However decision process theory
and game theory are not the same: we see strategic opinion as being
fundamentally dynamic. The concept of constant effort replaces the
502 Geometry, Language and Strategy—Vol. 2

concept of static choice. Some preference flows will thus be inactive and
others active.
By using weighted sums, game theory is consistent with this choice.
However, it does suggest that each of the player preference scales rK be
inactive or minimally that the decision preference scale r be inactive.
This does make some sense in terms of the typical values used for the
payoff matrix. If both time and the decision preference scale were
inactive, we would expect that the sum of the time components of the
payoff matrix would be zero. This is consistent with zero-sum and
constant-sum games. If all player preference scales were part of the
inactive strategies that are not worldview player strategies and so formed
the code of conduct, we would call such decision processes purely skill
driven processes. These would include ordinary games in which the
effort available to each player is set as part of the rules of the game. It
also includes predator and prey situations where the resources are
constrained.
In the prisoner’s dilemma, we started from game theory but then
generated the converse possibility (Chaps. 7-9). We argued that the
prisoners could choose a code of conduct consisting only of the relative
strategic preferences. Furthermore we argued that the sum of the
preference scales of the two prisoners might also be part of that code of
conduct. Either way, this is close to the extreme case in which every
player relative preference s jk is inactive. We call such behaviors purely
resource driven processes. Here, cooperation is more important than
personal gain. Here, players are competing for resources rather than
optimizing for personal gain, so we expect the difference of the player
preference scales to behave dynamically.
If time is also inactive, the player payoff time components F j mt of
each player would be equal. Moreover, since there are no closed loop
flows that are entirely within the player subspace, there are no sources
that generate payoff fields orthogonal to that subspace, Cf. Eq. (2.13) and
Eq. (2.19). In simple language, there are no self-payoffs. We observed
this behavior in our numerical calculations in Chaps. 7-9.
We shall cover both resource driven behavior and skill driven
behaviors. We allow for mixtures of these two extremes and hence
illustrate all of the above possibilities. We provide illustrative numerical
Vorticity and Active Strategies 503

examples (Cf. Sec. 7.6, where for the prisoner’s dilemma, we adopted the
same goals). We start by specifying the system in a state with known or
assumed properties as in Sec. 13.9.

14.4 Known Behaviors

Let’s “start” the system on a spatial boundary that contains a still point,
defined as zero active acceleration, which facilitates our study of
whether it stays at that point or not. We shall refer to the behaviors at our
“starting” point as the known behaviors. These behaviors provide the
boundary conditions for the time independent differential equations of
decision process theory. The equations include the phasor solutions
characterized by a frequency  . We are thus stating the conditions on a
boundary in the space of active strategies.
The general characteristics of a decision process in decision process
theory are contained in Eq. (4.5) and depend on the strategies we identify
as active and inactive in the normal-form coordinate basis:

d 2   jk  d j  Aaj dx a  d k  Abk dx b   gab dx a dx b . (14.3)

The invariant distance  is the proper-time.


Our statement in Sec. 14.2 that we are free to identify some or all of
the inactive strategies as active is based on our ability in differential
geometries to change bases. For example, we can change basis to the
normal-form holonomic basis where we put all strategies on an equal
footing. In this basis, the proper-time is:

d 2  ˆ dx  dx . (14.4)

We transform from one basis to another using Eq. (2.43) and compare
terms.
We can also change basis to one in which we formulate each decision
process identifying inactive strategies as those associated with players as
well as strategies that are inactive based on an agreed code of conduct.
We call this the symmetric normal-form coordinate basis whose form is
identical to Eq. (14.3), but the labels are different. We use the normal-
504 Geometry, Language and Strategy—Vol. 2

form holonomic basis Eq. (14.4) as a convenient common ground to


relate values from one basis to another.
In each application, we specify the known flows, set the values for
the orientation potentials for the symmetric example, transform these
values from the symmetric normal-form coordinate basis to the
symmetric co-moving orthonormal coordinate basis (or to the symmetric
co-moving holonomic basis) and deal with the specification of inertial
effects that have no analogy in game theory. We deal with each of these
topics in the following sections.
In the symmetric normal-form coordinate basis, the initial flow
components V  are specified by the components V a viewed as active on
the initial boundary and the components V j viewed as inactive on the
initial boundary that specify the coupling of the payoffs fields to the
flow. The decomposition between active and inactive depends on what is
known about these flows on the boundary.
The inactive strategies include the player worldview strategies and
the codes of conduct. Along a streamline, these flow components are the
rates of change:

dx 
V  . (14.5)
d

We normalize the flow to unit length, Eq. (7.9) and Eq. (14.4):

dx  dx
ˆ  1. (14.6)
d d

Based on an initial flow direction, there is a natural set of orthonormal


vectors that can be constructed. Initially, this set aligns proper-time along
the flow direction and constructs vectors transverse to the flow using the
Gram-Schmidt orthogonalization process, starting with the initial
strategy directions.
Vorticity and Active Strategies 505

14.5 Constructing an Ortho-Normal Frame

We know the flows. The flows allow us to determine the frame


transformations at the initial point. We sketch this process when there are
three active strategies and three inactive strategies, one of which is a
code of conduct.
The first step is to set the active transformation components
E a  V a and E j  V j to the initial point flows. We then choose the
remaining orthogonal transformation components in the order of the
player world views, the inactive strategy and then the active strategies:

1  2 rc u1 u2 u3  .

The remaining transformations can be viewed as vectors that are


independent of the flow, though not in general transverse to the flow.
We take a unit direction along the first strategy 1 , which defines the
direction U1 . We construct a new orthonormal vector that is both
orthogonal to the flow and of unit length:

U  1  U  1E E  
E 1   . (14.7)
ˆ U  1  U  1E E   U 1  U  1 E E   

We go through each vector and use the same algorithm. Each new vector
must be orthogonal to all previous vectors and have unit length relative
to the initial metric. This process can be done in any frame of reference.
The last strategies will be the active strategies uk . In the symmetric
normal-form coordinate frame, these strategies are orthogonal to each
inactive strategy. In picking the associative orthonormal directions k ,
we set E jk  0 E jk  0 for each inactive strategy j .
Since the initial active strategy flows are known, to complete the
specification of the flow vector, we need the magnitude of the time
component of the flow V 0  V t  m0 and the player inactive strategy
flows. We start the inactive flows at zero. We start with zero player
engagement (charge). This is not unreasonable if we assume that we start
at an equilibrium position qk  0 .
506 Geometry, Language and Strategy—Vol. 2

The inactive flows change with the active strategies as determined


from Eqs. (4.64) and (4.69):

 E j   q E j  2  E j
. (14.8)
 e   q e  2    e

We also provide the equation for the player engagement e . If there are
non-zero player interests  , the player will develop a non-zero
coupling away from equilibrium. Moreover, when there are two or more
active strategies, there can be a non-zero player passion p  , whose
equations are part of the summary in the next section.
Given the initial flow directions and the values of the complete
symmetric co-moving orthonormal set at some initial boundary, we
determine the symmetric co-moving orthonormal frame at all other
positions from the equations in Table 5.1:

 E j    E j . (14.9)

For diagonal solutions to the player bonding, if we start with the inactive
direction j aligned with its co-moving counterpart  j , this alignment
remains at all other values even though the scale is not preserved. These
equations provide coupled differential equations in the active strategy
directions k and can be uniquely solved once the orientation potentials
are determined, which we deal with next.

14.6 Surface Geometry

A common characteristic of differential geometries such as decision


process theory, general relativity and electromagnetic theory is that some
of the equations are independent of time. One can distinguish a time-like
vector such as the energy flow and a space-like hypersurface orthogonal
to the flow on which there is a surface geometry independent of time, Cf.
Sec. 5.2.2. Such equations are characterized as elliptic partial differential
equations and the initial conditions are taken on this hypersurface. It is
then proved that the resultant partial differential equations have stable
solutions at all other points in time.
Vorticity and Active Strategies 507

We made additional assumptions with the player fixed frame model


so that we have more equations defined on this hypersurface, which we
take to be the transverse directions in the co-moving orthonormal
coordinate basis. In this section we summarize the equations that were
derived in Chap. 6, which define the surface geometry. The equations
split into two groups: those that are gradients of the tensor fields and
those that are the “curl” of the tensor fields. Given both sets of equations
for each tensor field, one can show that there exist solutions to the
corresponding partial differential equations.
We start with the symmetric co-moving orthonormal vectors, which
we have dealt with in Sec. 14.5. Next we summarize from Chap. 6, the
gradients of the active and inactive metric components. These metric
components are determined in the co-moving frame by orientation
potentials, whose values need to be determined based on known
behaviors on some initial surface. In this section we detail our
expectations in general, setting the stage for the specific 2  1 and 3  1
dimension models used for our numerical examples. Once we determine
the co-moving frame orientation potentials we then determine the active,
Eq. (4.72) and inactive, Eq. (4.70) metric components.
We summarize the gradient differential equations (Chap. 6) that
determine the inactive components, assuming we know the player
interests  , player engagements e , scalar potentials q for the
acceleration q   q , player bonding potentials  for the player
bonding components    , and player passion p :
        2 q        
        q    q           
 p   p  q   p   p 
 p     p  q p   q  p    p      p
 p  . (14.10)
     q              ph  h 
 n  1 
        p
  q  q q  q       2       p  
 n 1 
          2q  
      q   2                 p
508 Geometry, Language and Strategy—Vol. 2

The form of these equations is based on the player fixed frame model.
The pressure components are:

 
0  h  
 h  1   h  
 p
 n 1 
 
      q   2    1 2  
   1

 1 2     3   3 2     ,


 p     q   q q           
2     2    2    
  p
  p   h  . (14.11)
 n 1 

With two or more active strategies, we have non-zero player passion and
player payoff fields, which were absent in the single strategy models.
In addition to gradient equations, we also have the “curl” equations
given in Tables 5.4 and 5.5, respectively, which were zero in the single
strategy models:

          
         
          0,
     q     2     
    q    2     
    q    2       0. (14.12)

For three or more active strategies, it follows that if the player payoffs
and player interests are initially non-zero, in general the composite
payoff gradients     will also be non-zero. For two or less active
strategies, the above equations are identically satisfied and provide no
additional constraints (Sec. 14.10). They become important only for three
or more active strategies.
Thus the payoffs provide the key indication of vortex type behavior.
The result for three or more active strategies is therefore of particular
interest. The composite payoff   reflects the decision vorticity
Vorticity and Active Strategies 509

behaviors as we move along a streamline solution. For single strategy


solutions this is zero and for two strategy solutions there can be one such
vortex but it is not coupled to the behavior of the players. When there are
three or more strategies however, Eq. (14.12) shows that movement
along a third strategic direction has a gradient    that depends on
q    2     . This contains the sum of the products of the player
interest flow and the payoff field for each player. It says that the resultant
decisions reflect the sum of the overlaps of each player’s interest flow
and payoff (Cf. Eq. (14.30)).

14.7 Geometry of Competitive Field

Given the importance of the player payoffs to generating vorticity, we


note the geometric meaning of the player payoff field. A geometric
interpretation of   can be taken from the discussion at the start of
Sec. 5.1.1, Eq. (5.3). If we move our observation point along the
worldview direction of player  , the orientation potential (player payoff)
  describes the frame rotation of the frame transformation E a :

DF Ea
   e  Ea    Ea    Ea  . (14.13)


For each proper player  , the player payoff field describes the amount
of rotation the player sees as their worldview changes.
In general, we see from the first line of Eq. (14.12) that the player
payoff for one player might influence the payoff of another through the
player bonding. However in the player fixed frame model, the coupling
 can be diagonalized. This means that for certain “ideal” players,
they each determine their payoffs independently. We shall say more
about this in Sec. 14.9.

14.8 Stress and Strain Fields

Stresses represent effects that are organizational constraints for players


not explicitly included in the model, but whose energy and momentum
510 Geometry, Language and Strategy—Vol. 2

must be accounted for. Stresses in general give rise to strains and vice
versa. The bond matrix is an example of strain.
Based on the player fixed frame model, the bond matrix is diagonal.
This is accomplished with diagonal inertial components   Eq. (6.20):

 p        2         


. (14.14)
       0

To complete the specification of the stresses, we specify the energy


density  , the vector viscosity  and the player stresses   .
We choose zero viscosity, zero inactive stress components and linear
relationships between the energy density and average pressure:

  0
   0 . (14.15)
 p

This allows us to solve the equation for the average pressure in


Eq. (14.11):


  h  
 h  1  h    p
 n 1 
 
 q   1 2    1 2  
 1 2     3   3 2      0. (14.16)

We require:

h  
h
  1   h 
 0. (14.17)
n 1

Knowledge of the initial values completes our specification of the


numerical problem for the full set of field equations.
We base our solutions on the Cauchy-Kowalewsky existence theorem
[Courant & Hilbert, 1962, p. 39] for coupled partial differential
equations. We require both gradient and “curl” equations to be known. In
Vorticity and Active Strategies 511

the next section we rewrite these equations in vector notation appropriate


to 3  1 dimensions, where we collect the gradients and “curls” for each
vector field.

14.9 3+1 Active Strategy Equations

The equations summarized in the last section hold for any number of
active or inactive strategies. Since we will spend some time with three
spatial dimensions, it will be convenient to develop the formulas with
that in mind.
Therefore, we focus on three active strategies, na  h  3 . For now
we leave the number of inactive strategies ni  h  variable. Let
n  na  ni be the total active and inactive strategies. The number of
inactive strategies is then h   n  3 . For three active strategies, we use
the notation 1  x, 2  y , 3  z .
For each of the payoff fields we use a notation that is common and
helpful in physics for an antisymmetric tensor with three spatial
dimensions. There are three payoff fields B1  23 for each (proper)
player  , with the cyclic choices of the active indices.
Similarly there are three decision composite payoff fields based on
1  23 and its cyclic permutations. We define 3-vectors as follows
for the player payoff vector Bα and composite payoff ω , along with
relevant scalar fields:


B α  23   3 1
  , 1 2

ω    23   3 1
  ,
1 2

I  2
α 1 22 23 , 
P  p
α 1 p2 p3 , 
q  q1 q2 q3  q, 
Θ   1 2 
3  ,
σ  
αβ 1         ,
2 3
512 Geometry, Language and Strategy—Vol. 2

    ,

    h . (14.18)
ni

We express the field equations in terms of these 3-vectors, which can be


visualized in ordinary space. Note the additional 3-vectors and that σ αβ
is a symmetric tensor in the inactive indices.
Each of the vectors above can be given a specific interpretation in
terms of the decision process. We have an example of systems thinking
in that these separate mechanisms can be individually isolated and tested
yet they arise by closing the loop in space and time, Sec. 13.9. The
process of closing the loop provides limits to growth and limits to choice.
We explore the distinctions that arise based on the field equations.

14.9.1 Pressure

We expect many of the effects to be similar to those found with the


single strategy models. In particular we see no qualitative changes in the
forms for the inertial terms, starting with the average pressure,
Eq. (14.16):

 n    1
  ni   p      q  Θ  1 2  1   Θ  Θ  1 2 σ αβ  σ αβ
 n 1   ni 
 3 4 I α  I  Bα  Bα  3ω  ω .
α (14.19)

We require the denominator to be positive. We look for solutions in


which the average pressure is positive. We have experience in Chaps.
7-9 with all but the last two terms of the right-hand side, demonstrating
that solutions of this type are possible. The new terms are the player
payoff and decision composite payoff contributions: the composite payoff
contribution to the average pressure is positive and the player payoff
contributions are negative.
Vorticity and Active Strategies 513

14.9.2 Acceleration

The acceleration terms are also similar. From Eq. (14.10) we have:

  p
  q  q  q  q  Θ  1 2 I α  I α  2ω  ω      p   . (14.20)
 n 1 

We use the fact that the “curl” of q is zero. These equations differ from
single strategy model calculations because the acceleration is generated
from a potential and the decision composite payoff term has been added,
whose sign is opposite to the pressure and energy density contributions.
This shows that the decision composite payoff reduces the strength of the
inertial term. In the numerical computations, we shall see when this
effect is important.

14.9.3 Player Bonding

A significant feature of the player bonding equations is that the player


bonding is generated from the player bonding potential and the size of
the player bonding is set in part by the energy density. As before,
Eq. (6.25), the player bonding reflects the inertial effects and generates a
“gravity” associated with the energy density and average pressure:

 1 
  Θ    1 2 σαβ  σαβ  1 2 1   Θ  Θ  3 4 Iα  I  Bα  B  3ω  ω .(14.21)
α α

 ni 

We use the fact that the “curl” of Θ is zero. The energy density, player
payoff, player shear and player bonding all contribute an attractive force
for the player bonding potential. The player interest I α and composite
payoff ω contributions are repulsive.

14.9.4 Player Shear Tensor

We can split the bond player bonding matrix into two terms: the above
player bonding scalar and the bond shear tensor, whose equation is
determined directly from Eq. (6.24):
514 Geometry, Language and Strategy—Vol. 2

  σ αβ   q  Θ   σ αβ      1
ni h     . (14.22)

We use the fact that the “curl” of σ αβ is zero. The scalar potential for the
shear has as its source the stresses associated with each player.
For steady state solutions we impose the constraint that the vector
components of shear commute as matrices in the inactive spaces. One
way to achieve that is to transform the known values of the shear vectors
σαβ  0 to a frame in which they become diagonal, σ αβ . In that same
frame, set the initial conditions of the transformed stresses   to be
diagonal. With these initial conditions set, transform all the equations
back to the original frame. Since in the transformed frame, the shear
solution to Eq. (14.22) will remain diagonal, the resultant shear in the
original frame will commute: there is a constant transformation matrix
that diagonalizes the shear at all points. This will be useful in simplifying
some of the equations below.
Note that we have started with a simplified model Eq. (14.15) in
which the stresses are zero. We can still have a non-zero transformation
for the shear components. We are free to consider more complicated
models in which the stresses are not zero. The strategy for solving these
models is the same.

14.9.5 Player Interest

We see qualitatively new behaviors with the equations for the gradient
and “curl” for the player interest Iα :

 n 1 
  I α  4ω  B α  σ αβ  I β   2q  i Θ   Iα
 ni 
. (14.23)
 1 
  I α   q  Θ   I α  σ αβ  I β
 ni 

We see that player interest is enhanced (or decreased) depending on


whether there is overlap or buy-in between the entitlement payoff and the
decision composite payoff. There is engagement: when there is at least
partial buy-in, there is a source for player interest. A simple concept of
Vorticity and Active Strategies 515

equilibrium is that the player payoffs align with the composite payoff.
This generates strong player interest, passion and engagement, where
player passion is the mixed stress components Pα   p  .

14.9.6 Player Payoff

The player payoff behavior is specified as well from Eq. (14.10) and
Eq. (14.12) for the “curl” and gradient:

 n 1 
  Bα   Pα  ω  I α  σ αβ  Bβ   q  i Θ   Bα
 ni  . (14.24)
1
  Bα  Θ  Bα  σ αβ  B  0
β

ni

The player behavior is not constant but is generated by the decisions


made by the player based on the passion, Pα , which determines the
strength a player holds for each decision, as well as on the cross product
ω  Iα of the composite payoff with the player interest.
We have hinted at the idea that the payoffs may change. Here we
have an explicit mechanism. We see that if the player interest is in the
“eye” of the composite payoff, the cross product gives zero. We can
determine the payoffs solely based on the acceleration, player bonding
and player passion.

14.9.7 Player Passion

The gradient and “curl” equations for the player passion follow from
Eq. (6.18):

ni  1
  Pα  q  Pα  Θ  Pα  σαβ  P β
ni
. (14.25)
 1 
  Pα   q  Θ   Pα  σ αβ  P β
 ni 
516 Geometry, Language and Strategy—Vol. 2

This corresponds to the conductivity model with non-zero “curl”, Cf.


Ex. 1. We have a concept of player ownership, and gain the insight of
associating this model with one of high conductivity, one in which
equilibrium establishes itself quickly.
An alternate model, the ownership model, is one in which   Pα  0 .
The surfaces of constant ownership potential P determine the player’s
view of the stresses that lead to the payoff field.

14.9.8 Composite Payoff

The gradient and “curl” equations for the composite payoff are:

  ω  I α  Bα  q  ω
. (14.26)
  ω   Θ  2q   ω

We see that player impact on the composite payoff outcome is the


overlap of the player entitlement payoff and the player interest. If there is
no player impact, the composite payoff is determined by tidal effects that
are independent of any particular player. If there is player impact, then
the composite payoff is not zero.

14.9.9 Stress Components

There are six active stress components in Eq. (14.11), which are
determined from the above equations:

1
 p      q   q  q        
ni
2 B B    2Bα  Bα   2    2ω  ω 
  p
 1 2 I I             p     . (14.27)
 n 1 

For comparison the inactive stress components Eq. (14.14) are:

 p  2Bα  Bβ  1 2 I α  I β    . (14.28)


Vorticity and Active Strategies 517

The first two contributions are the stresses between players and reflect
the overlap of the player payoffs and the overlap of the player interests.
In other words, when players share common ground there will be
stresses.

14.9.10 Inactive Frame Transformations

To complete the specification of the field equations, we need the frame


equations for the inactive components, Table 5.1:

E j   E j e
1
E j  σ α β E j   ΘE j . (14.29)
ni
1
e  I α  qe  Θe  σ α β e
ni

We see that if the pure inactive strategy    j is initially set along the
inactive strategy j , the pure inactive strategy continues to be along that
strategy.
For each player, surfaces of constant player bonding generate the
same frame transformation. We also reconfirm our understanding that the
player interest is directly related to the gradient of the player
engagement for that player. We can use this information to determine the
remaining variables.

14.9.11 Seasonal Vector Potential

The magnetic fields determine the related seasonal vector potential


A  2e q a based on Eq. (6.14), which guarantees that the commutation
rules are satisfied for the gradients and provides the gradient and “curl”
equations for the seasonal potential:
518 Geometry, Language and Strategy—Vol. 2


A  2e q a1 2e q a2 2e q a3 
  2a  q
. (14.30)
f    A  2e q  ω  e Bα 
A  0

In other words, the seasonal payoff is determined by the player payoffs


and the composite payoff.
We have eliminated the gauge ambiguity with the covariant gauge
condition by applying it on the surface z  0 along with an additional
condition (Sec. 15.15, Ex. 7) that ensures that the gauge condition is
satisfied at all values of z . We also have the constraint that on that
surface, we must match the boundary conditions of the vector potential
with those of the composite and player payoffs.
The seasonal potential generates a frame rotation of the active
system. For example, there may be periodic activities that are not
associated with making a strategic choice such as seasonal variations or
quarterly earnings. They play a role in the decision process even though
they are external. The equations for the seasonal potential show what
sources generate the rotations; the rotational energy can be transferred to
or from other fields. Also recall that in other theories based on
differential geometry such as general relativity, there is precedent for
such fields (gravitomagnetism), so we have some justification for
including this and expecting it to play a role.
The equations for the seasonal potential can be written as second
order differential equations using the field equations (see Ex. 13,
Sec. 6.10):

 1 2 e  q 2 A   e Pα  ω   Θ  3q  e Iα 
  2 
 Bα   I α  e q   1   e Θ  2σ αβ e   . (14.31)
  ni  

We see that the player passion Pα contributes to the generation of a


seasonal potential A  2e q a . However, the overall behavior is more
complex as the player payoffs and composite payoff are also coupled. We
Vorticity and Active Strategies 519

strengthen our understanding of player impact, whose source is the


overlap between the engagement and passion of the player.

14.9.12 Proper Time Equations

The seasonal potential for the averaged magnetic fields creates a frame
rotation in the central co-moving holonomic frame and provides the basis
for taking co-moving frame derivatives with respect to the proper active
strategies, Eq. (6.58) expressed using normal partial derivatives (that
mutually commute):

   
. (14.32)
     2a  q   

We showed in Sec. 6.6 that we can use these expressions to determine


the time dependent partial differential field equations Eq. (6.33) with the
covariant gauge condition:

1  e e


 4a  a  4a  q  q  q 2   2 x a    x a
2  2a  q    x a   q  Θ   x a (14.33)
  2Θ  a  6a  q  q  Θ    q  2q  q   x a  0.

We then use the harmonics of Sec. 6.3 to solve these partial differential
equations for the position scalars x a  x, y , z ,  based on the scalar
potentials that are independent of proper-time.
We add to our set of partial differential equations coefficients that
depend only on the proper active strategies. We have demonstrated this
strategy earlier in Chaps. 7-10. This is the logical generalization to the
case of two or more spatial strategies. We call these n  1 dimensional
models, where n is the number of spatial dimensions. These equations
appear to have a dependency on proper time that might be hard to
eliminate. Though we showed this not the case in Chap. 6, a more direct
approach is to go to the central co-moving frame.
520 Geometry, Language and Strategy—Vol. 2

14.9.13 Central Time Equations


We get a more direct view of the phasor harmonics by transforming to
the central co-moving holonomic frame, Sec. 6.6, Eq. (6.79), where
   and central time   e q  :

  x a  1  e e  4a  a  e 2 q   x a  4e q a   x a
  q  Θ   x a  2e q a   q  Θ   x a  0 . (14.34)

We see that the time dependence of the coefficients in Eq. (14.33)


disappear as well as the divergence factor   q . In the central holonomic
frame, time is an isometry (the metric is independent of central time),
Sec. 6.10, Ex. 23. The general solutions are superpositions of the
following zero frequency and non-zero frequency contributions,
Eq. (14.1):

  U 0 a   q  Θ   U 0 a
2 eq a   q  Θ V0 a  4e q a  V0 a 0,
  V0 a   q  Θ   V0 a 0,
  U a   2 e 2 q 1  e e  4a  a U a
  q  Θ    U a
4 e q a  V a  2 e q a   q  Θ V a 0,
  V   e
a 2 2q
1  e e


 4a  a V a

  q  Θ   V a
4 e q a  U a  2 e q a   q  Θ U a  0. (14.35)

The solutions are attenuated frame-waves that propagate from the initial
conditions as a function of central time, Sec. 6.10, Ex. 37. The
attenuation is determined, among other factors, by the acceleration and
player bonding q  Θ . The equations can also be written as a single
equation using complex numbers. The content of this equation is then the
magnitude and phase of the solution, called the phasor.
Vorticity and Active Strategies 521

14.9.14 Waves

We made a distinction in Sec. 6.6 between frame-waves and


gravitational waves that are theoretically possible for differential
geometry theories such as the decision process theory. We expanded this
discussion to travelling waves and standing waves in Secs. 10.10 and
10.11. We concluded that there will always be near zone waves,
associated with the fields from a moving source. There is a requirement
however on the presence of far zone waves, called radiation.
Thus it is possible for there to be radiation. If the necessary
constraints are satisfied, we would associate gravitational waves with the
active coordinate system metric and associate light waves with each
payoff matrix.
In the near zone, we have frame-waves that are associated with waves
in the transformation from the central co-moving basis to the normal-
form coordinate basis. Frame-waves may be attenuated as in
transmission-lines, or may be travelling waves. We look for the
reflection of their wave behavior in various tensors. In steady state, we
will see standing waves that sum both the travelling wave and its
reflections. We understand such behaviors by appealing to the equations
of motion in the normal-form coordinate basis along with the associated
boundary conditions.
For the solutions under consideration here, the metric tensor is
constant in the central holonomic basis and so there is no wave
propagation in that basis. If we transform from that frame to any other
frame in which there is an initial harmonic “twist,” that “twist”
propagates as a wave. We have changed the boundary conditions. Thus
the choice of the harmonic gauge makes it possible to study phasor
solutions in exactly the same way as one might do for electromagnetic
phenomena in electrical engineering. The propagation properties of such
system response harmonics reflect the underlying dynamics.

14.10 2+1 Preview of Lattice Models

Though we have chosen to focus on models with 3 or more spatial


dimensions, there are insights to be gained by focusing first on 2 spatial
522 Geometry, Language and Strategy—Vol. 2

dimensions. This will help build our understanding of the numerical


results and identify their strengths and weaknesses. In 2 spatial
dimensions, we see effects due to an additional dimension without the
interplay of different payoffs and passions.
The simplest calculation that extends zero acceleration is to simplify
the acceleration Eq. (14.20) to the non-linear Laplace equation:

 2 q  q  q
. (14.36)
   x z

We have dropped all of the terms except the quadratic term. From our
single strategy models, we expect that the acceleration flow moves
purely along a single (say the z ) direction. We expect no movement
along the other (say the x ) direction.
If this other direction is active, we can have movement along that
direction. Part of the effect follows directly from the behavior of
Laplace’s equation:

2 q  0 . (14.37)

We expect that flows qx   x q along the x direction, with suitable


boundary conditions, will appear to circulate as in Fig. 14.1. We see
basically the same effect in the non-linear Laplace equation Fig. 14.2,
though the range has decreased substantially in the z direction.
In both cases we demonstrate critical aspects that must be addressed
with solving equations with more than a single active strategy. We must
decide on the boundary conditions. In this case along the surface z  0 ,
we assume that we know the spatial values of the acceleration potential
q0  x   q  x, z  0  and the derivative  z q0  x    z q  x , z  0  . It
follows that we have specified all derivatives of these functions with
respect to x .
Vorticity and Active Strategies 523

Figure 14.1. Acceleration flow in Figure 14.2. Acceleration flow in


Laplace's equation non-linear Laplace's equation

Our practical method of solving the equations will be to consider


these functions on a lattice of points along the x direction:

q  x, z   q  xm , z  : m   N , N   qm 
. (14.38)
qz  x, z    z q  xm , z  : m   N , N    z qm 

Whenever we require derivatives of these functions along x we


approximate the derivatives by these (second order) symmetric lattice
derivatives:

 1 
 x q    qm 1  qm1     x qm 
 2  
. (14.39)
 1 
 x  x q   2  qm 2  qm2  2qm     x  x qm 
 4 

In this expression, the spacing between points along x is  .


One difficulty with this approach is that for a lattice of finite size, the
derivatives at either end are inaccurate unless we know what the function
is outside the lattice. We resolve that difficulty here with the simplifying
assumption that the boundary conditions are periodic. The points outside
the lattice are determined by the values inside. Note that these equations
hold at all values of z so that the 2-dimensional partial differential
524 Geometry, Language and Strategy—Vol. 2

equations are replaced by multiple ordinary differential equations, one


for each lattice point. This approach was used in generating the above
figures. We employ Mathematica [Wolfram, 1992] to carry out the
numerical work.

14.11 Locked Behavior

We are now in a position to comment on the characteristics of the results,


results we could obtain analytically in the case of the Laplace equation.
The characteristics are more general and will be seen to hold. The flow
reflects the periodic boundary conditions. For the Laplace equation, we
have conserved flow whereas for the non-linear Laplace equation, the
acceleration is also a source for the flow. In either case, the possibility of
lateral flow results in a picture reflecting that the flow circulates as well
as the fact that there can be sources and sinks.
We take this to the next level by considering the 2  1 dimensional
results for Eq. (14.10), observing that Eq. (14.12) are identically satisfied
unless there are three or distinct strategies. We employ a notation
modeled after Eq. (14.18):

B   xz
   xz
I α  2x 2z 
Pα   p x p z  . (14.40)
q  qx qz 
Θ   x z 
   0

We make the simplifying assumptions Eq. (14.15), assume that the


shear components are zero and get the equations for the orientation
potentials in Exs. 3-5. We explore using Ex. 6 the behaviors that
generalize the 1  1 model calculations in Sec. 8.1ff, picking the number
of inactive strategies and the initial values to roughly correspond to the
prisoner’s dilemma computation.
Vorticity and Active Strategies 525

Figure 14.3. Pressure in 2  1 Figure 14.4. Acceleration  qz in


dimensions 2  1 dimensions

We reproduce the locked behavior seen in the prisoner’s dilemma in


which there is a high pressure and hence an energy density peak. In
particular we replicate the pressure p , Fig. 14.3 and the acceleration
component qz , Fig. 14.4, assuming the z direction corresponds to the
single strategy active direction (see Ex. 8). We compare these results
with Fig. 8.1 and Fig. 8.2. We have not chosen identical parameters;
nevertheless we see the strong similarity.
In this 2  1 dimensional model, the boundary conditions have been
chosen so that there is no variation along the x direction and so we
approximate the single strategy 1  1 model.
We emphasize that these curves are based on a numerical calculation
on a lattice. This gives us confidence that our lattice results represent the
behaviors of the theory.
We see similarities from our single strategy model calculation. The
pressure for the prisoner’s dilemma model, Fig. 8.1, peaks at the origin
based on an assumption that the stress components are analogous to a
perfect fluid, Eq. (7.46).
The pressure for 2  1 dimensions, Fig. 14.3 is based on the steady
state hypothesis, which insures we have solutions to the field equations.
This leads to the stress components being determined by Eqs. (6.20),
(6.28) and (6.18). Despite slight structural differences in the acceleration,
the qualitative shapes are the same.
526 Geometry, Language and Strategy—Vol. 2

Figure 14.5. Pressure with x structure Figure 14.6. Acceleration  qz with


x structure

We see new behaviors in the pressure, Fig. 14.5 and the acceleration
Fig. 14.6 when we change the boundary conditions to allow structure in
x , which we do in Ex. 9. In this example, we assume that the composite
payoff and player payoffs are small.
The structural changes in the acceleration flow, Fig. 14.7 show the
onset of the lateral flow we identified as possible in the Laplace
equation, Fig. 14.1 and the non-linear Laplace equation Fig. 14.2. The
single strategy model assumes no lateral flow. We hope to learn the
consequences of these effects.
The small pulse in the acceleration generates effects in all the other
scalars. For example, the player interest for player 1 is modified,
Fig. 14.8, so that there is more flow along the center x  0 than at the
edges. If the pulse is turned off, these streamlines become uniform flows
independent of x .

Figure 14.7. Acceleration flow q Figure 14.8. Player-1 interest flow I1


Vorticity and Active Strategies 527

14.12 Player Bonding Effects

Though we obtain more insight into the player payoffs for 3  1


dimension, we already see interesting effects for the composite payoff
(Fig. 14.9) and the player payoff (Fig. 14.10). The effects are result from
the acceleration and player bonding scalars and our assumed non-zero
scale factor (Ex. 9). Though we start the payoff at z  0 and for all x ,
the field equations for the payoff in Eq. (14.43) generate a non-zero
gradient due to the player interest term  I z  , which is non-zero for
player 1 and player 2. We thus expect each to have a payoff, with player
1 having opposite gradient based on the initial conditions. We thus
expect each to have a payoff, with player 1 having opposite gradient
based on the initial conditions.

Figure 14.9. Composite payoff  Figure 14.10. Player 2 payoff B2

We have an extremely useful set of models for which we can obtain


quantitative behaviors (e.g. Exs. 10-12). We will return to these models
for support of our qualitative arguments. However, the results are
limiting in that the composite payoff and player payoffs in these models
are decoupled (Ex. 4). For 3  1 and higher dimensional models, the
composite payoff is expected to be influenced by the player payoffs. Thus
to further understand that issue we turn our attention to those lattice
models. The anticipated tradeoff is that lattice calculations of such
models may not be as numerically accurate: a 100 point lattice will be 10
points along each direction. We address this in the next section.
528 Geometry, Language and Strategy—Vol. 2

14.13 3+1 Active Strategy Lattice Models

In the last section, we showed that the 2  1 decision process models,


using a numerical lattice for the points, achieved reasonable accuracy,
making the tradeoff between providing sufficient lattice points and the
demands of solving elliptic partial differential equations which are not
always stable when expanding from an initial spatial boundary. This
approach may lead to instability under small changes of the initial
conditions that becomes manifest when the number of lattice points
becomes too high [Sofroniou & Knapp, 2008].
In the work that follows, we believe a significant improvement in the
use of lattices is to use NDSolve and the numerical method of lines that
is provided by Mathematica. In many ways it is similar, though it has
evolved over time and seems to handle the instability issues better. Also
it works well when the non-focus directions satisfy periodic boundary
conditions. It can be used to provide the behavior of the equations
described in Sec. 14.9. Moreover, as the technology for solving
differential equations improves, the upgrades can be used transparently.
By using NDSolve, we take the previous 2  1 model discussions
(Sec. 14.10) and extend them to a 3  1 dimensional model (Ex. 13) with
characteristics similar to the example in Ex. 12 in which the player
payoffs for the two players were offset.
We thus obtain a three dimensional view for three players: two
individual players associated with worldview strategies and one
institutional player associated with a code of conduct strategy.
The meaning of player will depend on the context of the specific
model. In general terms, there will be players that have the ability to
make decisions for a subset of strategies that they own exclusively. There
will also be players that are really seasonal, corporate or institutional
players that own many strategies that may include up to all the strategies
or no strategies at all.
Vorticity and Active Strategies 529

Figure 14.11. Payoff vector Figure 14.12. Payoff vector


field B1 field B 2

We put ownership of the code of conduct (Sec. 11.2) in this context.


For the numerical examples in this section, we assume the ownership
model, that player 1 and player 2 have exclusive ownership of the active
strategies x and y . We assume that player 3 is an institutional player,
administers the code of conduct and has joint ownership of all strategies,
including the focused strategy z . We say that a strategy is focused when
the inertial forces (pressure) produce one or more maxima and allow us
to treat the behaviors away from such points with a Cauchy development
of the differential equations.
The player ownership is clear in the
player payoffs, Fig. 14.11 and Fig. 14.12.
They show that player 1 owns the x, z
directions and player 2 owns the  y, z
directions. The vertical axis z corresponds to
the axis along which pressure has a sharp
peak, which is similar to that in Fig. 14.3.
The periodicity in the directions transverse to
z result in (pseudo) periodicity for z .
From the form of the equation for the
composite payoff ω  e Bα , Eq. (14.30) and
based on the boundary conditions, we expect Figure 14.13. Composite
and see that the composite payoff (Fig. 14.13) payoff field ω
530 Geometry, Language and Strategy—Vol. 2

is an amalgam of the payoffs for both


players, with no contribution from the third
player since we assumed a zero initial
condition for player 3 charge, Ex. 13. Though
the two players have picked quite opposite
behaviors, a common cause behavior results
and is exhibited by the composite payoff.
The approximate form for the composite
payoff is modified by the behavior of the
seasonal potential A  2e q a , Eq. (6.10),
Fig. 14.14. We meet the gauge condition
Figure 14.14. Characteristic constraint (Sec. 15.14, Ex. 7) by fixing the
potential a
boundary values for a z  x a z  to be zero.
We see that the resultant vector field has small components along the
z axis. The significance of this field is that it adds the commutator terms
to the time dependent Eq. (14.33) and determines the composite payoff.
We find that the stresses p , i.e. the player passion Pα , describe the
gradient of the charge (ownership model) for each player (for player 1,
Fig. 14.15 and for player 2, Fig. 14.16). Because the player passion flow
is conserved, the flow is constrained, as we have anticipated (Ex. 11).

Figure 14.15. Passion vector Figure 14.16. Passion vector


field P1 field P2
Vorticity and Active Strategies 531

14.14 Institutional Vortex

The player passion flow acts like a current that generates the player
payoff. The assumption of ownership shows that this flow moves in the
x, z direction producing a payoff for player 1 that moves along the y
axis (Fig. 14.11); the corresponding behavior for player 2 has a payoff
that moves along the x axis. We noted that the common cause of the two
players is a circulation around the z axis.

Figure 14.18. Composite


Figure 14.17. Payoff vector
payoff ω with code of
field B 3 with code of conduct
conduct

This suggests that a code of conduct could be added with the same
basic behavior (Ex. 14). In this case we obtain a new player payoff,
Fig. 14.17. This modifies the composite
payoff, resulting in Fig. 14.18.
The player passion for player 3 that
generates this payoff, Fig. 14.19, has
essentially a flow along the z axis at the
center of the x, y plane. To conserve the
flow, we see that there is an inward flow
towards this axis all along the z direction.
The player interest, Fig. 14.20 is different in
that there is a non-zero initial gradient so that
the code of conduct payoff contributes to the Figure 14.19. Passion vector
composite payoff at z  0 . field P3 with code of conduct
532 Geometry, Language and Strategy—Vol. 2

We have sketched how the numerical


method of lines provides the means to obtain
the behavior of the scalar quantities of the
decision process theory. We suggest that the
way to approach decision problems is to use
these behaviors as a backdrop for discussion
in which one first starts with the problem
statement (Cf. Sec. 10.2) followed by putting
the problem into normal form (Cf. Sec. 10.3).
The behaviors we have suggested here
are applied to the problem in normal form
following a discussion of the possible
Figure 14.20. Interest vector boundary conditions that are appropriate.
field I3 with code of conduct Because of the sensitivity of the outcomes to
the initial conditions for the numerical method of lines, care must be
taken that the results have sufficient accuracy. This issue might be solved
using other methods such as finite element analysis [Bhatti, 2005]. The
finite element is planned for a future release of Mathematica; it currently
works for linear partial differential equations. The advantage of using
NDSolve from Mathematica is that as new approaches are incorporated,
they can be incorporated into existing calculations with little change.
When we can identify a focus direction, we assert that we can use
the numerical method of lines to carry out our program of investigating
3  1 dimensional models.

14.15 Outcomes

The student will have learned how to identify and apply known
conditions to the decision process field equations for 2  1 , 3  1 and
higher dimensions. In the process, the student should be able to identify
those scalar fields that are steady state in the steady state models and
those fields that depend on proper-time. At a high level the student
should understand in what sense the flow equations generates “vorticity”
solutions that are harmonic and distinguish between the harmonic
components that are part of the initial conditions and vorticity
components that are intrinsic and set by the composite payoff, player
payoffs, engagements, entitlements, passions and interests of the decision
process.
Vorticity and Active Strategies 533

The student should be able to distinguish between the steady state


fields whose values are set independently on a subspace such as z  0
and steady state fields whose values are constrained on that same
subspace.
The attainment of the outcomes of this chapter is facilitated by doing
the exercises in the following section. Based on this investment, the
student should achieve the more detailed outcomes below.
 From Sec. 14.2: The student should be able to describe the difference
between player preferences and player effort. Furthermore the student
should be able to distinguish between “worldview” players and “code
of conduct” or institutional players.
 From Sec. 14.3, the student should be able to understand the
difference between purely resource driven processes and purely skill
driven processes.
 From Sec. 14.4: The student should be able to explain why the still
point is characterized by zero acceleration, zero engagement and
player entitlement. The student should be able to explain why the
metric of differential geometry provides the mechanism to define
utility for decision process theory.
 From Sec. 14.5, the student should be able to use the metric distance
between points to describe events in a variety of reference frames,
including the normal coordinate frame, the co-moving coordinate
frame and the symmetric co-moving coordinate frame.
 From Sec. 14.6, the student should be able to identify the surface
geometry Eqs. (14.10)-(14.12), which provide a complete set of
equations. The student should be able to identify which equations are
essentially divergence equations and which are “curl” equations.
 From Sec. 14.7, the student should be able to see the striking
difference with the competitive field equations that can generate
vortices. These are essentially the “curl” equations. The student
should understand the geometric interpretation of such fields.
 From Sec. 14.8, the student should understand the set of equations for
the stress and strain fields.
 From Sec. 14.9: The student should be able to rewrite the field
equations in 3  1 dimensional notation. The student should
appreciate that although the original field equations were very
534 Geometry, Language and Strategy—Vol. 2

compact they lead to a vast number of individual “laws” that can be


investigated.
 From Sec. 14.9.1, the student should be able to describe the
contributions to the average pressure and explain why each
contribution is either positive or negative. The student should
understand that the average pressure p and the energy density
   p are dependent scalars and hence determined when all other
scalar fields are determined.
 From Sec. 14.9.2, the student should be able to qualitatively describe
the acceleration Eq. (14.20) and its contribution and explain why the
orientation potential q and its gradient  z q along z can be specified
independently on the boundary surface z  0 .
 From Sec. 14.9.3, the student should be able to qualitatively describe
the player bonding Eq. (14.21) and explain why the player bonding
potential  and its gradient  z  can be specified independently on
the boundary surface z  0 . Based on the fact that the energy density
is a source, the student should be able to argue why this equation has
some similarity to a gravitational equation.
 From Sec. 14.9.4, the student should be able to qualitatively describe
the shear components Eq. (14.22), as cooperative effects between
players and explain why the shear component potentials   and
their gradients  z  if specified independently would provide a
complete solution. The student should understand why this is not
sufficient because the gradients considered as matrices must mutually
commute. The student should be able to explain how these constraints
can be imposed in practice.
 From Sec. 14.9.5, the student should be able to qualitatively describe
the player interest Eq. (14.23) as providing a complete divergence and
“curl” equation from which to solve for the player interest I . The
student should further understand that the player charge Eq. (14.29) if
used automatically satisfies the “curl” equation. Thus the player
charge e and the player interest component I z , which specifies the
charge gradient, are sufficient to determine the player charges and the
interest component I z . The remaining components I x I y  are
dependent fields, determined by Eq. (14.29). The student should
Vorticity and Active Strategies 535

understand that these equations may impose constraints on the


boundary z  0 .
 From Sec. 14.9.6, the student should be able to qualitatively describe
the player payoff Eq. (14.24) and explain why the payoffs B are
determined by the divergence and “curl” equations. The student
should be able to explain why one of the “curl” equations is a
constraint as it involves no derivatives along z . This constraint can be
used to determine Pz on the boundary as well as overall space.
 From Sec. 14.9.7, the student should be able to qualitatively describe
the player passion Eq. (14.25) and explain why in the ownership
model; only the player passion potential P needs to be specified to
obtain a complete solution. The student should be able to explain why
we obtain the same situation for the conductivity model.
 From Sec. 14.9.8, the student should understand that the composite
payoff ω on the boundary z  0 is determined by Eq. (14.30) and
can be used to determine the components B3 , one of the constraints
on the boundary.
 From Sec. 14.9.9, the student should understand the stress
components and the model that has been used to construct them.
 From Sec. 14.9.10, the student should understand that the frame
transformations are constrained in the player fixed frame model, so
that they correspond primarily to changes in scale not direction.
 From Sec. 14.9.11, the student should be able to qualitatively describe
the seasonal potential A , Eq. (14.31) and explain why the potential
components A and their gradients  z A would in the absence of the
gauge condition, be sufficient to determine a unique solution. The
student should understand that the gauge condition specifies  z A ,
removing one of the choices. The student should understand that an
additional choice can be made, Sec. 15.15, Ex. 7, which provides a
differential equation on the boundary z  0 that determines Az on the
boundary. Thus the only freedom will be the remaining seasonal
potential components and their gradients. If the gauge components are
zero, that in fact leaves only the gradients.
 From Sec. 14.9.12, the student should learn that the proper time
equations can provide harmonic solutions that maintain the stationary
aspect of the surface geometry.
536 Geometry, Language and Strategy—Vol. 2

 Similarly, in Sec. 14.9.13, the central time frame equations provide an


equivalent description.
 From Sec. 14.9.14, the ideas come together with a description of the
travelling and standing wave solutions that result from the wave
equations.
 From Sec. 14.10 the student should be able to solve the simplified
non-linear Laplace equation and generate the contour plots Fig. 14.1
and Fig. 14.2. The student should be able to rewrite the field
equations to apply to 2  1 dimensions.
 From Sec. 14.11, the student should be able to explain locked
behavior and how it arises.
 From Sec. 14.12, the student should be able to explain the effects of
player bonding.
 From Sec. 14.13, in 3  1 dimensions, the student should be able to
explain the player payoffs Fig. 14.11 qualitatively from the boundary
conditions. The student should be able to explain qualitatively the
composite payoff vector field Fig. 14.13. The student should be able
to explain qualitatively the player passion vector field Fig. 14.15.
 From Sec. 14.14, the student should understand how vortices can
arise, for example in the player and composite payoffs.

14.16 Exercises

(1) Show that for the conductivity model, the passion vectors in
Eq. (14.25) can be determined in terms of scalars using Sec. 6.10
Ex. 14, Pα  P  P :

n 2  
2 P   i h   2     P . (14.41)
 ni 

(2) Show that the integrating factor technique can also be used to write
Iα  I  i , Eq. (14.23) and ω    , Eq. (14.26) in terms of

scalar potentials and Bα  B  A Eq. (14.24) in terms of vector
potentials. Determine the relationship of i to the player
engagement e .
Vorticity and Active Strategies 537

(3) We expect new and interesting features with three or more active
(spatial) strategies. Nevertheless, there are useful things to learn
from models with two active spatial strategies. Show that in the case
of Eq. (14.15) and zero shear    0 , the acceleration, player
bonding and player interest behaviors from Eq. (14.10) can
be expressed in terms of I α  2x 2z  , q  q1 q2 ,  
Θ   x  z  , Pα   p x pz  , B   xz and    xz as
follows, showing that the player composite payoff and player payoff
fields are not coupled:

   x  z ,
h  n  2  ni ,
 1 
e  I α   q  Θα  e ,
 ni 
 n 1 
  I α  4 B   2q  i Θ   Iα ,
 ni 
 1   1 
 x I z   z I1    qx   x  I z   qz   z  I x ,
 ni   ni 

  p
  q  q  q  q  Θ  2 2  1 2 I α  I α      p  ,
 n 1 
n 1
      1 2 i Θ  Θ  3 4 I α  I α  3 2  B B . (14.42)
ni

(4) With the same set of assumptions as Ex. 3, show that the player
passion, player payoffs and composite payoff are determined in the
ownership model by:

Pα  P ,
 n 1 
  Pα   q  i Θ   Pα ,
 ni 
 n 1 
 z B   qz  i  z  B  I z    x P ,
 ni 
538 Geometry, Language and Strategy—Vol. 2

 n 1 
 x B   qx  i  x  B  I x    z P ,
 ni 
 z   z    2 q  ,  x   x    2 q  ,
   w exp    2q  . (14.43)

(5) With the same assumptions as Ex. 3, show that the average pressure
is determined by:

ni  1
q  Θ  12 Θ Θ  3 4 Iα  Iα  B B  3 2
ni
p . (14.44)
n 
 ni
n 1

(6) With the same assumptions as Ex. 3, show that the active and
inactive pressure components are:

 p      q   q q    


1
     1 2 I I 
ni
  p
  p      2   B B    ,
2 

 n 1 
 p    B B  1 2 I α  Iβ  . (14.45)

(7) With the same assumptions as Ex. 3, show that the equations reduce
to the following generalization of the single strategy model when
the player payoffs and composite payoff are zero:

 n 1 
 I α   2q  i Θ  Iα ,
 ni 
 1   1 
 x I z   z I x    qx   x  I z   qz   z  I x ,
 ni   ni 
  1
 q  q q  q Θ  1 2 I α I α   p    1  ,
 n 1 
Vorticity and Active Strategies 539

ni  1
    p  1 2 Θ Θ  3 4 Iα Iα ,
ni
Pα  0 . (14.46)

(8) Using Ex. 7, generate the pressure Fig. 14.3, and acceleration curves
Fig. 14.4, using n  6 spatial dimensions, the implied ni  4
inactive dimensions,   1 , resiliency   5 , 3 mesh points for x
in the interval  1, 1 , z in this interval and the following initial
values:

q  x,0   qz  x,0   0
  x ,0    z  x,0   0
. (14.47)
I x  x,0   e  x,0   0 0 0 0
I z  x,0   4 4 0 0

(9) Building now on Ex. 8, adding back the possibility of vorticity and
player payoffs, provide structure in x with x dependent boundary
conditions and an increase in the number of mesh points to 11 as
given below, and a resultant smaller interval z   0.65, 0.65 ,
generate the pressure Fig. 14.5 and acceleration surfaces Fig. 14.6 as
well as the streamlines for the acceleration flow Fig. 14.7 and player
interest flow Fig. 14.8. Do the results depend sensitively on the
number of mesh points above about 10? Note that the boundary
conditions for the composite payoff (up to a scale), the player
interest and for the ownership model, the (gradient of the) player
passion are determined by the field equations on the boundary.

 1 
I x  x,0    x e  x,0    qx  x,0    x  x,0   e  x,0  ,
 ni 
 z P  x,0    x B  x,0   I x  x,0    x,0 
 n 1 
  qx  x,0   i  x  x,0   B  x,0  ,
 ni 
540 Geometry, Language and Strategy—Vol. 2

x
q  x,0    1 5 cos2 ,
2
qz  x,0   0 ,
  x,0    z  x,0   0 ,
e  x,0   0 0 0 0 ,
I z  x,0   4 4 0 0 ,
B  x,0   P  x,0   0 ,
  x,0   110 exp    x,0   2 q  x,0   . (14.48)

(10) Using the same model as in Ex. 9, and making the changes below to
the initial conditions, show that the player interest (Fig. 14.21) and
player passion (Fig. 14.22) are as indicated below using a mesh with
11 points.

q  x,0   0
 x 1 x 
B  x,0    110 cos2 ,  10 cos2 ,0,0  . (14.49)
 2 2 
e  x,0   0 0 5  5
1 1

Figure 14.21. Interest flow I1 for Figure 14.22. Player passion P1 for
Ex. 10 Ex. 10
Vorticity and Active Strategies 541

(11) Using the same model as in Ex. 9, and making the changes below to
the initial conditions, show that the player interest (Fig. 14.23) and
player passion (Fig. 14.24) are as indicated using a mesh with 11
points.

q  x,0   0
. (14.50)
e  x,0    1 20 sin  x  1 20 sin  x 1
20 cos  x  1 20 cos  x

Figure 14.23. Player interest I1 for Figure 14.24. Player passion P1 for
Ex. 11 Ex. 11

(12) The various factors can lead to complex behaviors. Using the same
model as in Ex. 9, and making the changes below to the initial
conditions, show that the player passion for player 1 (Fig. 14.25)
and player passion (Fig. 14.26) are as indicated below using a mesh
with 11 points. This shows that they can be chosen to be exactly out
of phase using the player payoffs and can be said to provide each
player, ownership of the stresses along his or her flow direction,
which as shown are independent.

q  x,0   0
e  x,0   0 0 1
10  110 . (14.51)
B  x,0   110 cos2 1 2  x, 110 cos2  1 2  x  1 4   ,0,0
542 Geometry, Language and Strategy—Vol. 2

Figure 14.25. Player passion P1 for Figure 14.26. Player passion P2 for
Ex. 12 Ex. 12

(13) Use Mathematica and NDSolve on the 3  1 model described in Sec.


14.9, along with the gauge condition from Sec. 15.15, Ex. 7 to
generate Figs. 14.11-14.16 in Sec. 14.13. For the initial conditions,
use Eq. (14.29) for I x  x, y , z  I y  x, y, z  in terms of the player
engagement e  x , y , z  and take the following conditions, which
are similar to those in the above exercises for the 2  1 lattice model
with endpoint set at 5 :
q  x, y ,0    z q  x, y ,0     x, y ,0    z   x , y ,0      x, y , z   0
e  x , y ,0   0 0 0
I z  x, y ,0   4 4 0
P  x, y ,0   110 cos2 1
10  x  110 cos2 110  y 0 . (14.52)
Bx  x, y ,0   B y  x, y ,0   Bz  x, y ,0   0 0 0
a x  x, y ,0   a y  x, y,0    z a x  x, y ,0    z a y  x, y ,0   0
a z  0, y ,0    x a z  0, y ,0   0

(14) Modify Ex. 13 to allow for player 3 to influence behavior. Interpret


player 3 as having a passion for a common cause solution, with the
following changes that allow the common cause solution to own all
three strategies. Show that the resultant changes are primarily to the
payoff for player 3 Fig. 14.17 and the corresponding composite
Vorticity and Active Strategies 543

payoff, Fig. 14.18. We also see new behaviors in the passion for
player 3, Fig. 14.19 and in the interest, Fig. 14.20.

e  x, y ,0   0 0 1
5 
. (14.53)
P  x, y ,0    1
10 cos 2 1
10  x  110 cos2 110  y 1
10 cos2 1
10  x cos2 110  y
Chapter 15

Game Equivalency

In Sec. 14.6, we explored the steady state surface geometry attributes of


decision process theory. In this section we explore the dynamic aspects
and in particular, those aspects that use game theory to define values on
the initial surface geometry boundary; as an example, the harmonic
behavior attributes of the energy flow that are associated with the game
theory concept of Nash equilibrium. To make the correspondence to
game theory, we distinguish two concepts, still point and free fall. We
illustrate this for decision processes with three active strategies that have
a non-trivial time development; they are not static or steady state.
Consider that at some initial point in time, we specify the strategic
behaviors on a hypersurface orthogonal to the time direction that we take
to be the direction of the energy flow. The orthogonal hypersurface
defines the co-moving frame. We have considered models that are steady
state in the sense of Chap. 6. We are not free to specify these steady state
fields arbitrarily as they are constrained by the field equations. The field
equations then provide the time evolution from this hypersurface.
A time evolution will provide the fields that describe decision
processes. We can obtain all of the time dependence from knowledge of
the scalar coordinate fields, which act like potentials in electricity: they
define surfaces of constant value that provide meaning to the coordinate
position. They have the same usefulness as surfaces of constant voltage.
We continue the analogy: the time dependence can be decomposed into
harmonic frequencies so that the time behavior is translated into a study
of phasors for each frequency. The resultant phasor equations provide
additional steady state fields to consider on the hypersurface.

545
546 Geometry, Language and Strategy—Vol. 2

To obtain numerical results and start comparisons with real world


behaviors, we must compute these fields on the hypersurface. We
compute the fields by specifying values on an initial boundary. We then
use the field equations to compute their value on the hypersurface.
The goal in this chapter is to identify the values on the initial
boundary. First of all, the fields depend on what we identify as the active
strategies, transported to the orthonormal coordinate frame. For a given
decision process, what are these active strategies? We propose a game
equivalency principle that is helpful in choosing the active strategies,
defining the initial boundary surface and setting values on that surface.
To carry out this program, we identify all the strategies normally
considered in game theory. There are a few caveats. First of all, we don’t
include the internal strategies that each player maintains for their own
worldview.
Next we make a tentative identification of inactive strategies, which
are symmetries of the problem, which may later turn out not to be
symmetries. Our first tentative choice is that the total effort expended is
inactive, as that effort though included, is normally not considered
significant in game theory.
We do include relative player efforts. We note that these relative
efforts appear to correspond to the imputations introduced by Von
Neumann & Morgenstern [1944] and so in fact may be relevant as
strategies.
We organize the chapter by defining in Sec. 15.1, the circumstances
in which decision process theory might reduce to game theory. In Sec.
15.2, we extend the discussion of equivalency to include the distinction
of free fall or no active forces causing changes to decision flows. The
concept of still point is related to this, which is a natural extension of
Nash equilibrium. The field equations require a focus boundary on which
we specify the initial values of each of the scalars. This discussion is not
limited to two-person games.
From these initial values, the field equations then provide values at all
other positions. This defines the steady state geometry. We discuss two
interesting aspects of this geometry: player engagement in Sec. 15.3 and
entitlement in Sec. 15.4. In Sec. 15.5, we consider the behavior
orthogonal to the focus boundary, which we call the focus direction.
Game Equivalency 547

In Sec. 15.6, we use the fact that we can write the payoffs and their
potentials in terms of a symmetric metric to deduce values for metric
components. We use these results to translate the game payoffs into
potential field gradients, Sec. 15.7, and entitlement payoffs, Sec. 15.8.
The player interest or game theory value helps set the player
engagement, Sec. 15.9. New aspects of decision process theory not
covered by game theory are the cooperative payoffs, Sec. 15.10 and the
seasonal vector potentials Sec. 15.11, which lead to natural resonances.
We return to the inactive strategies in Sec. 15.12 with an explicit
accounting of scale invariance for two-person games with three active
strategies. We summarize how to set initial values for such models in
Sec. 15.13.

15.1 Equivalency

The idea of game equivalency is a helpful guide to setting the boundary


conditions for the dynamic model. There are various aspects of this. At a
theoretical level, game theory provides our basis for strategy distinctions.
We start with that idea and assume that on our initial boundary, we have
identifiable players, which means we have a given set of inactive
strategies that play no direct role in the decision making. We have
identifiable game strategies that play a strategic role by means of the
payoff matrices, which on the initial boundary we take to be the same as
the game theory.
For each player, we identify the player effort strategy rk , Sec. 14.2,
which is one of the game strategies. We identify the remaining player
relative preferences as the player skill preferences. With these
distinctions, we express the symmetric payoff matrix in terms of the
player skill strategies, the focus strategies rj  rk representing the
relative player efforts, and the total effort strategy 1 n  rk , Sec. 14.2.
These distinctions rearrange the given set of game strategies into an
equivalent set.
We make a minimal set of dynamic assumptions that leave the
dynamic possibilities open and reflect key aspects of game theory. If we
choose two or more game strategies as inactive, the payoff between these
548 Geometry, Language and Strategy—Vol. 2

two strategies must be zero. As a general rule, this is an unjustifiable


limitation on the initial game theory values. However, we can choose one
game strategy as inactive. For a game with any number of players, there
is some justification for choosing the total effort strategy as inactive,
since there is no strategic consequence in game theory based on this
strategy’s value. We assume this scale invariance as part of equivalency.
It does not appear to close significantly the dynamic possibilities.
We make simplifying assumption about the focus strategies. Though
in general, we don’t take them to be inactive, we set their values on a
specific initial boundary. Consider the focus subspace spanned by the
focus strategy directions. We define the focus boundary as a specific
closed subspace from a consideration of two, three and more players.
Consider a decision process with two players: there is one focus
strategy z12  r1  r2  z . The focus boundary that makes the most sense
is z  0 , which aligns with game theory for two players. Without
changing the strategic content, we can always force the player efforts to
match. The focus boundary is defined as that closed point.
For three players, the situation is more complicated. We have three
focus strategies zij  ri  rj constrained so that their sum is zero:
z12  z31  z23  0 . The focus subspace is two dimensional. There is a
natural triangular boundary formed by these three variables, analogous to
the imputation diagrams used by Von Neumann & Morgenstern [1944, p.
282ff.].
Consider an equilateral triangle such that the center of the triangle
corresponds to all three focus strategies being zero. Through this point
draw three lines parallel to each of the triangle legs. Associate each leg to
one of the focus strategies and assume the vertical distance from the leg
to the center is s . This defines a coordinate system with three variables
appropriately constrained. For any s the triangle defines the focus
boundary. In similar manner, one can identify the focus boundary for
any number of players, given a distance s from the center of a regular
convex solid to one of its faces.
On the focus boundary, we assume the game theory values for the
payoffs. However, there are some caveats. The fields, such as the payoff
fields, depend on the focus strategies and the skill strategies, assuming
scale invariance. We are free to specify that the payoffs are independent
Game Equivalency 549

of the focus strategies if we wish. We can specify that the payoffs have a
functional dependency on the skill strategies if we wish. For equivalency
as used here, we require that the dependency on the skill strategies be
periodic, which assumes that for each skill strategy, there is a distance
specifying the period. This defines the initial boundary.
With these assumptions and the further assumption that the
dependency on the skill strategies remains periodic, it is plausible that
the partial differential equations have unique solutions. We have
specified the equations on the initial boundary that consists of the closed
focus boundary with arbitrary skill strategies and the periodic boundary
for each of the skill strategies for any value of the focus strategies. We
find in practice that this leads to unique solutions.
Our practical set of cases will be two-person games where each
person has two strategies. There are four game strategies. The focus
subspace is one dimensional and the focus boundary is the point z  0 .
Let’s call the skill strategies x, y and the scale invariance strategy r .
We specify every scalar   x, y , z  on the initial boundary as follows.
For z  0 we specify an arbitrary function   x, y ,0  subject only to the
constraints that the function is periodic in the skill strategies. For every
z  0 we impose the boundary condition that the solutions remain
periodic in the skill strategies. The partial differential equations are then
expected to yield unique results for this scalar function. The same holds
true for vector and tensor quantities.
We now identify those functions that can be specified by game theory
and hence invoke a game equivalency. A necessary ingredient of
equivalency is that in some sense, the competitive forces are the only
forces that matter. Even with just competitive forces, for multi-player
games, we have a form of cooperation through the focus strategies. Other
than that, there should be no other forces. We explore this in the next
section.

15.2 Still Point and Free Fall

Suppose there are only competitive forces. First, we are assuming that the
payoffs are independent of time and that the flows are independent of
550 Geometry, Language and Strategy—Vol. 2

time: we have a steady state condition. Second, we are assuming that


there are no other forces causing changes to the flow. We need to work
out the dynamic consequences of these assumptions. We start with
steady state flow.
A steady state flow does not imply that the flow value necessarily
reflects Nash equilibrium. However, we would like to connect to that
concept since it is an important underpinning of game theory. We think
of this equilibrium as a still point on the focus boundary that has special
characteristics akin to the eye of a hurricane. For example, we assume
that the still point has zero active acceleration Q .
The motion around such a point is steady state but not static; the
forces at that point are exactly balanced and there may be many such
points. Thus a still point might be thought of as the Nash equilibrium,
without the corresponding assumption of static behavior throughout the
neighborhood. There is cyclonic behavior that is steady state. This allows
us to identify player payoffs at the still point with game theoretic
equilibrium values and still identify the cyclonic behavior with the
payoffs. Our concept of steady state here is also consistent with the
literature of steady state being associated with the existence of a frame in
which time is an isometry (the central holonomic frame).
We specify the values of the scalar fields over the whole focus
boundary, which goes beyond the standard game theory (Cf. Sec. 14.13).
To do this we invoke the notion of free fall, which in physics is the
notion of falling in an elevator, which provides the local neighborhood.
Locally, inside the elevator, there is no notion of gravity or rotational
forces. The acceleration is zero. With this concept we move away from
the behavior at a point to behavior in a local neighborhood.
To measure the true acceleration, we need the covariant rate of
change, Eq. (2.40). For example, the covariant rate of change of the flow
can be expressed in a way that is independent of how we choose the
active strategies or the frame used to measure them: V a   a bcV bV c ,
Eq. (2.44). This expression is covariant for all transformations of the
active strategies among themselves.
This rate of change depends on our perceived acceleration of the
active strategy flow, the active geometry acceleration Q , Ex. 29 from
Chap. 6, which is the sum of the absolute acceleration, the competitive
Game Equivalency 551

acceleration component and the cooperative acceleration components.


These terms have explicit definitions in any holonomic basis for the
active strategies.
In the central holonomic frame, the expressions can be related to the
potential fields of the co-moving basis Exs. 19-29 from Chap. 6. We
apply this to our current analysis in Ex. 2 and Eq. (15.31). When the
player engagement vanishes, all that remains is the absolute force,
supporting our provisional definition. In practice not all components of
the player engagement vanish, so we modify our provisional definition
of still point to mean the vanishing Q  0 of the active geometry
acceleration in the central holonomic frame.
When we are at the still point V a   a bcV bV c  0 , we find that the
space may not be locally flat,  a bc  0 for at least some components. For
example in the central holonomic frame, we have V a  0 identically
and the active geometry acceleration is Qa   a e 2 q . In general this
vanishes only at the still point.
Moreover, this suggests only some of the components  abc vanish;
not all. This is really a new situation that suggests that the acceleration is
dependent on forces that can be transformed away at a point by viewing
the strategies in a locally flat frame. However, in general there is no
transformation that will simultaneously transform points in an extended
region to be flat. We say that the system is in free fall if we are both at a
still point and the choice of strategies is locally flat. This is the case of
falling in an elevator.
If we take this behavior as setting the known conditions on the initial
boundary, then we not only obtain game equivalency at the still point,
but extend it so that locally, the only forces operating are the competitive
forces. The basic conservation law, Eq. (2.49), has no contributions from
cooperative, organizational or frame transformation effects. This means
that the rate of change of the flow is determined entirely by the product
of the payoff matrix and the flow.

dV b
g ab  Vk Fabk V b . (15.1)
d
552 Geometry, Language and Strategy—Vol. 2

The solution to this equation has only discrete harmonic solutions


corresponding to the eigenvalues of the payoff matrix. This was used in
Vol. 1 to demonstrate natural resonance solutions with discrete
frequencies. We explore this in more detail in Sec. 15.11.

15.3 Engagement

To set the values of all of the scalars on the initial boundary, we


supplement game equivalency with two concepts. The first is the player
engagement e , a non-game theoretic distinction associated with the
still point. This is directly related to the inactive flow V j in Eq. (15.1).
The second, which we deal with in the next section, is the player
entitlement payoff. This is the payoff a player perceives when their
player engagement is zero.
The inactive flow represents a charge that would normally be
incorporated into the definition of the payoff. This charge is related to
the player engagement by Eq. (4.54):

V j  E j   e E j . (15.2)

Given that we align the frames initially in such a way that the frame
transformation is the unit matrix, the initial boundary value of the charge
is the (negative of the) player engagement.
For game equivalency, the active acceleration is Q  0 . In addition,
the co-moving frame is aligned with the normal coordinate frame
(indicated by  ). Therefore we obtain constraints based on Eq. (4.74):

f j
 
 2 q E j       E j  e E j    e E j   1  E k
 Ek  
. (15.3)
f j
   2  E j  2  E j

Requiring the active acceleration to vanish provides a constraint on the


acceleration q , Eq. (6.78):
Game Equivalency 553

q  2 e   e e 
Q  0
1  e e 

 2

 q   I e   e e  . (15.4)

These relationships relate the player interest to the player strategy bias
on the initial boundary:

f j     
k
 2 j e  2 e  e e j  I j  I e  e j
1  e e k
j
e j f 
  2  e e  e I  . (15.5)
1   e e 
 2

We rewrite these relations to express the player interest at the still point
in terms of the game value, player bonding and player engagement.

j f j e jVk f k   
I    2 j e . (15.6)
1  e  e 1   e  e 2

We conclude that when the player engagement is zero, the player interest
is set by f j .
We use this result and Eq. (4.69) to determine the player engagement
gradient in terms of the game value and player bonding on the initial
boundary:

 e j  I j  q e j   j  e 
 f j   
. (15.7)
 e j  
  j e  e j   e e
1  e e

We see that if the player engagement is zero, e j  0 , then the player


interest and player engagement gradients are zero. In drawing these
conclusions, the active strategies are the transformed game theory pure
strategies in the co-moving orthonormal coordinate frame. We associate
the proper strategies in that frame on the initial boundary and assume
554 Geometry, Language and Strategy—Vol. 2

that the player bonding matrix is diagonal. Later we modify these


assumptions.

15.4 Entitlement

We suggest that when the player engagement is zero, we open up an


additional and important distinction that we call player entitlement,
which is related to the player’s view of the payoff unrestrained by
external interactions. We consider the non-pejorative definition
suggested by Gladwell [2008, p. 105]:

“But Lareau means it in the best sense of the term: ‘they acted as though they had a
right to pursue their own individual preferences and to actively manage interactions in
institutional settings.’”

To get to this distinction, we start with the game theory related


concept that pure strategies are characterized by the player strategy bias
f j . The player strategy bias f j is related to the game value. Our
result for zero player engagement is the expectation that the game value
is also zero.
We propose that the player engagement is a property of the player,
not of the interaction. Thus we believe it represents something about how
the player approaches the decision process. The player can choose to
engage in the decision process, in reasonable analogy to the way charge
interacts with a magnetic field. Just as the strength of the magnetic field
is independent of the charge, we suggest that the player engagement is
independent of the player payoff.
There are other consequences to charge as well, such as the existence
of an electric field; with that analogy, here we would refer to the
existence of the player strategy bias. Following this analogy, if there is a
charge, then there is an electric field generated by the charge. That
suggests that if there is non-zero player engagement, there will also be
something analogous to the electric field, namely a non-zero game value.
This is the result that we have obtained.
Consider the co-moving player payoff f j  , Eq. (15.3):
Game Equivalency 555

j 
f j
   2  E j  2  E j  2    2 e j . (15.8)

Player entitlement occurs when the player engagement is zero. This


allows us to define the co-moving player payoff by the player entitlement

payoff  j   , with no contribution from the collective decision behavior
or composite payoff   . In general, the co-moving player payoff will be
the sum of the two. The composite payoff  e emphasizes that player
engagement is required to move a player away from “pursuing their own
individual preferences.”
We add player entitlement and player engagement, non-game
theoretic attributes, to the properties of the still point: we assume that for
real worldview players (but not for institutional code of conduct players),
their player engagement e needs to be set along with the player interest.
For the latter we expect guidance from the game value. By considering
different initial conditions, we may be able to measure the player
entitlement and the player engagement.
One of our choices is to identify one or more strategies as part of an
institutional code of conduct. Though we may have reason to believe that
the players act as if they are entitled, there is no reason to believe that the
institutional players behave in the same way. Moreover, based on the
equations, there is no reason to assume that the resulting bonding matrix
in Eqs. (15.6) and (15.7) is diagonal.
As an example of what might happen, consider the scenario in which
we have scale invariance. In this case, the active strategies can be chosen
to be the preference differences. Typically, the game values are the same
for all pure strategies for each player. So if  is a player skill preference,
it is the difference between two pure preferences for a given player. We
then expect that the player strategy bias components f j to be zero for
each worldview player. In this case, the player interest will depend only
 
on  j e .
If this is true for institutional players as well and for the differences
of that player’s pure strategies, we say that the players make their choice
with no strategic bias. We don’t expect that this will be true in general.
Thus even if we start at a point with zero player engagement, we
maintain zero player interest only if the player bonding that mixes actual
 
players and institutional players  j e , is zero. This is true if the player
556 Geometry, Language and Strategy—Vol. 2

bonding matrix  is a diagonal matrix, in which case we say that the
decision process is neutral: the frequency of choice between two pure
strategies is equal.
For a neutral decision process, player entitlement means that there is
no variation of the player interest or the player engagement. We see that
the physics analogy was helpful. If there is no charge, there will be no
electric field. If there is no player engagement, there is no player
interest. Conversely, if there is non-zero player interest, then we expect
that a non-zero player engagement will develop in the focus subspace. In
general the decision process is not neutral; we allow for non-diagonal
solutions and obtain non-zero player interest.
At the still point, we think it may make sense to take the player
engagement to be zero for worldview players as a way to investigate the
effects of the player interest. This agrees with the common sense notion
that at the still point, the players need to pay attention only to their view
of the world. Also, their choice of player engagement is optimal. Though
this makes sense for the game theory players, it does not make sense for
the code of conduct players. As we shall see with the numerical
examples in the next section, in general we expect the still point player
engagement for the code of conduct players em to be non-zero and
relatively large as the value is set by Nash equilibrium values.

15.5 Focus Direction

Let’s recap what we have so far. We have enumerated the decision


process theory equations in Sec. 14.9. We have identified the initial
boundary and initial boundary conditions: specifically those associated
with game equivalency in Secs. 15.1-15.4. To obtain unique solutions,
we must complete the specification of these initial boundary conditions.
We then solve the equations by integrating them away from the
boundary. We have characterized the initial boundary as the closed focus
subspace along with the skill boundary, Sec. 15.1, which is the boundary
to the skill preference subspace on which we impose periodic
conditions.
Game Equivalency 557

Based on our numerical experiences, we are able to solve these


equations with Mathematica using its method of lines and other features
appropriate to periodic boundary conditions. This technology has
improved over time.
We have restricted our attention to a single focus dimension
appropriate to our study of two-person decision processes. We call this
single direction the focus direction. The method of lines depends on such
a direction, which we recall is a generalization from the prisoner’s
dilemma model suggested in Chap. 11, in which two players have two
strategies: either to honor a contract or break the contract. There are four
strategies and it was natural to take as inactive, the total player
preference, r1  r2 , the sum of each player preferences, Cf. Eqs. (7.18)
and (14.2).
For each player, we specified a payoff matrix that represents the
player’s view at some common contextual point. The relative player
preference z  r1  r2 was the focus direction. We generalized this
discussion in Sec. 15.1. We expect decisions to be made in which the two
players have roughly equal player preference: they make decisions in
such a way as to focus the desired (game equivalent) outcome to be at
z  0 . This was in fact the result we found for the prisoner’s dilemma.
However, this is not the general case along the focus direction: we expect
behavior along this direction to be asymmetric for some solutions to the
field equations.
The remaining skill preferences can be active. We carry out
calculations as demonstrated in Sec. 14.13. We assume values for the
scalar fields along the focus boundary with game theoretic input for the
still point. The resultant scalar fields are then specified at all other points
using the field equations. Because we make the steady state hypothesis,
including harmonic analysis of scalars representing dynamic features, we
obtain the dynamic behaviors from the resultant phasors. With this
approach, we describe a large class of models such as those in the
exercises in Sec. 12.7, in which each player has two strategies, in which
the total player preference is inactive and in which there remain three
active strategies.
We have suggested that game equivalency will help specify the initial
boundary conditions. However, we point out one difference from our
558 Geometry, Language and Strategy—Vol. 2

general discussion. Consider the prisoner’s dilemma as formulated in this


chapter. Game theory indicates four strategies. We take one of these
strategies as inactive, so in the co-moving coordinate basis, we are
dealing with only three active strategies. What we call player payoffs,
player engagements etc. must be appropriate to having three active
dimensions, not four. This means that we must perform a transformation
to relate these two views as part of game equivalency.
So the question is how to specify the scalars in this reduced space of
three at the still point. For example, we need to specify the entitlement
payoff,     , or in the notation of 3  1 models, Bα . We need the values
of the player engagements e . Though the notation is the same as in
previous sections, we now specifically mean to include not just two
worldview players, but a third institutional player. We need to treat these
players on an equal footing. We must be able to integrate their values
away from the initial boundary along the focus direction and then be able
to combine these values appropriately back into fields that relate to the
original problem statement of a two-person decision process.

15.6 Symmetric Form Transformations

The two-person decision process is defined in terms of the player payoff


components F j ab in the normal-form coordinate basis. In the process of
going from one view to another, we require additional scalar fields,
whose values we also need on the initial boundary. Without them, we
will not be able to determine the inactive metric components and their
gradients and in the end, we will not have a prediction for the evolution
of the strategic frequency choices that we could then compare with
observed behaviors.
Therefore to predict behaviors along the focus direction and indeed
throughout the entire skill subspace, we must go beyond game theory.
We believe the concepts we have developed in the last few sections for
the behaviors of the co-moving potentials will help our understanding of
the decision process and will suggest how we go beyond game theory.
Even though our fields are in three dimensions, we call Bα the
entitlement payoff. These three dimensional payoffs contribute to the
Game Equivalency 559

independent field, the composite payoff ω , through Eq. (14.26). This


determination requires that in addition to the entitlement payoff fields, we
must know the level of player engagement e through which each
payoff couples to the collective behavior. We must know the dynamics
as expressed by the decision process acceleration q away from the still
point. This reflects information on the organizational forces. Similarly
we must know the player bonding, Θ and σ αβ , which determine the
cooperative forces.
There is a new concept, expressed in terms of the seasonal vector
potential A  2 e q a that is determined from the composite payoff field
and player payoff fields, Eq. (14.30). The entitlement payoffs are
determined by the player stresses or passions, Pα , which contribute to
the player field equations.
Thus it is not enough to know only the entitlement payoffs. We need
to know more than just the competitive behaviors when we view the
system in terms of active and inactive strategies and worldview and
institutional players. We must consider some aspects of cooperative
forces.
By choosing two-person games where each player has two strategies,
we are led to the idea of a focus direction and an institutional player.
To these strategies we add two skill preferences. Together they make up
the original four strategies. Our goal is to transform from one view to
another and indicate how at least some of the needed scalars can be
determined from game equivalency.
To discuss all of the issues that might arise and to make sure we can
connect with any frame of reference, we proceed as follows. We start
with the orientation potentials Eq. (14.3), corresponding to the normal-
form payoffs Eq. (2.47) and normalized flows Eq. (14.6) that we have
used before, Cf. Sec. 7.6. The invariance of the distance measure
Eq. (14.4) provides the mechanism for transforming between different
views. This works for all of our examples in any coordinate frame, such
as in the normal-form coordinate basis, the symmetric normal-form
coordinate basis, the holonomic basis or the symmetric co-moving
orthonormal coordinate basis:
560 Geometry, Language and Strategy—Vol. 2

ˆ jk   jk
ˆ ja   jk Aak . (15.9)
ˆab  g ab   jk Aaj Abk

Specifically, the metric potentials ˆ in the normal-form holonomic


basis provide the method to tie these different descriptions together. Here
we have on the left the metric potentials in the normal-form holonomic
basis and on the right we have the potentials that describe utility
behaviors in the normal-form coordinate basis, including competitive
and cooperative payoffs. The indices  , can be active or inactive: the
requirement is that they be holonomic (Sec. 2.7).
Assuming the transformed basis from Sec. 7.5 and further assuming
that the metric potentials depend only on the active
dimensions u1 u2 u3 t , we relate the metric potentials to the
normal-form coordinate basis potentials:

j, k  1 , 2 ,
m  i1 , i2 ,
a , b  t , u1 , u2 , u3 . (15.10)
 aˆ jm   a jk A   jk  a A   a jk A   jk F
k
m
k
m
k
m
k
am

 aˆ jb   bˆ ja   a jk Abk   b jk Aak   jk F kab

In this basis, we treat the decomposition as if all the strategies were


active except the individual (worldview) strategies 1 2  . The
above notation suggests how to extend our treatment here to more than
two players and more than two inactive strategies. No result of the theory
depends on the still point value of the vector potentials Aaj  Amj  0 , so
there is no loss in generality setting them to zero; this is a consequence of
gauge invariance. Similarly there is no loss in generality in taking the
metric potentials to be equal to the Minkowski metric m (Sec. 2.6) at
the still point.
We next compute the streamline solutions of Sec. 6.2. We obtain the
following with u  u1 , u2 , u3 :
Game Equivalency 561

 uˆ jm  m jk F um
k

 tˆ jm  m jk F ktm
. (15.11)
 tˆ ju   uˆ jt  m jk F ktu
 uˆ ju   uˆ ju  m jk F uk u

Implicit in this description is that the normal-form coordinate basis and


co-moving basis are aligned at the still point using the Gram-Schmidt
process described in Eq. (14.7). We use the notation
1  x  2  y  3  z in the orthonormal frame. At the still point,
t     , we start the orthogonalization process with time along the
flow (proper-time) direction and uk  k .

15.7 Game Payoffs and Potential Field Gradients

We now have two views. One characterized by Eq. (15.11) and the other
by Eq. (15.13) below. The link between the two views is the left-hand
side of each equation. This allows us to take values in one view and
compute the consequences in the other view. The two equations have
been idealized. There will be small changes as a result of the
orthogonalization process because the energy flow is not identically
along the time direction. For the discussion below, we ignore these
effects but do take them into account in our numerical calculations.
We relate the player payoffs to the gradients of the potential fields.
For the symmetric normal-form coordinate basis, we have the same
normal-form holonomic metric potentials ˆ , which are expressed in
the form Eq. (15.9), with the below noted differences in active and
inactive variables and a different tensor A j a for the payoff potential:

a , b  t , u1 , u2 , u3
j, k  1 ,  2 ,, i1 , i2 ,
ˆ jk   jk . (15.12)
ˆ ja   jk Ak a
ˆab  g ab   jk A j a Ak b
562 Geometry, Language and Strategy—Vol. 2

We make the decomposition with the identified set of indices


u1 u2 u3 t active.
The metric tensor is independent of all other strategies. We thus have
a redefined set of symmetric normal-form payoff fields F j ab derived
from the payoff tensor. There are fewer competitive payoffs and hence
more cooperative payoffs. Therefore, of special interest will be the
gradients that can be written in terms of the normal-form payoff values,
Eq. (15.11):

j, k  1 ,  2 ,
m  i1 , i2 ,
u, u  u1 , u2 , u3
 u jm  m jk F kum . (15.13)
0   t jm  m jk F ktm
 tˆ ju   uˆ jt  m jk F tuk  F k ut  F utk
 uˆ ju   uˆ ju  m jk F kuu  F k uu  F uu
k

These normal-form “game theory” payoffs determine the still point


behaviors of the gradients  u jm for the indicated restricted set of
indices.
We also determine the still point behaviors of F k ut F k uu  . From
game theory alone however, we do not know the corresponding payoffs
F m ab for the symmetric inactive directions m  i1 , i2 , , corresponding
to the institutional players; those associated with codes of conduct. Nor
can we justify from game theory the assumption that at the still point we
have  t jm    jm  0 . For this reason, we suggest that we look
elsewhere for insight into all such quantities, though we can still use the
relationships above, Eq. (15.13).

15.8 Game Value and Entitlement Payoffs

We take the behaviors that we have determined and translate them to the
symmetric co-moving orthonormal coordinate basis. We start with the
Game Equivalency 563

j
co-moving player strategy bias field f  and co-moving player payoff
field f j  from Eq. (4.74):

f j
 
 2 q E j    E j  e E j    e E j 1  E k  Ek   
f j
  E a E b F j ab  F j tu . (15.14)
f j
     E j
   E j
 E  E F
a b j
ab F j
uu

We recall that we use the equality with the “dot”, “  ”, relationships as


those that hold assuming the frames are aligned at the still point1 and that
further, the active frame acceleration is zero, Q  0 . These
relationships hold on the initial boundary.
With three active strategies, the last equation determines the co-
moving player payoff f j   F j uu , which can be expressed in terms of
the composite payoff behaviors ω and the entitlement payoff Bα :

f j
12  B3 E j  3 E j  E a1 E b2 F j ab  F j u1u2 ,
αj 
B 3  e j ω3  F uj1u2 ,
& cyclic permutations. (15.15)

Using the concept of player entitlement above, we see that for worldview

players we have e j  0 and so the assumption here is that on the initial
boundary, worldview players make decisions based only on their
entitlement payoff.
We find that this assumption is not particularly restrictive since in
reality, non-zero player interest generates player engagement in the
neighborhood of the initial boundary. Rather, this suggests how we
isolate the player entitlement payoffs by finding regions where the player
engagement might be zero.

1
For the numerical work, we don’t make this approximation but use the vectors as
determined by the orthogonalization process and take the exact expressions for the scalars
in terms of the payoffs and other known quantities, imposing all necessary constraint
equations. The approximations here are only to provide insight into the choices that are
subsequently made.
564 Geometry, Language and Strategy—Vol. 2

What we learned from Sec. 14.9.6 is that the entitlement payoff is


determined by the gradient Eq. (14.24) for   Bα in terms of many
terms including one that depends on the player passion or stress Pα .
Decision process theory predicts that players change their utility function
for decisions as a result of applied stress or energy described here as the
player passion. The justification for this is not just the mathematics of
the field equations, but common sense that utility assigned to a
comparison of choices may in fact depend on the context.
One possible context is the desire of a player or agent to make a
change. Whether the agent may continue along that path is itself subject
to constraints, Eq. (14.25). Though we may debate the details of the
theory, we suggest that it is perfectly reasonable that the utilities should
in fact vary depending on the level of preference one has towards a
certain decision. Our choice in food may in fact depend on how hungry
we are. This dependence of payoff on preference is one key difference
between decision process theory and game theory. When this behavior is
small we obtain game equivalency away from the initial boundary.
Next, we look at the co-moving player strategy bias field f j at the
still point in Eq. (15.5) and see that it is zero along the skill preference
direction   x, y . Along the focus direction   z it depends on the
player interest (charge gradients), Eq. (15.6) and simplifies when the
players are entitled:

j  1 ,  2 ,
  1 , 2 , . (15.16)
f j
 
 I
j
 2e  
 j
 1  e e   F


j
tu

The player interest I α along the focus direction is a natural


generalization of the game theory game value.
We suggest this because the time component of the normal-form
payoff field F tuj is the game value for a symmetrized game Cf. Eq. (1.12)
and Luce & Raiffa [1957]. However, we see that a player bonding
contribution is needed to determine the normal-form player payoff. The
player interest is related to the gradient of the player engagement e by
Game Equivalency 565

Eq. (14.29). So again we have a relationship that involves scalars being


functions of strategic preferences.

15.9 Player Engagement Values

We note (Sec. 15.3) that the player engagement is a distinction that goes
unnoticed in game theory. A multiplicative factor in the payoff or utility
makes no change in the strategic content of game theory. When these
payoffs are not the same, the player engagement factors can make a
difference and in decision process theory, can change the Nash
equilibrium. In game theory, this is viewed as a change in the player
utility scales, which may be considered as not comparable.
It was suggested by Thomas & Kane [2010] that such differences
might distinguish players that are egoists from altruists. Such a
difference might provide a means to distinguish decision process theory
from game theory. We believe that the player engagement is a natural
concept and one that expands on the meaning of the value a decision has
to a player.
To measure the player engagement, we need two steps. First we need
to know the absolute utility function so that we have knowledge of the
unit of measure for the player payoffs F k ab . Then we need to measure
the product by looking at the consequence of the force Eq. (15.1) that
yields oscillatory behavior whose frequency is determined by the product
of the player engagement and the player payoff, Vk F k ab . One measures
the frequency. This approach is superior to simply setting the player
engagement to zero. The oscillations represent natural seasonal or cyclic
behaviors that characterize the decision process (Sec. 15.11).
Though we have the option of setting the player engagement to zero
for worldview players, this is not typically an option for institutional
player. Nash equilibrium determines a flow, which with game
equivalency becomes an inactive charge and hence the player
engagement for the institutional player.
When there are both worldview players and institutional players, we
use the same forms Eq. (15.7) to relate payoffs and player engagement to
the co-moving player bias f j , with the appropriate change in meaning
566 Geometry, Language and Strategy—Vol. 2

for the variables since we now have fewer active strategies and more
inactive strategies. We need information about the cooperative player
bonding components  .
We get insight on these player bonding components by looking at the
behavior of   jm . This behavior can be derived from Eqs. (15.13) and
Table 5.1 (Cf. Ex. 23 from Sec. 4.11, Eq. (4.85), specialized to the
streamline solution):

  jm  m jk E u F kum  m jk F kum
.(15.17)
  jm  2q E j Em  2  E j Em  E j Em   2 E j Em 

k
At the still point, the player payoff F um for the worldview players
determines the components of the player bonding that lie along the
mixed axes of the worldview player directions and the institutional
player directions:

j , k  1 , 2 ,
m  i1 , i2 ,
2
j
F um  2 q e j e m  2  j  m   h j  m  I j e m  Im e j
ni
e j  0 and u  focus  direction  . (15.18)
j
F um  2  j  m  I j e m
j f j 
I   2  j e
1  e  e
 j em F tuj
j
F um  2  j  m  2  e e m 
1  e e

In general, the bond shear tensor does not need to be a diagonal matrix.
It is essentially the co-moving frame player j payoff between an active
strategy u and an inactive code of conduct strategy m . The bond shear
is determined by the player game value and the player payoff between
the scale invariance direction and one of the player skill preferences.
The bond shear generates cooperative forces.
Game Equivalency 567

15.10 Cooperative Payoffs

Let’s take an arbitrary player bonding matrix at the still point and
transform to a frame in which the matrix is diagonal. In this basis, we
have defined a set of players that act as if there is no cooperation
between them. Using this form, we compute the cooperative payoffs  jk ,
Eq. (4.70): the initial metric and its gradients are diagonal, which is
consistent with our notion of what it means to be an independent player
or agent.
The diagonal nature of the player bond components maintains that
notion at all other strategic values away from the still point. If we
transform back to the original frame, these effects will not be as
apparent. Even though the player fixed frame model requires a fairly
strong constraint on the form of the bonding matrix, it still allows for
some non-trivial effects based on such coordinate transformations. We
suggest this effect becomes more complex if we generalize from the
player fixed frame model to one in which we impose only the condition
that time is an isometry: a general central frame model.
The possibility of gradients of the cooperative payoffs raises new
distinctions. We found in the previous section as well as in other
numerical calculations that the cooperative payoffs, Eq. (14.21) generate
locked behavior or strategic clustering that appears to us similar to
galaxy formation in the physical sciences. Because of this strategic
clustering, at the still point we have a higher energy density than
elsewhere. Because the energy density is assumed to be proportional to
the average pressure, we also see clustering of the average pressure. We
see a second cooperative shear effect in Eq. (14.22). This may associate
the strategic clustering to a specific player. We expect to see limits to
choice (Sec. 10.9) and limits to growth (Sec. 10.12) as we did in one-
dimensional models.
The effect of these steady state scalar fields is to determine the
decision flow V  , along with the acceleration, player bonding and
vorticity tensors along that flow. We recall that the flow of decision
processes can be envisioned as the motion of a point in the space of
preference frequencies of players or agents of the decision process,
making decisions observable and measurable. When we have
568 Geometry, Language and Strategy—Vol. 2

institutional players, Nash equilibrium determines not only the flow


along the skill directions, but along the institutional directions. These
flows are in the null space of the game equivalent payoff matrix.
We identify the cooperative force law from the general force law. If
we know the normal-form payoffs F abj for each player and assume the
metric components are locally flat, then the resultant force law, Eq.
(5.57), modified to include the constraint stresses, will generate the
vortical motions around an equilibrium direction, as described in Vol. 1.
As part of that force law, we have gradient forces due to the cooperative
payoff potentials  jk . Since the force law is the statement of the
conservation of energy and momentum in the theory, a full solution of
the field equations will automatically require the flows to satisfy the
force law. Our requirement only needs to ensure we identify values for
all of the scalars (and tensors) on the initial boundary.
The purely cooperative payoffs between the original players are
determined by the player bonding matrix:

j , k  1 , 2 ,
e j  ek  0  . (15.19)
2
 u jk    h   2 u j  k
ni u j k

The diagonal terms are analogs to the physical centripetal acceleration in


a rotating frame. Again, there is no reason that such cooperative payoffs
be diagonal at the still point. The mixed term is an example of influence
(cooperation) that one player imposes on the other that is not part of any
payoff matrix. We can always diagonalize the full set of cooperative
payoffs for every strategy direction  with a common transformation.
Even such a simple model generates mixing.
In getting to this point we recognize a viewpoint change from game
theory. We consider decisions to be more than solipsistic, though there is
a strong component of this. We do hold that decisions are based in part
on individuals pursuing their own solipsistic self-interest as reflected in
their respective payoff matrices. There is a component of entitlement.
This is similar to game theory. However, in decision process theory,
there are additional effects and forces that bring in codes of conduct,
Game Equivalency 569

cooperation and opportunity (Chap. 13.6), as well as external periodic


and possibly seasonal effects based on the seasonal payoffs. Such new
effects reflected in these additional gradients are no less important or
observable. Indeed they are essential aspects of closing the loop to obtain
a global system solution (Sec. 13.9). The cooperative effects in the
steady state models are steady state in behavior whereas the opportunity
effects are time-dependent.
We now have a practical method for carrying out the calculations that
include all of the effects along with our best guess as to which effects can
be computed using game equivalency. The one exception is how we deal
with frame rotations generated by the seasonal vector potential, which
we do in the next section.

15.11 Seasonal Player and Natural Resonances

In Vol. 1, we found that there were natural resonances in the solution


which followed from the form of the weighted sum of the player payoffs,
V j F j ab . The eigenfrequencies of this antisymmetric matrix provide the
natural resonances. We are now in a position to argue that such natural
resonances arise from the field equations of decision process theory and
are not restricted to the approximations made in Vol. 1.
We start with the concept of game equivalency, which in particular is
the requirement that the active acceleration Q  0 . In Exs. 1-5, the
student is asked to work out the consequences of this assumption. To put
the assumption in context, we look at how the reference frame rotates as
we move along a streamline using the Fermi derivatives, Eq. (4.44):

DF Ea
   Ea   1 2 I   e  Ea . (15.20)


We have inserted the player fixed frame model values, Eq. (4.49).
We learn something very interesting. The frame will maintain its
orientation along the streamline if we have small or negligible player
engagement and if we have zero composite payoffs. We propose that we
add the requirement of zero or small composite payoff to our set of
requirements for game equivalency. Even if the player engagement is
570 Geometry, Language and Strategy—Vol. 2

not zero, we see that the frame rotation is mainly along the flow
direction, which suggests that there is more flow but no change in
direction.
With this assumption, we get the following result from Ex. 2:

Vk f k    e  q  f  . (15.21)

The weighted average of the player payoffs is given by the seasonal


payoff f  , which is determined by the seasonal vector potential
A  2a e q , Eq. (14.31) in Sec. 14.9.11. Moreover, from Ex. 5 we see
that the natural resonance frequencies will be the eigenvalues of
 1 2 e  q  f  , which is the product of “charge” e  q  of the time
isometry and the seasonal payoff.
We achieve essentially the same result as in Vol. 1, with the caveat
that the payoff that matters is the one in the active space: we get one less
eigenvalue for two-person games, whose players have two strategies
each.
The seasonal player behaves like an institutional player. It makes
sense because in differential geometry, it is a theorem that for every
isometry, there will be a conserved charge, a payoff matrix (Coriolis
force) and associated gradient fields (centripetal forces or in this case
“gravity”) that describe the forces. It is thus with some justification that
we may think of the seasonal payoff as the payoff of a fictional seasonal
player. We called the inactive strategy the hedge strategy in Sec. 1.7.
We think the result also makes sense. Based on the composite payoff
being zero,    0 and Eq. (15.20), we conclude that the orthonormal
frame is not rotating. In the active space however, the system is rotating
in such a way that the active rotation component is balanced by the
weighted average of the player payoff contributions. We verify that this
is so by computing the frame effects  a bcV bV c and find that they cancel
the rotation contribution exactly.
The active rotation part is similar to the earth that is rotating. As
observers on the surface, we believe that the earth is not moving. If we
see motion of a charged particle in a magnetic field, whose motion is
exactly opposite to the earth’s rotation viewed from space above the
North Pole, then from that point in space, the charged particle will appear
Game Equivalency 571

to be stationary. Thus the fluid appears to be moving, but will be


stationary when seen from space.
We conclude that the seasonal payoff field represents a real,
observable and dynamic effect, one that we identified in Vol. 1. Now
however, we want to compute the effect in the full decision process
theory model. To solve for the seasonal vector potential, we must solve
the field equations including the gauge condition, Eq. (14.30).
The field equations still depend on the composite payoff ω , which
reflects the fact that even though we might set it to be small on the initial
boundary, it need not stay small since the player impact is a source in its
field equation that need not be zero, Sec. 14.9.8. The gauge condition has
some complexity, which is resolved in Exs. 6-8, using the covariant
gauge.
We can equally well use the seasonal gauge, Sec. 6.10, Ex. 56:

Az   z Az  0
A  2e q a . (15.22)
f    A  2e  ω  e B
q α

This sets the seasonal vector potential to be zero at z  0 . Since we have
a complete set of equations that determine the composite payoff ω and
player payoffs Bα , the seasonal payoff f is determined. We show in
Sec. 6.10, Ex. 13 that the “curl” of the seasonal payoff is zero, f  0 ,
as a consequence of the equations for the other fields. Hence the payoff
can be expressed in terms of a vector potential. With these constraints,
we only have to write the equation for the remaining two components of
the vector potential.
The first equation in the list below is imposed on the initial boundary:

 x Ay  x, y ,0    y Ax  x, y ,0   f z  x, y ,0  ,
 z Ay  x, y ,z   f x  x, y ,z  ,
 z Ax  x, y ,z   f y  x, y ,z  . (15.23)

The other two equations are differential equations that hold everywhere.
There is no ambiguity in setting the focus component Az to zero
everywhere since there are no conflicting equations. It follows from the
572 Geometry, Language and Strategy—Vol. 2

differential equations that the first equation is automatically satisfied


everywhere so these equations need to be applied only on the boundary.
We see that the seasonal vector potential is the resultant of the player
payoff fields, player engagement, player interest and cooperative
payoffs. We improve upon game theory with these new concepts and
their emphasis on how they change with strategic preference. In
particular we see that they determine dynamic behaviors, not only from
the seasonal payoffs, but through the seasonal potentials, there are frame
effects on the phasor solutions, Eq. (14.34).

15.12 Scale Invariance

We analyze the 3  1 models, Sec. 14.13, applied to decision processes


with two players who each have two strategy choices available. The
formulation covers the exercises from the literature in Sec. 12.7 that
involve two players where each player has two strategies. We treat these
models as scale invariant in which the overall scale r is inactive:
composite and player payoffs depend only on relative actions and are
independent of overall scale. We show that this assumption is consistent
with game theory in that we can specify four independent payoffs and a
game value on the initial boundary.
At the still point we assume a Nash equilibrium strategy  a , so that
this strategy vector is in the null space of each player’s payoff matrix2.
We can then define new strategy variables that put each payoff matrix in
a standard form, [Vol. 1, p. 81ff.]:

1 1
y1   r2  u2  , y3   r1  u1  ,
2 2
1 1
y2   r2  u2  , y4   r1  u1  . (15.24)
2 2

2
This assumption can be relaxed however. The assumption is however a convenient one
for gaining a preliminary understanding the results.
Game Equivalency 573

With a further transformation to the total player effort r  1 2  r1  r2 


and relative player effort u3  1 2  r1  r2  , the payoff matrix for player
j has the following form in terms of the player’s game value

T j S j
vj  :
2S j3

 u1 u2 u3 r t 
u 0 S j3 S j2 T j1 0 
 1 
u S j3 0 S j1 T j2 0 
F abj  2 j
. (15.25)
u3 S 2 j
S 1 j
0 T 3 j
2 v j m 
r T j1 T j 2 T j 3 0 0 
 
t 0 0 2vj mj 0 0 

This provides a seven-parameter characterization of the payoff:


S j T j m j  . Note that in this basis the two active strategies are skill
preferences and their corresponding “game value” components are zero.
The relative player effort however has a non-zero value for its “game
value”.
In the original basis, this payoff matrix is:

 F j ab y3 y4 y1 y2 t 
 
2
 y3 0 1 T j 2  S j1  a j c j vj mj 
 
 y4 1 2 T j
2  S j1  0 b j d j v j m j 
(15.26)
 
 y1 aj bj 0 1
2 T j
1  S j2  v j mj 
 
 y2 cj dj 1 2 T j
1  S j2  0 v j mj 
 
 t v j mj vj mj vj mj vj mj 0 

This payoff matrix is sufficiently general to cover all two-person games


with two strategies each. There are a total of 7 independent constants that
describe such games. We see that the game value is v j and that we allow
the possibility of internal payoffs.
The internal payoffs or factions account for two variables. There are
four variables that specify the game matrix:
574 Geometry, Language and Strategy—Vol. 2

aj  2
4  T  S  T  S   T  S 
j
1
j
2
j
2
j
1
1
2
j
3
j
3

b 
j 2
4 T  S  T  S   T  S 
j
1
j
2
j
2
j
1
1
2
j
3
j
3
. (15.27)
c 
j 2
4 T  S  T  S   T  S 
j
1
j
2
j
2
j
1
1
2
j
3
j
3

dj  2
T  S  T  S   T  S 
4
j
1
j
2
j
2
j
1
1
2
j
3
j
3

The remaining variable is the scale m j , which sets the relative size of
the null vector for the payoff:

 1 T j1 1 T j1 1 T j2 1 T j2 
Nj   ,  ,  ,  , m j  . (15.28)
 2 2S j 3 2 2 2
j j j
2S 3 2S 3 2S 3 

We assume the still point is Nash equilibrium, so the scale parameter will
be the same for each player as well as the other components of the null
vector. Away from the still point, all these parameters may vary and no
single null point may exist for all payoff matrices.
The assumption of scale invariance is that the effort scale r is
inactive. The conservation of effort is analogous to the idea in game
theory of a zero-sum or constant-sum game. In this case, we are making
a dynamic assumption that no scalar or tensor in theory depends on the
effort scale. This means in particular that the payoff elements
F j ar   a A j r   r A j a are independent of that scale as well as the payoff
potential A j r . We see the constraint: if chose another direction a that is
also inactive, then the payoff is identically zero. With a single inactive
strategy, the player payoff F j ar   a A j r can have any arbitrary value on
the initial boundary.
The payoff between two active strategies, F j ab   a A j b   b A j a can
take on any value on the initial boundary. The real distinction between
these payoffs and those that have one inactive direction is the spatial
dependencies. From differential geometry we determine that when there
is an isometry, the flow along that direction will be conserved, which is
not the general case. Of course on the initial boundary, we are fixed at
one point in time and so can’t observe whether a charge is conserved or
not.
Game Equivalency 575

15.13 Three Active Strategy Models

We have identified a large class of models that have three active


strategies. They cover many of the cases the elementary game theory
literature has focused on: two-person games with two strategies for each
player in which the overall scale is inactive [Williams, 1966]. See for
example Exs. 9-19.
There are other interesting cases that are covered as well. For
example we have a production and consumption model in which the
production process might have two strategic choices and the consumer
choice is limited to the amount consumed. In this case we study the
consequence of the scale as an active parameter. We leave this case as
an exercise.
We maintain our focus on the form Eq. (15.26) and its consequences.
Because we require the payoff to each player to be  v j , we assume that
at the still point the players make their choice with no strategic bias and
exclude the possibility that there are payoffs  f j x f j y  , so we are
implicitly assuming these are zero (Cf. Ex. 11 for the prisoner’s dilemma
however). The mixed strategy case with zero factions is a staple in game
theory examples of two-person zero-sum games. We see no reason to
exclude factions however (Cf. Ex. 12).
We first put each model into the frame specified by Eq. (15.24) with
the subsequent transformation to relative and total player effort, Ex. 9,
Eq. (15.36). Cf. Exs. (11)-(17). We then analyze the decision process in
that frame, starting with the initial boundary conditions based on game
equivalency. A necessary requirement for scale invariance in the steady
state model is that the inactive metric components  jk be independent of
proper-time. In particular at the still point, we must have F trj  0 ,
Eq. (15.13), which is satisfied by the general form Eq. (15.25).
This provides a formal definition in decision process theory of a zero-
sum game. This form also shows the following connections between the
game theory values in normal form and the co-moving frame values for
j corresponding to player 1 or player 2:
576 Geometry, Language and Strategy—Vol. 2

j  1 ,  2
 j e r F tuj
F  2  j  r  2 
j
e e r 
1  e e
ur

  1  e
j
f j
z

 I z j  2er  zr
j
e
 r r   F  2mv
j
tu3 j

  x, y : f j
  
 I j  2e r  3 j  1  e e   0
 r r

2 z j  r  I z j e r  F j
u3 r T j
3
. (15.29)


2 x j  r 1  e r e r  T j1 
2 y j  r 1  e r re T j
2

    e    E j  f
j
f j
 
j
xy B z  F uj1u2  S j 3
j
B x  F uj2u3  S j1
j
B y  F uj3u1  S j 2

We have simplified the expressions by assuming that the players (not the
institutional player) are entitled.
After setting values for the co-moving scalar fields at the still point,
these equations determine the seven parameters for each player, which
we connect to game theory values if available (Cf. Exs. 11-17).
The equilibrium null vector N a associated with the transformed
payoff Eq. (15.25) is T j 2 S j 3  T j1 S j 3 0 1 m j  . In particular the
flow V r along the scale direction N r and the player engagement at the
still point, e r   e E r  E r  V r , are not zero either. Although this
charge is non-zero, we must inquire about the value of the player
worldview charges e . They determine the player engagement at the still
point. In our example for the Prisoner’s Dilemma, we assumed that
initially the player charges were zero (Sec. 7.9). We justified this by
suggesting at the still point that the players will behave as if they are
neutral or uncharged.
As discussed in Sec. 15.4, our justification for worldview players is
that these flows are zero based on player entitlement at the still point. It
is not a strong argument however. It may be more relevant to note that
Game Equivalency 577

the player engagement changes rapidly away from the initial boundary in
our solutions. Since we don’t have a good argument at this time for an
initial value, zero is perhaps the best start point since in any case we will
be looking at the behaviors in a neighborhood.
In decision process theory, we specify a great deal more information
that bears on the dynamic behaviors than we would specify in game
theory. In particular our co-moving frame values may extend the usual
discussion if for example there are factions or if the components
  x , y : f j are not zero. Such changes leave the core game matrix
alone but may change the null vector. Our perspective is that the null
vector plays much less a role in decision process theory than in game
theory since the actual behavior is determined by the collective effects of
many fields, such as the seasonal payoff f .
Of course, it is plausible that under appropriate conditions the null
vector of the seasonal payoff and the null vectors of each player payoff
will be proportional. In the more general case, we still expect there to be
still points, steady state behaviors and non-steady state behaviors. Note
that Exs. 18 and 19 as forms for honoring or breaking contracts (Sec.
11.7, Ex. 1) are examples of this. All of these are consequences of
decision process theory.

15.14 Outcomes

In this chapter, the student will have learned the distinction between
strategies and preferences as well as between skill preferences and effort
preferences. The dynamic behaviors of decision process theory rest on a
set of initial boundary conditions and on time evolution equations.
The student will have learned that the initial boundary and the initial
boundary conditions are set in part by game equivalence, a principle that
relates many of the concepts and attributes of game theory to the
dynamic theory. This includes the payoff matrices from game theory,
which set the values of both competitive and cooperative fields.
There are some new terms not set by this principle: the player
engagement is new and is like “charge” in physics. The product of the
578 Geometry, Language and Strategy—Vol. 2

player engagement and the payoff matrices determines the natural


resonances of the system, leading to observable dynamic behaviors.
As in other global dynamic systems that are strongly coupled, there
are many effects to consider. Each effect in isolation may be
understandable, yet it may be difficult to appreciate how they all
interconnect. Even worse, there may be a tendency to ignore what may
turn out to be important effects because of this complexity. For this
reason, it is important to account for all of the effects and close the loop
(Sec. 13.9) by incorporating all of the equations as part of a
comprehensive approach. We do this with decision process theory.
The attainment of the outcomes of this chapter is facilitated by doing
the exercises in the following section. Based on this investment, the
student should achieve the more detailed outcomes below.
 From Sec. 15.1, the student will have learned the essential concept of
game equivalency. This includes the distinctions that we make
between strategies and preferences, efforts and skill.
 From Sec. 15.2, the student will have learned to extend game
equivalency to free fall in which no active forces cause changes to
decision flows. The concept of still point is related to this, which is a
natural extension of Nash equilibrium. The field equations require a
focus boundary on which we specify the initial values of each of the
scalars. The student should be able to explain why the still point has
zero active geometry acceleration. The student should also be able to
distinguish free fall as a special case of still point behavior.
 From the initial boundary condition values, the field equations
provide values at all other positions. The student should have some
appreciation of the initial boundary geometry and why it is steady
state. In Sec. 15.3, the student will have learned one interesting aspect
of this geometry: player engagement.
 in Sec. 15.4, the student will have learned another interesting aspect
of the steady state geometry: entitlement.
 In Sec. 15.5, we consider the behavior orthogonal to the focus
boundary, which we call the focus direction.
 In Sec. 15.6, the student will learn the technique to transform between
different views of payoffs. One corresponds to the game equivalent
view where we consider all strategies whether they are active or
Game Equivalency 579

inactive. The other corresponds to the decision process view where


one or more of these strategies is inactive.
 In Sec. 15.7, the student should understand how these results are used
to translate the game payoffs into potential field gradients. The
student will have learned that cooperative payoffs in decision theory
on the initial boundary are constrained by game equivalent payoffs.
 In Sec. 15.8, the student should understand how these results are used
to translate the game payoffs into entitlement payoffs. The student
should understand how to identify player entitlement payoffs in the
co-moving orthonormal coordinate basis.
 The student will have learned in Sec. 15.9 the relationship between
the player interest and the player engagement on the initial boundary.
The student should appreciate that the value of the player engagement
need not be zero on the initial boundary, but that it may provide a
convenient starting point.
 In Sec. 15.10, the student will have learned new aspects of decision
process theory, not covered by game theory: the cooperative payoffs.
 In Sec. 15.11, the student will have learned that with game
equivalence and zero active acceleration, decision process theory will
exhibit natural resonance behavior, with a frequency set by the
product of the player payoff and the seasonal payoff. This is
determined in terms of the seasonal vector potential and on the initial
boundary, will be equal to the weighted sum of the products of the
player engagements and their payoffs. The student will have learned
that this result assumes that the composite payoff is zero on the initial
boundary.
 In Sec. 15.12, the student will have learned how to deal with scale
invariance on the initial boundary for two-person games with three
active strategies. The student should be able to transform any two-
person game to one in which a transformed strategy is along the sum
of the original strategies. The student should be able to understand the
specific transformation chosen that results in the payoff form Eq.
(15.25). The student should be able to identify that Eq. (15.28)
provides the known condition for player decision flows.
 In Sec. 15.13, the student will have learned how to set the initial
conditions for two-person games with three active strategies.
580 Geometry, Language and Strategy—Vol. 2

15.15 Exercises

(1) Develop the equations below and explain why one would expect
helical structures based on them and how they relate to helical
structures in [Vol. 1, p. 124]. Note that the structures assume that
the payoff and most but not all metric components are constant
along the streamline as seen in Eq. (2.49), which can be re-
expressed using Eq. (4.32) showing the possibility of helical
behavior because of feedback loops both between the time
components and space components as well as between the space
components:

 E a   q  1 2   jk E j E k   Ek f k  E a   a bc E b E c ,

  E a   1  e e  E
1
2 

k 
f k   q    e E a

    1 2 Ek f k   E a    a bc E b E c ,


 E a      e      E a   a bc E b E c  . (15.30)

(2) If you ignore the bond forces  abc , show that the time dependence
of the frame equations Eq. (4.32), which we simplified in Ex. 1, Eq.
(15.30), can be put into the form (Exs. 2-5 and Cf. Chap. 6, Exs. 19-
29):

X a  E a x Ea y Eaz E a ,
1
  1  E k  Ek  ,
1  e e
q  2 e   e e 
Q  , Q  Qx Qy Qz ,
1  e e  
 2

    2 Ek f
1 k
   1 2 e  q  f   e    1 2 e  q f  ,
      1 2 Ek f k    1 2 e  q  f  , Ω   yz  zx  xy ,
Game Equivalency 581

 0 z  y   1Qx 
 
dX a
 z 0 x   1Q y  a
 X . (15.31)
d  y  x 0   1Qz 
 
 Qx Q y Qz 0 

(3) Show that the solution of Eq. (15.31) is a superposition of


eigensolutions to the following eigensystem equation:

X a   Aa    X 0    e

 0 z  y   1Qx 
 
  z 0 x   1Q y 
X 0      X 0    . (15.32)
 y  x 0   1Qz 
 
 Qx Q y Qz 0 

 Ω Ω   QQ   4  1  QΩ 
2 2
 ΩΩ   1QQ  1

  2

(4) The solutions to the eigenvalues equation in Eq. (15.32) consist of


two real eigenvalues with equal magnitude and opposite sign and
two purely imaginary and conjugate eigenvalues. The degenerate
case is when QΩ  0 in which case the two real eigenvalues are
zero. For the non-degenerate case show that for the two natural
resonance eigenfrequencies nh inh  :
X a   Aa    X 0    e ,

X 0    1
2  X     X     ,
0 0 X  0     1
2  X     X       X    ,
0 0

0

X 0  i   X 0  i   X  i   Re X 0  i  X 0 i   i Im X 0  i  ,
*  
0

A  i   A  i  A     1 2  Aa     Aa     ,
a a * a

X a  2 A a  nh  X  0  nh  cosh 1  2 A a  nh  X  0  nh  sinh nh


2 A a  nh  X  0  nh  cosh nh  2 A a nh X  0  nh  sinh 1
2 Re Aa  inh  Re X 0  inh  cos nh  2 Im Aa  inh  Re X 0  inh  sin nh
2 Im Aa  inh  Im X 0  inh  cos   2 Re Aa  inh  Im X 0  inh  sin nh . (15.33)
582 Geometry, Language and Strategy—Vol. 2

(5) In Ex. 4 we covered the case that the eigenvalues of Eq. (15.31) are
not degenerate and there are no tidal forces. We assumed that the
active acceleration Q  0 was zero: the rate of change of the flow
(in the normal-form coordinate basis) vanishes,
 E a   a bc E b E c  0 . We call this the still point. We define free
fall behavior as being at the still point with no active geometry tidal
forces  a bc  0 . Show that the eigenvector equations now simplify
since we have one solution with E a an arbitrary constant in
proper-time. Is the converse true: if this flow is constant and if the
active geometry acceleration is zero, will  a bc  0 ? There will also
be one solution with E a constant, though constrained to be in the
null space of the rotation Ω . Show that the remaining
eigenfrequency is   Ω . Thus show that the general solution for
the degenerate case is:

E a  Aa  B a sin   C a cos 


. (15.34)
E a  D a

(6) Assume that a vector function A is given that satisfies Eq. (6.65) of
Ex. 13 of Sec. 6.10,    A    f  , where A  2 e q a . We
impose boundary conditions on a surface z  0 that includes the
gauge condition  A  0 . Assume that all the boundary conditions
are known on this surface except the vector potential Az , and the
gradient  z Az . Assume that the direction z can be used to specify
Cauchy conditions for the partial differential equations. The partial
differential equations will then have a solution for each choice of
the vector potential Az . Using these solutions show that the gauge
function    A satisfies the following partial differential
equation for any initial function Az :

    0 . (15.35)

(7) Using the same assumptions as in Ex. 6, for the case of 3  1


dimensions with the notation Eq. (14.18), show that the form of the
gradient along z of the gauge function   x, y , z  is given below.
Deduce that if the gauge function and its gradient are set to zero on
Game Equivalency 583

the surface z  0 , along with periodic boundary conditions in x, y


for all z , then they are zero everywhere. Observe that the equation
below has no second z derivatives. Observe that Az  x, y,0  is an
unknown scalar function on the surface, with  z Az determined in
terms of known scalar functions on the surface. Based on these
observations, show that the equation below on the boundary surface
provides a second order partial differential equation in x, y for
this unknown scalar function Az  x, y,0  . Therefore prove that the
gauge condition and this equation on the surface are sufficient to
insure that the gauge condition is satisfied everywhere. Since
  x, y , z  is unchanged by adding a pure gauge term to
A  A    , with   cz , you should be able to show that the
right-hand side of the equation below is unchanged by the
transformation Az  Az  c . Show that this remaining gauge
freedom allows us to set the value of Az  0,0,0  to be zero at the
origin. Determine the source J from Eq. (14.31). Why is this
conserved?

J    f 
   A  J . (15.36)
 z   x, y , z    x  z Ax   y  z Ay  J z   x  x Az   y  y Az

(8) The gauge condition imposed in Ex. 7 is a linear partial differential


equation in   Az . If we choose periodic boundary conditions for
the scalar fields, then the coefficients of the differential equation are
also periodic. In general these coefficients are not polynomials. For
example, we can choose the active acceleration to be zero
everywhere on the surface, Sec. 16.13, Ex. 8. To deal with this, we
create the harmonic power series approximation to these
coefficients to every order as follows. For each harmonic function
for a particular mesh such as sin 3x , whose mesh is 3, replace it
with a polynomial in s with a power that reflects the mesh (degree)
of the harmonic, so in this case s 3 sin 3x . Treat every harmonic the
same, so sin 3y is replaced by s 3 sin 3 y . The idea is that higher
order harmonics contribute less strongly as their mesh increases. For
584 Geometry, Language and Strategy—Vol. 2

a mesh level of M , expand the coefficients as polynomials in s up


to the power 2 M . Now set the expansion variable to one, s  1 .
Thus, for successively higher mesh levels, we have an
approximation to the coefficients and can use these approximations
to compute solutions to the partial differential equation. Show that
with these assumptions on the form of the equation, we can expand
the solution as well as the coefficients in a Fourier series. Assuming
that the boundaries (for simplicity) are set by the interval     ,
for a series of  2 M  1   2 M  1 Fourier terms  mn , show that the
solution below provides an effective means to carry out the
numerical computations to obtain periodic solutions to the potential:

2  f  x  g y  h  k ,
M M
  m 2  n 2  mn  kmn  i   mf m  m,n n mn
m M n  M
M M
i  
m M n M
ng m  m,n  n g mn
M M
  
m M n M
hm m,n  n mn ,

f  f mn exp  imx  iny , g g mn exp  imx  iny  ,


h h mn exp  imx  iny , k  k mn exp  imx  iny ,
(15.37)
  mn exp  imx  iny  .

(9) Show that the coordinate transformation that goes from Eq. (15.26)
to Eq. (15.25) is orthogonal U T U  I and is given below:

F j  U T F jU,
 0 1
2
 12 1
2 0
 0  1  12 1 0 
 2 2
. (15.38)
U 1 2 0 1
2
1
2 0
 1 
 2 0 0
1 1
2 2
 0 0 0 0 1 

Game Equivalency 585

(10) If there are no factions, there are no self-payoffs. In this case show
that the usual rules of game theory give the same game value and
strategy associated with Eq. (15.28). Furthermore, show the
following relationships, where the first two follow because there are
no self-payoffs. Prove that in game theory, for mixed strategies it is
necessary that S j 3  0 . Also show that in game theory, one can add
to the game matrix components a common constant so one can
always choose values so that T j 3  0 . Finally show that the mixed
strategy solutions are explicitly as shown:

2T j1  2 S j 2  1
2 ( a j  b j  c j  d j ),
2T j 2  2 S j1  1
2 (  a j  b j  c j  d j ),
S j3  1
2 ( a j  c j  d j  b j ),
T j3  1
2 (a j  b j  c j  d j ),
S j1S j 2
vj  1
2 T j3  ,
S j3

  2S j 2   2S j2   2 S j1   2 S j1  j

N j   1 2  1  ,  1  S j  ,  1  S j  ,  1  S j  , m  . (15.39)
1 1 1
S j 3 
2 2 2
   3   3   3  

(11) Using the same transformation Eq. (15.38), show that the prisoner’s
dilemma model (Chap. 7) for player 1, Eq. (7.3), has non-zero time
components in the new basis and S 13  0 :

 0 0 1
10 2
 1
10 2
1
10 2m 
 0 0  9 10 2  9 9 
 10 2 10 2m 

F 1
ab    110 2 9
10 2
0 1 7
5m  . (15.40)
 1 
 10 2
9
10 2
1 0 1
2m 
 1 0 
 10 2m  10 2m  5m  1 2m
9 7

(12) The alternative formulation of the prisoner’s dilemma model (Chap.


7) for player 1 is Eq. (7.12). Show that with the same game sub-
matrix we have the following payoffs using the transformation
Eq. (15.38). What are the null vectors of this payoff matrix and do
they share a common null vector with Ex. 11 above?
586 Geometry, Language and Strategy—Vol. 2

 0 9
10
1
10 0  9 10 m 
9 0 1 9  9 10 m 
 10 10

F 1ab    10 1
1 0 1
10
9
10 m  ,
 
 0  10  10
9 1 0 9
10 m 
9 9  10 m  10 m
9 9 0 
 10m 10 m

 0 0 1
5 2
0 0 
 0 0  5 2 09 0 

F 1ab   5 2
1 9
5 2
0 1 9
5m  . (15.41)
 
 0 0 1 0 0 
 0 0  5m 0
9 0 

(13) Transform the game matrix from Ex. 3 in Sec. 12.7 using the
transformation Eq. (15.38) and show that it is given below. Also
show that the null vector before transformation is stated correctly
below:

 0 30 5 2 5 2 0 
 
 30 0 5 2 5 2 0 
 520 

F abj   5 2 5 2 0 170 
3m  ,
 5 2 5 2 170 0 0 

 520 
 0 0 0 0 
 3m 
1 2 1 2 
N   , , , ,m . (15.42)
3 3 3 3 

(14) Transform the game matrix from Ex. 4 in Sec. 12.7 using the
transformation Eq. (15.38) and show that it is given below. Also
show that the null vector before transformation is stated correctly
below.
Game Equivalency 587

 0 40 5 2 5 2 0 
 
 40 0 5 2 5 2 0 
 5 

F abj   5 2 5 2 0 0
2m  ,
 5 2 5 2 0 0 0 

 5 
 0 0  0 0 
 2m 
5 3 3 5 
N   , , , , m . (15.43)
8 8 8 8 

(15) Transform the game matrix from Ex. 5 in Sec. 12.7 using the
transformation Eq. (15.38) and show that it is given below. Also
show that the null vector before transformation is stated correctly
below:

 1 1 
 0 2  0 
2 2
 
 2 1 1
 0   0 
2 2
 
1 1 13 
F abj    0 6  ,
 2 2 2m 
 
 1 1
6 0 0 
 2 2 
 
 0 13
0 0 0 
 2m 
3 1 1 3 
N  , , , , m . (15.44)
4 4 4 4 

(16) Transform the game matrix from Ex. 6 in Sec. 12.7 using the
transformation Eq. (15.38) and show that it is given below. Also
show that the null vector before transformation is stated correctly
below:
588 Geometry, Language and Strategy—Vol. 2

 1 1 
 0 2  0 
2 2
 
 2 1 1
 0 0 
2 2
 
1 1 3 
F abj    0 2 ,
 2 2 2m 
 
 1  1 2 0 0 
 2 2 
 
 0 3
0 0 0 
 2m 
1 3 3 1 
N  , , , , m . (15.45)
4 4 4 4 

(17) Transform the game matrix from Ex. 7 in Sec. 12.7 using the
transformation Eq. (15.38) and show that it is given below. Also
show that the null vector before transformation is stated correctly
below:

 0 450 25 2 25 2 0 
 
 450 0 25 2 25 2 0 
 1300 

F abj   25 2 25 2 0 150 
9m  ,
 25 2 25 2 150 0 0 

 1300 
 0 0 0 0 
 9m 
5 4 5 4 
N   , , , ,m . (15.46)
9 9 9 9 

(18) Transform the game matrix for player 1 from Ex. 1 in Sec. 11.7
using the transformation Eq. (15.38) and show that it has the form
given below. Also show that the null vector before transformation is
stated correctly below. Are there other null vectors?
Game Equivalency 589

 0 0  2z 0 0 
 
 0 0  2y 0 0 
 2 w 
F 1ab   2 z 2y 0 x
m ,

 0 0 x 0 0 
 
 0 2w
 0  0 0 
 m 
 w w  (15.47)
N 1    , ,0,0, m  .
 z z 

(19) Transform the game matrix for player 2 from Ex. 1 in Sec. 11.7
using the transformation Eq. (15.38) and show that it has the form
given below. Also show that the null vector before transformation is
stated correctly below. Are there other null vectors?

 0 0 2y 0 0 
 
 0 0 2z 0 0 
 2 w 
F 2ab    2 y  2 z 0 x 
m ,

 0 0 x 0 0 
 
 0 2w
 0 0 0 
 m 
 w w 
N 2    , ,0,0, m  . (15.48)
 z z 
Chapter 16

Two-Person Decision Processes

Based on our work in the last chapter, we believe we have sufficient


grounding to specify the known behaviors. The reasonableness of this
specification may not be entirely obvious until we have worked out a few
examples. In thinking about what examples to pick, we note that game
theory practioners are familiar with a large number of two-person zero
sum games. There is a large literature around such games. Many of those
games for simplicity have been reduced to ones in which each player has
just two strategies. This seems to be a good place to start.
For historical reasons, many of these games focus on war strategies.
Over the years the focus has changed, so we think it would be helpful to
also give one example that is more of a social decision process than a
war situation.
We consider examples in which there are two players, each of which
has two strategies. We assume that the players agree to a code of conduct
in which the sum of all strategies is inactive; this leaves three active
strategies. We explore the harmonic solutions defined in Sec. 6.7, based
on the player fixed frame model. Though restrictive, we still have a wide
variety of models open to our investigation. This model framework
provides significant insight into both the vorticity free fall harmonic
behaviors and the system response harmonics (Cf. Sec. 14.1).
We organize the chapter as follows. We start with a discussion of the
types of models we might consider and focus on two such models in Sec.
16.1, one of which is a war scenario and the other a social scenario. In
Secs. 16.2-16.4, we describe a war scenario, an example we have taken
from one of the examples from Chap. 12, Ex. 5, which is the attack-
defense game from Williams [1966]. We draw heavily from Chap. 15 for
specifying the initial values. For example, the preliminary transformation

591
592 Geometry, Language and Strategy—Vol. 2

Eq. (15.25) can be used as illustrated in Ex. 15, Sec. 15.15, which also
provides the null vector (Nash equilibrium) Eq. (15.44). For this model,
we focus on game equivalence including an investigation of the free fall
behaviors. We finish our discussion of the attack-defense model in Sec.
16.5 with a discussion of gauge invariance for the seasonal vector
potential.
In Sec. 16.6 we define a social process using a work-wealth-wisdom
model that illustrates differences in our approach from game theory.
We use this model in Sec. 16.7 to illustrate streamlines when there are
only natural resonances and Sec. 16.8 when there are added forced
harmonics. This leads to a discussion of travelling waves in Sec. 16.9,
standing waves in Sec. 16.10 and striking behaviors in recurrence plots
in Sec. 16.11.

16.1 War Games versus Social Decision Processes

War games epitomize the short term concept of winning versus losing
whereas social decision processes are more about the effects of decisions
over time. This concept is not yet precise and is clearly an over
generalization. From decision process theory, there is some substance to
this idea. A necessary construct of the theory is the payoff associated
with each player, including institutional players associated with codes of
conduct. The payoff generates the concept of winning or losing and is
associated with Nash equilibrium. From this standpoint there is an
optimal strategy for behavior.
The payoff generates the concept of game equivalence in which there
is a natural or resonant frequency associated with the same payoff. We
make that concept more precise by identifying the payoff as the seasonal
payoff f  for the 3  1 dimensional models we will be concerned with
in this and the next few chapters. The idea of a resonant harmonic
(seasonal harmonic) cycle implies that there is a time scale, the seasonal
timeframe, associated with the repetition of the event; the larger the
payoff, the shorter the time scale. These are natural harmonics as
opposed to the forced harmonics that are associated with the general
expansion Eq. (14.1).
We believe that these different possibilities generate different types of
questions. We have previously illustrated one such question associated
Two-Person Decision Processes 593

with business decisions concerning the delivery schedule of software


products (Sec. 13.6). Social scenarios force us to ask questions of
sustainability and time behavior. We want to know when something will
happen next and if that behavior will continue over time. In contrast, war
scenarios help us focus on the outcome of a specific decision: whether
we win or lose a battle, for example.
We think both such questions are important and for that reason we
provide an example of each. We start with the war scenario of the attack-
defense model from Chap. 12, Ex. 5, based on the game from Williams
[1966]. We take the game aspects as reasonably well known or at least
knowable. We use that game to demonstrate how we might lock the
winning behavior in place by shortening the limits to choice (Sec. 10.9).
The war scenario highlights the thought that must go into specifying the
payoffs.

16.2 Attack-Defense Model

We recall that the attack-defense (AD) game posits that “Blue has two
installations. In normal form, he is capable of successfully defending
either of them, but not both; and Red is capable of attacking either but
not both. Further one of the installations is three times as valuable as the
other.” The attack or defense of the lesser installation is labeled “one,”
the other is “two.” The payoff matrix for blue is:

4 1
GBlue  1
10  3 4 . (16.1)
 

We highlight a number of assumptions involved at this early stage.


In arriving at the payoff matrix, the argument is made that one
installation is three times more valuable than the other. Hence to “Blue”,
a defense of the less valuable installation and a corresponding attack on
that installation is worth 3  1 units. If the attack is made on the more
valuable installation, “Blue” sees only 1 unit of value. Similarly, an
attack on the lesser installation while defending the valuable has value 3
units while an attack on the more valuable one (and defending it) is
594 Geometry, Language and Strategy—Vol. 2

worth again 3  1 units. There is an ongoing value for the existence of


the installations of 4 for “Blue” and 4 for “Red”.
These numbers make sense, but remember that in game theory any
other payoff matrix in which we change the scale or add a constant will
have the same strategic consequences. In decision process theory this is
no longer the case, so the scale factor and additional constants are
dynamic variables.
We scale the payoff for “Blue” by 110 . This changes the model. We
make the same argument for “Red” and consider the following payoff:

 4 1 
GRed  1
10  3 4  . (16.2)
 

We argue that if “Red” attacks the lesser of the two targets and it is
defended, “Red” loses 4 ; similarly if “Red” attacks the other target and
it is defended. The loss is 1 unit if the attack is made on the greater
value target and it is undefended and 3 for the other case. The payoff
matrix here is from “Red’s” point of view. For our numerical work we
also scale “Red” payoffs by 110 .
The scale factor changes the seasonal timeframe (Sec. 15.11) based
on the resonant harmonic frequency. The competitive aspects of the
attack-defense model (AD model) however, are independent of this scale
factor. In this case we have a zero-sum game. We would expect to see a
difference in the numerical results if we changed this into a constant sum
game.
One expected difference is that the limits to choice (Sec. 10.9)
narrows as we increase the scale factor and hence reduce the seasonal
timeframe. These results are different from the expectation in game
theory where nothing depends on a linear scale transformation. We
expect that the scale changes the “size” of the strategy space; the smaller
the scale the larger the effective space. We have scaled down by a factor
and get a space that has approximately unit size.
We lay out these behaviors in the following sections for the f0AD
model and our choice of parameters (see Ex. 1), which reflects purely
free fall behavior (Sec. 15.1). In the next few sections we elaborate on
this model, after which we consider a social model in Sec. 16.6. Of
Two-Person Decision Processes 595

particular interest in the social model will be a variant of the free fall
model where we add a forced harmonic behavior by means of an added
frequency component (f1WWW model). Some of the lessons learned
there also apply to the f0AD model here.

16.3 Locked Behaviors in f0AD Model

Game equivalence from Chap. 15 sets many of the parameters of the


model, but not all. In particular, there are six parameters associated with
the payoffs for the institutional player. In the original framework, these
parameters would specify the metric elements or utilities between the
institutional direction and the active strategies.
Three of these parameters can be used to set the locked behavior. In
the prisoner’s dilemma, we found that the elastic force populated a high
pressure zone, Fig. 8.1. Associated with this behavior was a
characteristic rise in the acceleration gradient, Fig. 8.2. In the inquiry
into behaviors for 2  1 dimensions we again found corresponding
behaviors in Fig. 14.3 and Fig. 14.4, which we characterized as locked
behavior. It has been found in other fields, such as the study of weather
phenomena, that there is value in focusing on areas of high pressure and
low pressure. Based on the locations of the highs and lows we then get a
more complete picture of the overall weather patterns. For that reason we
inquire further into the meaning of locked behavior.
For the attack-defense model, the parameter choices by themselves
don’t lead to the locked behavior observed in the aforementioned
references. In fact the initial choice of parameters leads to a pressure that
is negative. We see from Eq. (14.19) that terms contribute with both
signs. Our analogy to physical systems suggests that the average pressure
and the energy density should both be positive [Hawking & Ellis, 1973].
An elastic system is usually thought to be one that resists being pressed.
A positive energy density is usually thought to be related to the causality
of the system; effects occur after causes.
If our initial parameter choices generate a negative pressure, the
solution is to modify those parameter choices by increasing the size of
one or more terms that provide positive contributions. We call solutions
596 Geometry, Language and Strategy—Vol. 2

that have positive pressure and energy density, solutions with locked
behaviors. We find that we get positive values if we increase the
decision mass to md  40 . We are not forced to get a maximum
however. This gives us a minimum pressure at z  0 . To get a maximum
near the origin, we argue as follows.

Figure 16.1. Attack-Defense f0AD model q z  0,0, z 


There are several other ways we can provide positive contributions:
we can add elastic pressures    , which are currently set to zero; as with
the prisoner’s dilemma we can add strong player bonding terms ΘΘ ;
we can add rotational terms ωω , which is a classic way to balance
negative pressure or collapse; or we can add player interest terms
 I α  I α , which also generate rotational effects.

Figure 16.2. Attack-Defense f0AD model p  0,0, z 


Two-Person Decision Processes 597

All of these provide positive contributions to the pressure and would


generate locked behaviors. We have chosen to generate our numerical
examples by increasing the decision mass mentioned above and increasing
the institutional strategy bias field f 3  0, 0, 5 , which is not
specified by game equivalency. This determines I 3 through Eq. (15.6).

Figure 16.3. Attack-Defense f0AD model z  0,0, z 

We obtain the following locked behaviors for the acceleration (Fig.


16.1) and the pressure (Fig. 16.2). The key to achieving this is to focus
not on the game value but on the player strategy bias field, f 3 .

Figure 16.4. Attack-Defense f0AD model e  0,0, z 


598 Geometry, Language and Strategy—Vol. 2

Up to now we have assumed that the game for each player is simply
related to its player strategy bias field f j as in Eq. (15.29), allowing
for slight discrepancies due to the orthogonalization process. We have
assumed that the null space of the payoff vector is always related to the
Nash equilibrium.
This assumption need not always hold, though in our numerical
examples we maintain the assumption. We obtain locked behavior
consistent with this assumption because we have no information about
the game properties for the institutional player. However, we could have
changed the player strategy field for “Blue” and “Red”. This would
break our assumption about their relationship to game values. Thus, we
have reason to question that assumption.
We can choose the player strategy bias field independently from the
equilibrium flow, which leaves us free to set the form of the flow at what
we have designated as the still point. We are also free to choose the
player strategy bias field to conform to our expectations for the growth
of the player interest along the various strategic directions.

Figure 16.5. Attack-Defense f0AD model initial flow V a  0,0, z 

As an example, we could modify Eq. (16.2), so that “Red” sees a


different payoff value for attacking the two sites:
0 1
GRed  1
10  2 0 . (16.3)
 
Two-Person Decision Processes 599

Figure 16.6. Attack-Defense f0AD model initial intercept U a  0,0, z 


As a consequence we would get a different still point equilibrium based
on what each player sees:
V a  2 3 1
3
1
4
3
4 4 . (16.4)

The order of indices has “Red” first and “Blue” second. The
proportionality constant is adjusted so that the flow is a unit vector.

Figure 16.7. Attack-Defense f0AD model composite payoff


ω  0,0, z 
We return however to the f0AD model we started with. The player
bias field locks the pressure and acceleration as well as providing a
600 Geometry, Language and Strategy—Vol. 2

distinct look for the player bonding (Fig. 16.3) and player engagement
(Fig. 16.4). In particular, “Blue” and “Red” exhibit relatively flat
engagement away from the still point, whereas the institutional
engagement becomes large, even though it is close to being “entitled” at
the still point.

16.4 Decision Flow in f0AD Model

In this section we provide suitable initial starting points for free fall
solutions (Sec. 15.15, Ex. 5). Conceptually, we start with the steady
state solutions Eqs. (6.81) and (6.82). These are from Exs. 34-35
associated with linear growth and resonant behavior respectively. Such
solutions are free fall solutions (Exs. 2-6), leading to the known-
condition-based behaviors, Fig. 16.5 and Fig. 16.6. The corresponding
streamlines are linear in the normal coordinate frame active variables
u1 u2 u t .

Figure 16.8. Attack-Defense f0AD model seasonal payoff


f  0,0, z 

Though the behavior is free fall on the hypersurface z  0 , away


from the hypersurface there are acceleration effects for the active metric
components. We can make other choices that lead to a harmonic
spectrum or summation over multiple harmonics of the form Eq. (14.1).
Here however, we obtain exact solutions of free fall with just the
Two-Person Decision Processes 601

resonant harmonics (such as Fig. 16.8) that were proposed in Vol. 1


corresponding to situations in which the metric components are not
constant and the initial conditions arbitrary.
The free fall behavior, which becomes our assumption of Nash
equilibrium (Fig. 16.7), arises from the initial flows and the construction
of an orthonormal set of vectors using the Gram-Schmidt process,
Sec. 14.5:

Ej ,,a ,,t 1 2 3 x y z 


1 1.00031 0 0 0 0 0 0.0250254
2 0.000626076 1.00031 0 0 0 0 0.0250254
r 0.000626076 0.000625684 1.00031 0 0 0 0.0250254
u1 0.000221351 0.000221213 0.000221075 1.00004 0 0 0.00884783
. (16.5)
u2 0.000221351 0.000221213 0.000221075 0.0000781342 1.00004 0 0.00884783
u 0 0 0 0 0 1 0
t 0.025043 0.0250274 0.0250117 0.00883987 0.00883918 0 1.00102

Figure 16.9. Attack-Defense f0AD model frame transformation


E a x  0,0,0, 

These initial conditions set the vectors of the harmonic flow,


Eq. (15.34), Figs. 16.9-16.12. Although we obtain streamlines for the
flow that are constant, for the other frame transformations E a we get
harmonic behaviors. We also get harmonic behaviors when z  0 .
602 Geometry, Language and Strategy—Vol. 2

Figure 16.10. Attack-Defense f0AD model frame


transformation E a y  0,0,0, 
The initial attack-defense harmonic behaviors for the frame
transformations reflect the boundary conditions we impose as part of the
known conditions. We see that at time   0 , we align x with u1
(Fig. 16.9). We align y with u2 (Fig. 16.10).

Figure 16.11. Attack-Defense f0AD model frame


transformation E a z  0,0,0, 
We align z with u (Fig. 16.11).
Two-Person Decision Processes 603

Figure 16.12. Attack-Defense f0AD model frame


transformation E a  0,0,0, 

We align  with t (Fig. 16.12). In each case the corresponding


“diagonal” transformation starts at 1 . The orthogonalization process,
Eq. (16.5) and the transformation equations between gradients and the
frame transformation, Eq. (6.72) determine the remaining components.

16.5 Gauge Considerations in f0AD Model

The f0AD model generates a rotational frame of reference in the centrally


co-moving frame specified by the seasonal payoff matrix f . We see a
sample of the generated behavior in Fig. 16.8. The rotational frame
influences the transformations that we have been computing in Figs.
16.9-16.12. The behaviors are directly related to the gauge quantity, the
characteristic vector potential. Although the vector potential depends on
the gauge, quantities of interest, such as the payoffs are gauge
independent. In this section we look at the gauge properties of the vector
potential, such as the phase space Fig. 16.13.
In physical systems, we look at the phase space consisting of
coordinates and their derivatives (or equivalently their momenta) to more
clearly see patterns of behaviors. Planetary motion for example shows up
604 Geometry, Language and Strategy—Vol. 2

in a phase space plot as a circle or ellipse: its characteristic is that it is a


closed orbit. Chaotic structures show up as structures that are not closed
and may in fact be space filling.

Figure 16.13. Initial phase space for characteristic potential for f0AD model

We have chosen to focus on the gauge properties of decision process


theory because in the covariant gauge, we find an interesting property in
the solutions of Sec. 15.15, Exs. 6-8: the equations on the initial
boundary lead to elliptic partial differential equations with periodic
boundary conditions, whose solutions appear to be space filling.
This is interesting to us since we hope to find examples in the theory
where the non-linear nature of the differential equations leads to
behaviors that indicate structures such as chaos. Such hints would further
our understanding of the theory as well as the decision processes we
hope to understand.
Two-Person Decision Processes 605

In the case of the covariant gauge solutions, the elliptic partial


differential equations are not linear; currently NDSolve in Mathematica
doesn’t deal with this case if we can’t frame it in terms of the numerical
method of lines. We have instead employed a harmonic power series
approximation, 15.15, Ex. 8. We find that the solutions are not simple
closed surfaces. As an example, the analogy to a phase space plot would
be the characteristic potential a z  x, y,0  considered as the “coordinate”
along with its gradients (“velocities”),  x a z  x , y ,0   y a z  x, y ,0  .
The surface structure we find in Fig. 16.13 is clearly not a simple
closed surface. With the parameters chosen for the f0AD model, the size
is very small; nevertheless we have found similar solutions when this is
not the case (Ex. 10).

Figure 16.14. Initial characteristic potential gradients  az  x, y,0  for f0AD


model
606 Geometry, Language and Strategy—Vol. 2

By construction, the solution is a superposition of a finite number of


harmonics N , so the solution is not strictly speaking chaotic.
Nevertheless, we see the hints of behaviors that we might characterize as
chaotic. The space is partially filled. We have found that the solutions are
not sensitive to the value of N after a reasonable number is chosen, but
it is an open question whether the solution converges in the mathematical
sense as N goes to infinity. A full analysis is needed, starting with the
underlying partial differential equation.
Another question we can address is the nature of the covariant gauge
and its relationship to the seasonal gauge in which we set a z  0
everywhere. In Fig. 16.14 we have a plot of all of the gradients on the
initial surface. We start with a z  0 on the initial surface. If the gradient
along z is also zero we would have the seasonal gauge. For our
parameters, this is very nearly the case.
We see that the initial gradient along this direction is small, but it is
no smaller than the gradients along the other directions. Moreover, we
again see the hint of chaotic structure in the gradients that we saw in the
previous phase space plot.

16.6 Work-Wealth-Wisdom Model

In Sec. 16.1, we suggested that war games epitomize the short-term idea
of winning versus losing whereas social decision processes are often
about the effects of decisions over longer timeframes. We understand
what it means to win or lose; the consequences are immediate. It is less
clear for social processes. Consider how issues are framed in the media
as an example. Winning and losing are very clear. What do we listen for
relative to social decisions?
For example, it may be relevant when we hear the word propaganda.
It might be in the context of promotion and sales. What kind of forces do
they represent? Propaganda is the ability to set the narrative; it is the
ability to change a person’s mind. Success in the arena occurs over long
timeframes and so fits with our suggestion above, that social decisions
are related to long timeframes. Let’s take this a step further and inquire
about the forces that are involved.
Two-Person Decision Processes 607

How does one include such a force in a detailed theory of behavior?


Is the narrative similar to a gravitational field that draws all under its
sway, forcing everyone to move in the same direction, friend and foe
alike? Does this field owe its strength to some type of concentration of
mass or energy at some strategic location? Might we call it strategic
capital? How do we see that such a narrative effectively describes the
process? Is this narrative distinct from other cooperative forces as well as
any competitive forces?
We have suggested that the theory provides a couple of answers: first
the narrative may in fact be a part of the competitive forces where we
focus not on the win-lose aspect but on the seasonal or time variation
aspect; second, the narrative may be part of the cooperative forces that
move ones strategic position from one point to another.
To approach the answers to these questions we must overcome a
possible concern that we are using numbers to describe human behaviors.
The concern may be general. Some people are uncomfortable using
numbers to describe even physical events such as weather. Are we more
comfortable in reducing weather to a yes/no (win/lose) prediction of rain
tomorrow than in understanding the complexities of airflow, humidity
and pressure as a function of time (numbers)? The former is like a person
deciding a win-lose situation. The latter involves taking a stand on
understanding. Many of us are more comfortable with the yes/no
prediction. Yet if we are a pilot, then we need to know the weather in
terms of the detailed numbers.
What is involved in the more complex understanding? Weather
phenomena are not the acts of a capricious god but are the result of a
process involving interacting parts spread out in both space and time.
This process occurs in a continuum over space and time. What is
happening here and now is dictated by what has happened elsewhere in
the past. This is true at every level of scale, not only from a macroscopic
but a microscopic perspective.
A process view is based not just on a qualitative and discrete yes/no
understanding but a quantitative description. The quantitative description
provides the geometry; without this geometry the process is hidden. The
quantitative description provides the time dependence of the effects.
608 Geometry, Language and Strategy—Vol. 2

But, you may raise the valid objection that social phenomena are
inherently uncertain and intrinsically about decisions which generate
really discrete yes/no type outcomes. How can such events possibly be
described as part of a continuum? A choice at this instant does not
continuously flow from choices made in the past. It is a gamble and we
might in fact, if given the chance, make exactly the opposite choice if
provided the opportunity.
Yet even in physics we have phenomena at the quantum level that
from a certain point of view are uncertain and when viewed in a certain
way, appear to be discontinuous. That fact has not prevented us from
looking at them from a continuous point of view described by a
differential geometry in both time and space (Sec. 3.9). Indeed, we
suggest that game theory has provided us the perspective that social
interactions in fact can be viewed geometrically if we focus on the mixed
strategies and how they evolve in time as opposed to the pure strategies
that focus on the uncertainties [Von Neumann & Morgenstern, 1944].
We don’t focus on the actual decision, but on the mixture of decisions
that are possible at any point in time.
As an example, consider that over the last decade, wealth has
redistributed itself dramatically [Piketty, 2014 and Stiglitz, 2012]. It has
not happened discontinuously. It has evolved over time and appears to
change continuously across social strata. Some members of the middle
class have become wealthy whereas others have become poor. The
changes reflect a process, not a capricious set of changes. The process is
even more in evidence over long time scales (centuries). This example
might help focus our attention.
Suppose we analyze the flow of wealth and assume for simplicity it is
distributed between two distinct populations. If the populations are
valued similarly and if the interaction is zero-sum, we expect each to
receive the same payoffs in the sense that what one wins, the other loses.
But what if one population believes their payoff, if they win, is much
higher than the other population? We still achieve a zero-sum game if
we balance the product of the population and value for each side.
If it is agreed that one side is 10 times more valuable, then this works
if the other side has 10 times the population. What matters here is the
interplay between competition and propaganda, since there may in fact
Two-Person Decision Processes 609

be no objective reason for the valuation other than a (possibly enforced)


agreement. This example supports the geometric view, if learning how
the flows change as a function of wealth distribution gives useful insight
into wealth inequality.
So we return to the question of propaganda and narrative, which we
view as a question of process. In the above example, the valuation of
wealth is felt equally by both populations, yet it may not be factually
based. From a theoretical point of view that is really fine. We are not
establishing the “truth” of the valuation, but the outcome given that both
sides adopt this “truth”.
Whether this is a useful exercise is ultimately a question of
measurement and quantitative analysis, not to mention ethics. In a
differential geometry theory of decision processes, we look for
confirmation in data (behaviors) that highlight the existence of the
process. For example, for weather predictions, it is not enough to predict
rain versus not rain; rather we must predict in addition behaviors that
change continuously with space and time like air flow. That insight
allows us to predict when catastrophic phenomena such as hurricanes
might occur. Therefore, for social behaviors we must look for flows, as
an example, which change continuously with strategic position and time.
Sufficient insight might help us predict similar catastrophic phenomena.
We seek to gain understanding from two distinct directions. We look
at data to be convinced social behaviors are in fact geometric processes
and we look at results of theoretical simulations to be convinced that
geometrical processes might explain such data. At some point we hope
that these two approaches will meet, even though we are not there yet.

16.7 Streamlines in f0WWW Model

Let us put these ideas into the format of decision process theory. We
imagine a model in which there are two players, the workers and the
wealthy: player 1 and player 2 respectively. The players have entered
into a contract for increasing the general welfare by innovation and work.
Possible strategies for the wealthy population are:
 Honor the contract by increasing the wealth through reinvestment to
create innovation and economic growth: HONOR
610 Geometry, Language and Strategy—Vol. 2

 Break the contract by transferring wealth from the worker through


rents without reinvestment: BREAK
Possible strategies for the worker population are:
 Honor the contract by increasing global wealth through labor:
HONOR
 Break the contract by not working as agreed: BREAK

For the worker, we imagine that their game matrix is:

work / wealth HONOR BREAK


G worker  HONOR 0  1100 . (16.6)
BREAK  100
9 0

For the wealthy, we imagine that their game matrix is:

wealth / work HONOR BREAK


G T
wealthy  HONOR 0 9
100 . (16.7)
BREAK 1
100 0

The model is a zero-sum game.


We deviate from game theory by formulating the equivalent game so
that it takes into account both value and population. For each player, we
propose the following form for the payoff matrix:

 0 GT vWealth mg 
 
F G 0  vWork mg  . (16.8)
 vWealth m vWork m 0 
 

We interpret vwork as the game value for the worker population and
vWealth as the game value for the wealthy population.
The payoff matrix has a null vector, which we call Nash equilibrium:

 f Wealth Y 
 
V Nash
  f Work X  . (16.9)
 mg 
 
Two-Person Decision Processes 611

Without losing generality, we assume the vectors X and Y each have


components that sum to unity, and assume that the game value vg is
defined in the usual way, based on the max-min rules (Sec. 12.3.1):

GX  v g X
. (16.10)
G T Y  vg Y

The condition that V Nash is the null vector then leads to the following
conditions:

vWealth   f Work v g ,
vWork  f Wealth v g ,
vWealth v
 vg    Work  vWealth f Wealth  vWork f Work  0 . (16.11)
f Work f Wealth

We have a new population value rule that the product of the population
and the value is zero-sum.
For numerical work, we assume vWealth :  vWork   9 :1 , so with these
assumptions, we have f Work : f Wealth  9 :1 . The number of workers
outnumbers the number of wealthy by a factor 9, the same factor by
which the game value of the wealthy outnumbers the worker game value.
As a result, we will have the payoff matrix for the worker (and an
identical one for the wealthy):

 9 vg 
10
 0 0 0 9
100 
 mg 
 9 v
10 g

 0 0 1
100 0 
 mg 
 1 v 
 0
10 g 
F 1,2 ab  1100 0 0  . (16.12)
 mg 
 
1 v
 9 10 g 
  100 0 0 0 
mg 
 
 910 vg 9
10 vg 1
10 vg 1
10 vg 
  0 
 m mg mg mg 
 g 
612 Geometry, Language and Strategy—Vol. 2

With this payoff, a game value vg   9 1000 and an inertial parameter


mg  4 , we have the null vector for each payoff:

 1Nash   2 Nash   1100 9


100
81
100
9
100 4 . (16.13)

Based on our assumptions, the sum of the preferences for the worker
(third and fourth components) is a factor 9 above the sum of preferences
for the wealthy (first two components). This is illustrated in Fig. 16.15.

Figure 16.15. f0WWW model initial flows

We thus have a Work-Wealth-Wisdom (WWW) model, where


wisdom is the collective wisdom of the players capturing their history of
what has happened. The theory generates the corresponding
consequences. This model allows us to explore both social and technical
aspects of decision process theory.
From the initial normalized flows, Fig. 16.15, we see that the
inequality flow for the worker and the wealthy,
V  2 V
u 1 work
V wealth
  0.1095... is positive. The conserved flow
 
V r  1 2 V work  V wealth  0.1369... is initially greater than the
difference. This reflects the wealth inequality of the initial conditions.
The oscillations are due to the natural resonance of the seasonal
player, Sec. 15.11. We were unable to find stable solutions unless we
chose a fairly small game value, as indicated in Eq. (16.12). This
correlates with a long seasonal timeframe in Fig. 16.15.
Two-Person Decision Processes 613

During the first cycle we note that the inequality flow exceeds the
conserved flow, which implies that the rate of growth of the wealth
population starts to decrease. As time progresses, this changes and the
rate of growth of the two populations become almost equal: the growth
of the inequality flow becomes zero. Neither condition is stable; rather
we alternate between these two extremes. We do see a trend however:
the inequality continues despite the fluctuations and tends to deepen if
you look at the effect on the overall preferences.
The figure also exhibits competitive information based on the natural
resonance structure. The initial normalized strategic flows in this basis
are:

V u1
,V u2 ,V u ,V r   0.1394..., 0.01549...,0.1095...,0.1369... . (16.14)

With the time component, the flows are normalized to unit length,
gabV aV b   jkV jVk  1 . The net work-strategy

V u1  1
2 V HONOR
work  Vwork
BREAK
,
represents the difference of the normalized worker strategies to HONOR
or BREAK the contract. It is positive and proportional to the Nash
strategy difference for workers, 72 2 200 from Eq. (16.13). The net wealth-
strategy

V u2  1
2 V HONOR
wealth  Vwealth
BREAK
,
represents the difference of the normalized wealth strategies to HONOR
or BREAK the contract. It is negative and proportional to the Nash
strategy difference for wealthy,  8 2 200 from Eq. (16.13).
We see the interesting dynamic in Fig. 16.15 that during the phase in
which the workers more strongly honor the contract, the wealthy more
strongly focus on rents. Conversely when the workers adhere less to their
contract, the wealth strategy is to invest and rent about equally.
614 Geometry, Language and Strategy—Vol. 2

16.8 Forced f1WWW Model

Based on our initial parameters (Ex. 7), the wealth inequality generates a
seasonal natural resonance with a large seasonal timeframe. The frames
rotate based on this same constant: see for example Fig. 16.16.

Figure 16.16. Frame rotation in the f0WWW model

It is not just our choice of parameters. It has been shown that the type
of wealth inequality effects we have been discussing oscillate over
periods of time spanning multiple decades [Piketty, 2014]. In addition,
there may be short term oscillations not directly related to the
competitive effect under discussion. In decision process theory, such
effects show up as forced harmonic effects and are set by the initial
boundary conditions, Sec. 14.9.14. We describe such effects in this
section.
Two-Person Decision Processes 615

Figure 16.17. f1WWW model initial flows

We explore the effects by superposing a small oscillating term


(Ex. 11) to the seasonal natural resonance of the WWW model from the
last section. At the initial time we start with the same flows as before, but
based on this assumption, will see small oscillations, Fig. 16.17. The
oscillations are significantly shorter than the seasonal timeframe. The
behavior is not dictated by the competitive nature of the WWW model or
by the wealth inequality. Rather it is set by the initial conditions: it is set
by the forced harmonic  and the values of the phasor preferences
U a  x, y,0  and V a  x, y ,0  at z  0 .
The seasonal harmonic is clearly visible if we extend the time
interval; there is a slight hint however of its effect in Fig. 16.17. For the
model parameters chosen, as the worker moves towards honoring the
contract less, we see the wealthy moving towards more rents. This is
quite different from the seasonal behavior that was generated by the
competitive force.
616 Geometry, Language and Strategy—Vol. 2

Figure 16.18. f1WWW model energy flow velocity

Figure 16.19. f1WWW model time flow


velocity
Two-Person Decision Processes 617

This decoupling of the time behavior from the competitive behavior


we believe is important and helps us understand why it is difficult to
relate the dynamics to an underlying process.
Next, we compare two frame transformations that relate to the
relative flow of the co-moving frame relative to the normal frame. This
may generalize. We start with the active flows, V a  E a  , which we
characterize as the frame transformation components of the energy flow
velocity E a  in the co-moving orthonormal frame, Fig. 16.18. We
consider only the components that are not conserved: the net-work-
strategy, the net-wealth-strategy and the inequality flow.
These frame transformations, at a nominal value of the coordinates,
characterize one view of transforming from the initial frame to the co-
moving frame. Another characterization is provided in Fig. 16.19. Before
studying these two characterizations, we again note that the structure of
the frame transformations in this figure is much different from those in
the f1WWW model where the forced harmonic effects are small, Fig.
16.16.
Each of these figures provides some information about the difference
between the normal frame and the co-moving frame. We describe the
model in the normal frame. For the f1WWW model we interpret the
inequality assumptions in this frame of reference. We identify the
strategies as flows, so there is a direction of flow and a magnitude. If we
were to move along the flow with that magnitude, we would be in the co-
moving frame. This is the interpretation of Fig. 16.18. Because the
strategy-flow changes in time, the transformation to that frame changes
with time as well.
Another way to view the flow of energy is to view the “flow of
time”. Each instant of time in the normal coordinate frame is a
hypersurface of strategies. As part of decision process theory, we assume
that we can measure each of the strategies on this hypersurface. A
surface is specified by a scalar field t  u  in terms of time and the
strategies in the normal frame. The direction “normal” to the surface is
given by the gradients   t   t , which is non-zero only if the gradient is
along the time direction. This defines a covariant vector, which can be
transformed to the co-moving frame: E    t , E    t , E   t  , with
components along the flow, the active directions and the inactive co-
618 Geometry, Language and Strategy—Vol. 2

moving directions. The vector in the co-moving frame is therefore


E t ,  e E t , t E t  . The frame transformation provides us the time
flow velocity E  of the surfaces of constant time as seen in the co-
moving orthonormal frame. In Euclidean geometries, the time flow
velocity and the energy flow velocities are proportional. This is not the
case in non-Euclidean geometries, as we see by comparing the figures.
Also see Ex. 14.

16.9 Travelling Waves

For Robinson Crusoe economics, we investigated the travelling wave


behaviors in Sec. 10.8. In this chapter, we find similar behaviors in the
forced attack-defense model and in the f1WWW model. In this section
we explore such behaviors for the f1WWW model. With the parameters
we have experimented with, we see a strong limit to choice (Sec. 10.9)
and see little of the distance oscillations of the Robinson Crusoe
example. We analyze the f1WWW models in a similar way.
We start with the wave Eq. (14.34) in the central frame in 3  1
dimensions. As before we look for phasor solutions characterized by the
frequency  , which we write using complex numbers:

x a  x, y , z ,   Re  x a ei 
. (16.15)

x a  x, y , z ,   Re Ha  x, y  ei
a
 x, y
ei  z 
This solution, valid up to second order in a power series expansion
around z  0 , allows us to interpret the general solutions as either
travelling or standing waves. The waves are characterized by a transfer
function Ha  x , y  e i  x , y  that is analogous to that in electrical
a

engineering, with a transfer modulus Ha  x, y  and a transfer phase


a  x, y  .
Here we envision a travelling wave moving along the z direction that
measures the relative population of the workers to the wealthy, though
we could equally well focus on any of the other two directions or any
combination of the three. We expand the solution around the origin z  0
and thereby determine the parameters:
Two-Person Decision Processes 619

  a   a  i  a ,
 2     q  4i e q a  
 2 e 2 q 1  e e  4a a   2i e q a    q   0 . (16.16)

The solution will in general have both an “attenuation” component  a


and a velocity component  a . We make these identifications by
looking at the solution for the parameters evaluated at the origin,
assuming that the gauge at the origin is zero, a z  0 :

a  1 2   z  q z 
. (16.17)
a   2 e 2 q 1  e e   1 4   z  q z 
2

The field equations allow the attenuation to be both positive and


negative, since energy can be taken or given back.
The attenuation is not analogous to resistance, but analogous to
capacitance or inductance in electrical engineering. The energy is not lost
but stored. In this case it can be stored in a number of ways including the
analogy of electric and magnetic fields as well as in gravitational fields.
The general solution is based on taking the real part of the coordinate
scalar Eq. (16.15):

Re  x a ei   Ua  x, y , z  cos  Va  x, y , z  sin


Re  x a ei   Ha e z cos   a z  a 
a

. (16.18)
Ua  x, y , z   Ha e z cos  a z  a 
a

Va  x , y , z   Ha e z sin  a z  a 


a

Thus in terms of the general solutions, for each frequency component we


have the reduced transfer function with reduced transfer modulus
H a  H a e   z and reduced transfer phase shift  a   a   a z . This
a

result is very similar to the analysis of circuits and transmission lines in


electrical engineering. The possibility of a non-zero reduced phase is the
possibility of effects that are analogous to inductance and capacitance in
620 Geometry, Language and Strategy—Vol. 2

circuits. The reduced phase shift and the transfer modulus are the
outcomes of the theory.
This analysis starts with parameters identified at the origin and
determines transfer functions that characterize the solution around the
origin. We can reverse the process and start with a general solution, Eq.
(14.35) and compute the modulus of the solution and its phase:
Ua  x, y, z   Ha cos a
Va  x, y , z    Ha sin a
. (16.19)
a  a  a z
a
Ha  Ha e z

These definitions give us a reference point for discussion whether


solutions have travelling waves or standing waves: standing waves will
have a phase that is missing the growth along z .

Figure 16.20. Transfer function phase gradient for


travelling wave inequality flow in f1WWW model
Travelling waves provide evidence that solutions have contributions
that propagate from one point of strategic space to another with a
specified velocity. It is a property of the theory that such possibilities
exist and a constraint on the theory to find evidence in support of this.
We want to show that these possibilities do occur.
Two-Person Decision Processes 621

In Fig. 16.20, we see the phase for both the seasonal harmonic
component and the forced harmonic component in the f1WWW model.
The seasonal harmonic component is a standing wave whereas by
construction, the forced harmonic component is a travelling wave
moving to the right with a a  0.92.. based on the input parameters and
Eq. (16.17). We compute the velocity in any direction for a frame wave
as  a , so we expect the gradient of the transfer phase to be zero for
the standing wave and 0.92 for the travelling wave, which is what we
see. Note also that the reduced transfer phase shift is not zero for the
forced harmonic.

Figure 16.21. Travelling wave inequality flow of forced harmonic component


Because we have a restricted spatial component, it is a little harder to
see the travelling wave component in the inequality flow, Fig. 16.21, but
it is there. We see that the peak moves somewhat to the right as we
increase the proper time. The analogy is that we have a calm situation,
such as a calm lake, on which we now impose a small breeze.
We suggest that we see the zephyrs on the water indicating the
propagation of the effects of the wind on the water. One way to capture
622 Geometry, Language and Strategy—Vol. 2

this behavior is to look at a frame transformation that is initially small,


such as E t z , Fig. 16.22, which is the time flow velocity along the z
direction as seen in the orthonormal co-moving frame (Cf. Fig. 16.19).
We see slight ripples that correspond to the travelling wave of the
example. We are looking at very long time intervals compared to the
forced harmonic frequency. We see the behavior of the seasonal wave,
which has much in common with tidal effects of waves in the ocean.
Superposed onto this tidal effect are the ripples that we have added. If we
increase the breeze, we expect the effects to increase correspondingly
(e.g. Fig. 16.46).

Figure 16.22. Frame transformation for travelling wave in f1WWW model

16.10 Standing Waves

We may think of a standing wave as a superposition of two travelling


waves that move in opposite direction with the same amplitude. We
would then expect that the velocity of the resultant wave would be zero
Two-Person Decision Processes 623

and so the gradient of the phase would also be zero. We have constructed
such a superposition at z  0 .

Figure 16.23. Transfer function phase gradient for the


standing wave in the f1WWW model

The resultant wave phase gradient, Fig. 16.23 shows that both the
seasonal and forced harmonic components are standing waves as
expected. This is further confirmed in Fig. 16.24. We have imposed the

Figure 16.24. Standing wave inequality flow of forced


harmonic flow in f1WWW model
624 Geometry, Language and Strategy—Vol. 2

condition on the forced harmonic component that it be a standing wave.


For the seasonal harmonic component, we have imposed instead the
condition that the total active acceleration be zero.

Figure 16.25. Seasonal phase gradient in f1WWW model

This does not necessarily imply that the gradient of the phase is zero.
Indeed, the phase gradient, Figure 16.25, indicates that although the
travelling wave components along z are small, they are not necessarily
small along the other two directions.
When we turn to the large time scale behavior, we see less structure,
Fig. 16.26 than in Fig. 16.22. We expect standing waves to exhibit the
structure of travelling waves, including reflections, after all transient
behaviors have died out. Once the transients have died out, we may no
longer see what is travelling and what is being reflected. We think it
makes sense that there is correspondingly less structure exhibited. It is
Two-Person Decision Processes 625

the difference between initiating a wave behavior and looking at the


wave behavior after the steady state behavior has settled in.

Figure 16.26. Frame transformation for standing wave in f1WWW model

16.11 Recurrence Plots

By looking at recurrence plots as we did in Sec. 9.9, we can investigate


properties of data and theory that relate to the various time scales of
causal interactions. For the f1WWW model, we show an example of such
a plot in Fig. 16.27. The frequency of the seasonal contribution is
0.0163... compared to the forced component frequency 1.00... The time
for one cycle is 2  for each frequency, or approximately 6.28 and
385. respectively. Thus in Fig. 16.27, we are considering approximately
one cycle for the seasonal harmonic.
626 Geometry, Language and Strategy—Vol. 2

Figure 16.27. Standing wave recurrence plot for inequality flow in


f1WWW model

We expect that the structure will change as we change time domains.


We expect additional structure if we superpose a number of frequencies
as opposed to the two frequencies here, such as in Ex. 19 and Fig. 16.51.
We anticipate that the more closely spaced the frequencies, the greater is
the possibility for generating not just behaviors that have different
periods, but behaviors that are chaotic as we speculated in Sec. 9.9.

16.12 Outcomes

In this chapter we have examined two-person decision processes, starting


in Sec. 16.1 with a discussion of two main types of models: competitive
models such as war scenarios, Sec. 16.2 and social models, Sec. 16.6. As
an archetype of the war scenario, we consider the attack-defense model,
Two-Person Decision Processes 627

starting in Sec. 16.2. As an archetype of a social scenario we consider the


work-wealth-wisdom (WWW) model in Sec. 16.2. The student will have
learned how to apply decision process theory to these two quite different
scenarios in which there are three active strategies. Of particular interest
are the locked behaviors, Sec. 16.4, which the student should recognize
as being related to limits to choice, Sec. 10.9. By setting parameters in
which there is a strong pressure high, we limit the size of the space
available for choices.
The student will have learned to see causal behaviors in decision
processes that are analogous to standing and travelling waves in fluids:
the student should be able to distinguish the standing wave of the
seasonal harmonic from a forced harmonic set by an initial condition.
The student should note that with only three active strategies, there is a
single seasonal harmonic; however in the general case with more active
strategies, there will be more seasonal harmonics. There will be a
discrete number of such harmonics set by an eigenvalue equation.
The attainment of the outcomes of this chapter is facilitated by doing
the exercises in the following section. Though not essential for a general
understanding, there is benefit for a detailed understanding to replicate
the numerical examples given. Based on this investment, the student
should achieve the more detailed outcomes below.
 From Sec. 16.1, the student will have learned the distinction between
competitive scenarios and social scenarios and how they are
approached in decision process theory.
 From Sec. 16.2, the student will have learned how to take a game
from, for example, a classic treatment by Williams [1966] and set the
various parameters for such model’s evaluation. In this case, the
student will have learned to set the parameters for the attack-defense
(AD) model.
 From Sec. 16.3, the student will have learned how to create a locked
behavior by adjustment of parameters for the AD model.
 From Sec. 16.4, the student will have learned the decision flow results
for the AD model. Of particular note are the seasonal payoffs that
result in this model for game equivalency and the resultant seasonal
harmonics.
628 Geometry, Language and Strategy—Vol. 2

 From Sec. 16.5, the student will learn how to set the gauge for the AD
model. The student will have learned the consequence of picking the
covariant gauge is the necessity of solving a set of differential
equations on the initial surface. This is accomplished using the
harmonic power series approximation.
 From Sec. 16.6, the student will have learned that game theory ideas
can be extended in the context of decision process theory to a
problem of two players who get different payoff values. This leads to
the work-wealth-wisdom (WWW) model.
 From Sec. 16.7, the student will have learned about the streamlines
that result from the seasonal harmonics in the WWW model.
 From Sec. 16.8, the student will have learned how to add forced
harmonic effects to the WWW model and how such a f1WWW model
with a single forced harmonic behaves.
 From Sec. 16.9, the student will have learned how to see the travelling
waves in the f1WWW model. They can be seen by looking at the
transfer function for each harmonic component. The magnitude and
phase of each component characterizes the behaviors. The phase
contains the information about the travelling wave velocity.
 From Sec. 16.10, the student will have learned how to create standing
waves in a f1WWW model. In this case the wave does not move,
being a superposition of waves moving in opposite direction with the
same amplitude.
 From Sec. 16.11, the student will learn the value of using recurrence
plots to see different harmonic contributions. The student should be
able to see how such plots could also indicate chaotic behaviors, such
as suggested in Sec. 9.9.

16.13 Exercises

(1) To set the parameters of the attack-defense (f0AD) model, show that
we need to set the payoffs as in Eqs. (16.1) and (16.2). We take the
initial flow to be the Nash equilibrium Eq. (16.4), which assumes a
decision mass of md  40 . For the inactive flow or charge for each
player we take V 1  V 2  1 . Do we have enough information to
Two-Person Decision Processes 629

determine the symmetric payoff matrices for each player, F j ab and


the initial frame transformation, Eq. (16.5)? If not, what information
is missing?

Figure 16.28. f0AD model streamline surface for


x y0

Assume that we look for solutions that are periodic in the transverse
directions  Lx  x  Lx and  Ly  y  Ly , but not in the z
direction. For this direction, there will be limits to choice,
zlow  z  zhigh that will come out of the initial variable values. To
specify the steady state scalars, show that we must at least provide
the following initial values, where the values have been chosen to
match those of the figures in this chapter and we use the notation
from Chap. 14 (here   x, y, z  represents any tensor field):

  5,   1,
Lx  5, Ly  5,
   Lx , y , z     Lx , y , z    x,  Ly , z     x, Ly , z  ,
ω  B 3  0 0 0,
630 Geometry, Language and Strategy—Vol. 2

f 3 x  0, f 3 y  0, f 3 z  5,
z11  310 , z 22  4 10 , z 33  510 , z12  0 . (16.20)

(2) As an application of the attack-defense model of Sec. 16.1, the


streamline surfaces (Figs. 16.28 and 16.29) were obtained using the
steady state solutions with appropriate boundary conditions and
only the common free fall components, Secs. 15.1 and 16.4.

Figure 16.29. f0AD model streamline surface


x  y 1

Can you explain the behaviors?


Two-Person Decision Processes 631

Figure 16.30. f0AD model frame transformation


E u z  0,0, z, 
(3) For the Attack-Defense model, the frame translations (Figs. 16.30-
16.32) contain a great deal of information. The transformation
E u z  0,0, z,  is set to unity by the initial conditions. That it varies
in both the inequality direction and in proper time is a consequence
of the dynamics. The other two transformations start out at zero

Figure 16.31. f0AD model frame transformation


E u1 z  0,0, z, 
632 Geometry, Language and Strategy—Vol. 2

Figure 16.32. f0AD model frame transformation


E u2 z  0,0, z, 
initially and take on non-zero values as a consequence of the
dynamics. Can you identify the standing wave behaviors in each of
these curves and how they relate to the seasonal timeframe?

Figure 16.33. f0AD model pressure contours along


the streamlines
(4) We gain additional insight into the behavior of decision processes
by studying the contours of constant pressure, analogous to the
insight gained in weather forecasting, as well as studying contours
Two-Person Decision Processes 633

of constant acceleration. Along a streamline, the pressure is a


constant. Nevertheless, we see dynamic behavior in the normal
coordinate frame, Fig. 16.33.

Figure 16.34. f0AD model acceleration potential


contours along the streamlines

Note that the player inequality appears to oscillate with the seasonal
harmonic. As a function of time, describe the forces that these
contours represent.
(5) There is a similar behavior with Ex. 4 in Fig. 16.34 for the
acceleration potential. Here the potential oscillates with the seasonal
harmonic frequency. Why is there a maximum? What does this say
about the forces? As we switch from normal coordinate frame to
orthonormal frame, it is helpful to see the behavior of the streamline
contours u t compared to  z   , Fig. 16.35. What happens
when the streamlines become bunched?
634 Geometry, Language and Strategy—Vol. 2

Figure 16.35. f0AD model streamline contours for


free fall

(6) We investigate the behavior of the player charges as functions of


time and the relative player effort u  u3 , the player aggression.
When it is positive, player 1 is more involved than player 2.

Figure 16.36. f0AD model defense (blue) charge


Two-Person Decision Processes 635

When it is negative, the situation is reversed. For the “blue” defense


player in the attack-defense scenario, we see that there are limits to
the size of the player aggression, Fig. 16.36.
We see a similar limitation for “red” attack in Fig. 16.37. What are
the sources of these limitations? The contours are for fixed values of
x y . Does the situation change if we vary these values? The
behavior for the code of conduct is quite different, Fig. 16.38.
Explain the source(s) of the differences.

Figure 16.37. f0AD model attack (red) charge


636 Geometry, Language and Strategy—Vol. 2

Figure 16.38. f0AD model code of conduct


charge

(7) To set the parameters of the work-wealth-wisdom (WWW) model,


show that we need to set the payoffs as in Eqs. (16.6) and (16.7).
We also need the initial flow, which we take to be the Nash
equilibrium Eq. (16.13), which assumes a decision mass of md  4 .
For the inactive flow or charge for each player we take V 1  V 2  1 .
Do you have enough information to determine the symmetric payoff
matrices for each player, F j ab and the initial frame transformation
below?
Ej ,,a ,,t 1 2 3 x y z 
1 1.03684 0 0 0 0 0 0.27392
2 0.0723665 1.03431 0 0 0 0 0.27392
r 0.0361832 0.0337401 1.00812 0 0 0 0.13696
u1 0.036843 0.0343553 0.016474 1.00829 0 0 0.139457
. (16.21)
u2 0.00409366 0.00381726 0.00183045 0.00183362 1.0001 0 0.0154953
u 0.0289466 0.0269921 0.0129432 0.0129656 0.00142864 1.00504 0.109568
t 0.289466 0.269921 0.129432 0.129656 0.0142864 0.100504 1.09568

If not, what information is missing? Assume that we look for


solutions that are periodic in the transverse directions  Lx  x  Lx
and  Ly  y  Ly , but not in the z direction. For this direction,
Two-Person Decision Processes 637

there will be limits to choice, zlow  z  zhigh that will come out of
the initial variable values. To specify the steady state scalars, show
that we must at least provide the following initial values, where the
values have been chosen to match those of the figures in this chapter
and we use the notation from Chap. 14 (here   x, y, z  represents
any tensor field):

  5  1
Lx  5 Ly  5
   Lx , y , z     Lx , y , z    x,  Ly , z     x, Ly , z 
. (16.22)
ω  B 3  0 0 0
f 3 x  0 f 3 y  0 f 3 z  10
z11  310 z 22  4 10 z 33  510 z12  0

(8) In the numerical work (e.g. Ex. 1), the assumption has been that we
can set the active acceleration to zero at a point. Take this idea one
step further. In a consistent way, set the active acceleration to zero
on the initial surface. Show that this implies that the acceleration
and its gradient along the focused direction are determined on the
surface:

q pulse   1 2 ln 1  e  e 
. (16.23)
 z q pulse    e I z  e  z e  
pulse

(9) In setting the active acceleration to zero on an initial surface, for


numerical purposes you might think that Eq. (16.23) complicates the
analysis using the harmonic power series approximation (Cf. Sec.
15.15, Ex. 8). By investigating the numerical results in a variety of
models for different mesh sizes, show that this is not the case. For
example, we have computed the following table of results with an
assumed set of initial conditions for the known conditions, using
gauge invariance to set the initial value a z  0,0,0  (on the initial
surface z  0 ) of the characteristic potential to zero. These
638 Geometry, Language and Strategy—Vol. 2

parameters characterize the limit cycle surface in the three


dimensional space of  x a z  x , y ,0   y a z  x , y ,0  a z  x , y ,0  .

Figure 16.39. Gauge condition surface with mesh  4


The mesh is the order of the highest harmonic term. For the WWW

Figure 16.40. Gauge condition surface with mesh  10


Two-Person Decision Processes 639

model we have the following table. Based on these numbers, is the


process converging?

Table 16.1. Mesh Analysis in WWW Model

Mesh  x az  0,0,0   y a z  0,0,0  a z  0,0,0 


2 -0.000347192 0.000218351 -0.000103749
3 0.000481534 -0.00121574 -0.000101537
4 -0.0000673091 0.0000530251 -0.0000685456
5 -0.0000922419 0.000190898 -0.0000697723
6 -0.000131238 0.000109448 -0.0000485912
7 -0.000131143 0.000191035 -0.0000488936
8 -0.000182861 0.000138392 -0.0000370691
10 -0.000233351 0.000152613 -0.0000295712

(10) For the numerical solutions in Ex. 9, investigate the stability of the
topology of the solutions for different mesh sizes. Compare with the
results shown in Figs. 16.39 and 16.40. Based on these results, we

Figure 16.41. Pressure in the f1WWW model


640 Geometry, Language and Strategy—Vol. 2

have used a mesh size of 10 in our computations in Part 3 of this


volume. Is this reasonable? Include in the analysis, that the larger
the mesh, the longer the calculation takes. A possible way to avoid
using this approximation is using the seasonal gauge, Sec. 6.10, Ex.
52. What would be the advantages and disadvantages?
(11) For the forced WWW (f1WWW) model calculations, we start with
the WWW parameters suggested in Ex. 7, along with the seasonal
component. We then add an additional harmonic with frequency and
weights (Cf. Eq. (14.1)) for the standing wave solutions Sec. 16.10
and appropriate changes for the travelling wave solutions, Sec. 16.9:
  1.0
  12
U a  0,0,0    1 20 1
30
1
100
1
.
50 (16.24)
V  0,0,0   0 0
a
 0 0
a  u1 u2 u t

Figure 16.42. Acceleration potential in the f1WWW


model
Two-Person Decision Processes 641

Provide the appropriate changes for travelling waves and for


standing waves.

Figure 16.43. Streamline “tube” in f1WWW


model

(12) We have seen the seasonal variations of the pressure, Fig. 16.33 and
the acceleration potential, Fig. 16.34. When we include an
additional forced harmonic, the pressure develops corresponding
structure, Fig. 16.41 as does the acceleration potential, Fig. 16.42.
The observed effects are quite small here. What would you have to
do to make the structures significant?
(13) The free fall harmonic solutions in Sec. 15.15 Ex. 5 and in this
section, display two interesting attributes: a linear solution and a
helical solution that circulates around it. This is an inherent property
of the free fall solution.
642 Geometry, Language and Strategy—Vol. 2

Figure 16.44. Normal coordinate frames as seen


in the orthonormal frame for f1WWW model

Specifically it is a property reflective of the value   , Eq. (15.31).


A general solution however, will start at an initial time from a
known state, specified over all spatial points. It can be represented
as an infinite sum over harmonics.

Figure 16.45. f2WWW model transfer function phase gradient


The time dependence of the general solution at x  y  z  0
defines a solution that is no longer linear. Away from this “still
Two-Person Decision Processes 643

point” will be fluctuations around this behavior. We may observe


the inherent structure either from an exploration of all of the
different conditions at an initial time or from an exploration of all of
the possible harmonic behaviors along z  0 . In this exercise we
choose the latter. Consider the following curves in the f1WWW
model with two harmonics, one of which is the free fall harmonic.
The behavior along z  0 is now itself a helix, Fig. 16.43, with
additional rotational terms around it. The transverse directions are
written as polar coordinates, x   cos y   sin   . The
contours are for  (solid) and  (dashed). See if you can generate a
model with similar behaviors. Note that the model predicts the
behavior for z  0 , which means that the behavior at the initial time
is now predicted.

Figure 16.46. f2WWW model travelling wave for frame transformation


E t z   cos ,  sin  , z, 
644 Geometry, Language and Strategy—Vol. 2

(14) In Fig. 16.44, the vectors represent the frame components


E u z E u  (dashed) and E t z E t  (solid), which are the
directions of the coordinate axes vectors u and t as seen in the
orthonormal frame, relative to the axes plotted, z   . Explain
why three regions might be excluded: g  0 (gray as shown);
tt

gtt  0 ; and using Sec. 17.1, E u n  0 . Compare your arguments to


m

those in Sec. 9.11, Ex. 7.


(15) There can be more than one forced harmonic: in fact there can be a
continuum of such harmonics. In the continuum limit, we have
solved the problem of setting the initial conditions on a spacelike
boundary orthogonal to a timelike direction. In this exercise,
consider two forced harmonics for the work-wealth-wisdom
(f2WWW) model. Take stronger parameters than in Eq. (16.24):

1  12  1  e
1  1 2 2  1 2
U a 1  0,0,0   U a 2  0,0,0   1 2 1
3
1
10
1
5 . (16.25)
V a 1  0,0,0   V a 2  0,0,0   0 0 0 0
a  u1 u2 u t
Two-Person Decision Processes 645

Figure 16.47. Streamline “tube” in f2WWW


model

The frequencies are chosen so that their ratio is not a rational


number, so as not to introduce any unanticipated correlations. The
transfer function phase shift is shown in Fig. 16.45 and a frame
transformation for the model is shown in Fig. 16.46. Do these
figures demonstrate a travelling wave solution? If so, what is the
velocity of the wave? Note that the phase shift gradients are not
constant, so the reduced transfer phase shift varies with position.
What does this imply?
646 Geometry, Language and Strategy—Vol. 2

Figure 16.48. Pressure in the f2WWW model

(16) The streamline “tube” in the f2WWW model, Fig. 16.47, has more
structure than the previous f1WWW model, Fig. 16.43. Discuss the
differences. Take into account the vastly different time differences.
For two forced harmonics, the proper time is in the interval
0    10 .

Figure 16.49. Acceleration potential in the


f2WWW model
Two-Person Decision Processes 647

(17) In a similar way to the previous exercise, compare the pressure


Fig. 16.48 and acceleration potential, Fig. 16.49 in the f2WWW
model to the corresponding quantities in the f1WWW model. Why
are the oscillations more pronounced?

Figure 16.50. Normal coordinate frames as seen in


the orthonormal frame for f2WWW model

(18) In what way does the frame transformation, as seen in the normal
frame for the f2WWW model Fig. 16.50, show the effect of a
stronger coupling than the corresponding curve for the f1WWW
model, Fig. 16.44?
648 Geometry, Language and Strategy—Vol. 2

Figure 16.51. Travelling wave recurrence plot for the inequality flow in
the f2WWW model

(19) We see new structures at shorter time frames in the recurrence plots
for the f2WWW model. Explain the structures in Fig. 16.51 for the
inequality flow in this model.
Chapter 17

Steady State Geometries in 3D

In this chapter, we continue our analysis of two-strategy two-person


games. Our analysis has much in common with engineering disciplines
such as electrical engineering. In those disciplines, specific models rarely
capture the full capabilities of the theory: however, one can isolate
effects that are locally verifiable, which the theory then puts together into
a connected whole. We gain significant insight into the causal behavior
of the theory by looking in detail at local properties of steady state
solutions which can then combined by the theory.
As in electrical engineering it takes mathematical sophistication to
apply a steady state analysis to what is essentially a time dependent
behavior. This chapter also provides insight into that mathematical
reasoning. In both local behaviors and time dependent behaviors, the
boundary conditions are an important part of the resultant behavior. We
shall focus not only on the boundary conditions of a simple example, but
shall indicate what might change if these boundary conditions were
different.
We consider several geometric pictures as guides. We use geometry
in the context of differential geometry, which may be different from the
casual definition we may use in conversations. As opposed to a thing that
we see, a space is described by its metric, which are tensor fields that we
may not directly observe. Thus we consider the fields that arise and their
behaviors. In this context for example, the electromagnetic fields are
attributes of internal symmetry geometry just as the space-time metric
defines the external general relativity geometry in physics.
In three spatial dimensions, we find that the differential equations in
engineering and in decision process theory that define geometries are of

649
650 Geometry, Language and Strategy—Vol. 2

two main types: “divergence” equations or “curl” equations. The


geometry that results from the “divergence” equation is represented by
the vector field normal to the surfaces in which the scalar is a constant.
The geometry that results from the “curl” equation is represented by a
vector field that circulates around a vector potential. In higher
dimensions, the “curl” equation and its interpretation are generalized.
In decision process theory, we often have many more dimensions
than three. We have three distinct geometries corresponding to three
distinct sets of fields: the active geometry (external geometry) with
metric fields g ab , the active-internal geometry with electromagnetic like
fields F j ab that generate Coriolis like forces and the internal geometry
(inactive geometry) with scalar fields  jk that generate gradient fields.
Because we don’t explicitly take into account every dynamic variable,
we nevertheless must take into account their effects on the energy and
momentum. Thus, we also have stress tensor fields T that drive each of
these geometries based on these other dynamic degrees of freedom.
For decision process theory, the harmonic gauge determines the
active geometry for the coordinate potentials Eq. (14.1). In addition to
the linear time solutions, we have harmonic solutions that are either
natural harmonics or forced harmonics, which were explored in some
detail in Chap. 16.
Each of these solutions is a solution to the decision process field
equations, as is any linear combination of these solutions. The harmonics
are described in terms of their reduced transfer modulus and reduced
transfer phase, Eq. (16.18):

Ua  x, y , z   Ha e z cos  a z  a 


a

. (17.1)
Va  x, y, z   Ha e z sin  a z  a 
a

As in electrical engineering, the time domain dynamics are explored


through these frequency domain steady state moduli and phases; we can
treat the dynamic problem as a steady state problem. Implicitly, we may
be ignoring transient effects, depending on how we treat the initial
conditions.
Steady State Geometries in 3D 651

The steady state fields from Sec. 14.9, and the transfer functions,
depend on the coordinates that are transverse to the energy flow as seen
in the co-moving orthonormal coordinate frame. Adding the transfer
function to the steady state fields expands their role in providing the
geometric picture of the fields. The steady state view provides one guide.
This view however is not complete. As we move along a streamline
solution, the motion may slow down or stop as we approach a region that
does not make sense. It would not make sense for example to have an
event from the future impacting the present. We need a geometric picture
to help us anticipate that possibility. That is a second category that may
help guide us.
We focus in this chapter on each of three geometries indicated above
as well as on the geometry of the regions that make sense for decision
processes. For the models under discussion, these geometries are in 3D:
three dimensions. For illustration purposes in this chapter, we use the
attack-defense model defined in Sec. 16.2 with two forced harmonics
(f2AD).
Decision process theory provides an integrated view of all of the
effects. We pointed out the importance of treating causal behaviors
globally in Sec. 13.1 as a generic goal of modeling the social and
economic behaviors. Our approach here is the same as the one described
in that section. We look for stable behaviors and see how the system
responds to stimuli. The theory closes all the loops so that all the
interactions are captured. We can easily change our assumptions about
initial values, which may change the relative importance of each of the
effects.
We start with a geometric picture of the valid regions in Sec. 17.1: we
specify the regions of validity based on the causal behaviors. In Sec. 17.2
we examine the transfer functions for the f2AD model, which describes
the active geometry. Next, we build on game theory by looking at the
payoffs as fields that describe the active-internal geometry. The internal
geometry relies on the concept of payoffs that in 3D have a vector field
structure articulated in Sec. 17.3.
One ingredient of this structure is an external current, the player
passion field, Sec. 17.4. After considering the geometry of this structure,
we address the coupling of these fields to actions through the player
652 Geometry, Language and Strategy—Vol. 2

engagement, Sec. 17.5. This opens up the discussion about the gradient
geometry.
The player engagement is generated by the player interest, Sec. 17.6,
which in turn generates the player strategy bias, Sec. 17.7, which
generalizes the concept of game value in game theory. Cooperation
between players is set by the player bonding and player shear, Sec. 17.8,
which provide a description of the internal geometry.
We return to the composite payoff, Sec. 17.9, which we take initially
to be zero on the initial surface. Finally we deal with the seasonal
payoff, Sec. 17.10, which can be viewed as an attribute of the active
geometry or of the inactive geometry. We show that the stationary
geometry helps us understand the significant dynamic features of the
theory and simplifies the computations. Note that there is significant
numerical work required to go from the steady state view to the dynamic
time dependent view, just as there is in electrical engineering.

17.1 Validity Regions

Before restricting attention to the frequency domain, we consider a time


domain aspect of the solutions that we obtain for the harmonic gauge
coordinates. The coordinate potentials specify a general linear
transformation from the co-moving orthonormal coordinate frame to the
normal coordinate frame. The transformation is continuous and well-
defined. However realistic solutions may be confined to specific regions.
This is because not all values of the field solutions are meaningful.
There are regions of validity and non-validity for any given solution. The
analogy from physics is that not all speeds are possible. You can’t go
faster than the speed of light nor can the speed of light be zero. Such
constraints are natural in a complete theory and come about in a natural
way. For example, you can’t go faster than the speed of light. If you try
to go faster, it takes more energy. The energy required increases, the
closer to the speed of light you get. To achieve the speed of light would
in fact take all the energy in the universe. Thus this region would be
excluded.
Steady State Geometries in 3D 653

There are also constraints in any description based on the coordinate


system used. For example you can project the sphere onto a plane, yet
the projection will have one point, the North Pole, that lies at infinity.
This is an artifact of the coordinate system, not an intrinsic problem of
the sphere. What it suggests is that you have to identify when you are
getting close to singular points and determine whether they are artifacts
of the coordinate system or the description. Curved spaces in general
require multiple local (flat) maps since no one map can cover the whole
space.
In this section we consider some of the restrictions that need to be
considered, starting with those restrictions based on the active coordinate
 
frame transformations E a E a E a . For the fixed frame model, we
have the requirement that E a   e E a  , Eq. (4.52), so we have only two
classes of transformations: those that transform the active coordinates to
the flow vector and those that transform them to the transverse
directions. For our models here, that means frame transformations to one
of the directions x y z in the co-moving orthonormal frame.
Our approach to understanding the validity regions is to consider
certain constraints that are implicit in our definition of the normal
coordinate frame.
 The causality assumption is that time t in the normal coordinate
frame should always increase:

t
 E t  0 . (17.2)


This assumption depends on more than just the steady state geometry;
it must be satisfied along every streamline: as the proper time
increases along the streamline, normal time also increases.
 For n a active strategies, the volume assumption is that the invariant
volume element  g du1    du na  dt that is used to define the
action, Eq. (3.29) is non-singular, so the determinant 0  g 1   .
The consequence is that the determinant of the frame transformation
must be non-singular:
654 Geometry, Language and Strategy—Vol. 2

m m 2
E u n Eu 
  1  e e 
na 
g ab
0. (17.3)
E t n E t

The determinant can be expressed in terms of a reduced frame


transformation:

g ab  1  e e  E t  E u n
2 2
 
na m
0
1  e e  0
E u n  0
m
. (17.4)

um
E E t n
E u n  E u  n 
m m 

E t

In addition to the causal assumption, we therefore have 1  e e  0


and E u n  0 . As an example, in the model calculation, Fig. 16.50,
m

we indicated the valid region by no shading, corresponding to


1  e e  0 .
 In any frame, the time component of the energy momentum tensor Ttt
is the energy and should be positive. It can be written in terms of the
energy density  by use of the frame transformations: Ttt   Et Et .
Since we start with initial conditions in which Et  0 , and since it is
not reasonable that the energy density changes sign, this condition
holds everywhere. This is the positive energy density assumption.

We can put a tighter bound on some of the frame transformations


such as Et . Through each point along a streamline, consider the null
geodesics, defined as the paths with zero invariant length, which satisfy:

du a du b
g ab  0,
dt dt
du m du n du m
m, n  t  g mn  2 gtm  g tt  0 . (17.5)
dt dt dt

The eigenvalues of the metric g ab must be Lorentz: there must be one


positive eigenvalue and the remainder must be negative in a valid region.
Steady State Geometries in 3D 655

In general, Eq. (17.5) describes a point (if all eigenvalues are the same
and the metric is not Lorentz), an ellipsoid shape or a hyperboloid shape
at each point along the streamline.
We set the value g tt  0 at the initial point along the streamline as
well as the remaining eigenvalues, which are set to be negative. The
shape of the ellipsoid or hyperboloid is determined by the eigenvalues of

Figure 17.1. Communication ellipsoid for f2AD model


the spatial components of the metric and the off-diagonal time term. The
initial shape is an ellipsoid.
Along a streamline, the eigenvalues g mm are assumed to stay non-
zero, which is true as long as gmn  gg tt  0 . We assume E t  0 and
E u n  0 so we conclude that g tt  0 . Based on these assumptions, the
m

equation for the null geodesic in terms of the eigenvalues is:

 du m du m du m 
 g
 mm
m  dt dt
 2 g tm
dt 
  gtt  0,
2 2
 du m gtm   g tm 
  g mm     g tt   g mm   . (17.6)
m  dt g mm  m  g mm 
656 Geometry, Language and Strategy—Vol. 2

The right-hand side is positive initially, since the spatial eigenvalues are
all negative. When an eigenvalue goes negative there is either no solution
or the solutions are an ellipsoid or a hyperboloid.
As examples of the ellipsoid along the streamline, we show two cases
for the f2AD model: one with g tt  0, g tt  0 and one with
g tt  0, g tt  0 . In both cases E t  0, Et  0 .
We see that when g tt  0, g tt  0 , the shape (Fig. 17.1) is a
communication ellipsoid, which implies that at each point, there is an
open set around the streamline and there is a null flow in any direction in
that open set.

Figure 17.2. Communication hyperboloid for f2AD model


If along the streamline one of the eigenvalues g mm goes to zero, then
g tt goes to zero and from Eq. (17.6), the shape is no longer an ellipsoid.
The shape becomes a communication hyperboloid, which in general is
unbounded (Fig. 17.2), suggesting that for some direction, “light” can
travel with infinite speed.
This suggests we impose the condition g tt  0 . With this condition,
when g tt  0 , the communication ellipsoid (or communication
hyperboloid) passes through the point in which all eigen-velocities are
zero. We therefore consider that g tt  0 is also a singular point, since
Steady State Geometries in 3D 657

this means that for an open set around that point on the streamline, there
will be parts of space that will never be reached. We impose the
communication assumption that you can communicate to any part of
space: g tt  0 . This ensures that the point where all eigen-velocities
vanish occurs strictly inside the ellipsoid.

Figure 17.3. Energy-momentum ellipsoid for f2AD model

We make a similar argument using the covariant flows. Consider the


invariant energy momentum flow for “light” g ab a b  0 for the energy
momentum  a at each point along the streamline. The argument
proceeds exactly as above and so we conclude that there is also a
covariant communication ellipsoid, such as Fig. 17.3, with the condition
g tt  0 .
658 Geometry, Language and Strategy—Vol. 2

If we further allow the possibility that g tt  0 at some point along the


streamline, then again it appears that light gets trapped. This is a
contradiction so we have the additional condition that g tt  0 and the

Figure 17.4. Offset energy-momentum ellipsoid for f2AD model

origin is inside the ellipsoid. If this is not the case, the origin is no longer
inside the ellipsoid as in Fig. 17.4. This argument extends the
communication assumption to require both conditions: g tt  0, g tt  0 .
These constraints in turn impose constraints on the behavior of the
flows and streamlines, Exs. 4-5. We see that the validity regions are
specified by the βn and its dual * βn, Ex. 8, along with E t  and the
reduced frame transformation E u n .
m

17.2 Transfer Functions

The dynamic content of the time domain is captured by the transfer


functions in the frequency domain. Since the harmonic functions in time
can be linearly superposed to reproduce any given set of initial
conditions, the transfer functions determine the time dependence of the
coordinate scalars in the harmonic gauge.
Steady State Geometries in 3D 659

Figure 17.5. Transfer phase gradient for f2AD model

We have provided examples of the (gradient of the) phase for the


f2WWW model in the previous chapter, Fig. 16.45. For the f2AD model,
the corresponding result is Fig. 17.5. The time dependence of the
coordinate frame transformations determines the active metric content.
See for example, Eq. (4.72). In this context, player engagements e are
part of the frame transformation.
In the holonomic central frame, the metric contributions
g g g  are independent of proper time, Sec. 6.10, Ex. 23.
These components depend on the elements of the determinant 1  e e ,
Eq. (17.4), the acceleration potential e q and the characteristic vector
potential a , which determines the seasonal harmonic. These fields
provide the stationary component of any solution. In addition, we have
the explicit time dependence as determined by the transfer functions, Eq.
(16.18).
Figures such as Figs. 16.45 and 17.5 provide important information
about the harmonics: they demonstrate whether the harmonics are
initially standing waves or travelling waves. For these two models we
have travelling waves for the forced harmonics and a standing wave for
the seasonal harmonic. The gradient of the phase gives the propagation
constant  a for each coordinate direction.
The ratio of the frequency to the propagation constant,   a is the
velocity of the wave. We interpret the harmonic   2 f as
660 Geometry, Language and Strategy—Vol. 2

proportional to the frequency f of the wave; we interpret 2  a as


the wavelength. We have picked out a single direction for the gradient of
the phase: we can in general pick any one of the three directions so that
the propagation constant is easily generalized to a vector defining a wave
in three dimensions.
In addition to the transfer phase denoting a travelling wave or a
standing wave, it also indicates that as we move along a particular
direction, the gradient of the phase changes, Fig. 17.5. Such changes are
common in electrical circuits as we move along a circuit and indicate the
presence of inductance or capacitance. Such components store energy
and release energy so that a wave can pick up a phase as we move over a
circuit element.
Here we have a more complicated situation, since the wave is
travelling. The phase shift effect is clearly spelled out in Eq. (17.1). We
see that as we increase z in Fig. 17.5, the gradient of the phase is not
constant, but increases: the phase increases as we increase or decrease
away from the initial point z  0 . This is an indication that there are
mechanisms in the theory for capturing or returning energy, just as there
are in electrical circuits.
The phase gradient for each harmonic component clearly specifies
some of the initial conditions for the forced and seasonal contributions.
The remaining initial conditions are specified by the transfer modulus for
each harmonic. For the f2AD model, we have Fig. 17.6, where we plot
the modulus in units of decibels (dB), 10Log10 H . We see the model has
been constructed so that the forced harmonics are much smaller than the
seasonal harmonic.
In the simplest theories of wave phenomena, such as geometric
optics, one initially ignores the effects of frequency. Such effects are
however important, since one observes aberrations or lack of focus of
light due to variations in frequency. Similarly, we see in decision process
theory that the shape of the transfer modulus does in fact depend on
frequency. This means that when we have a superposition of multiple
frequencies, which occurs when the initial wave form is not a pure
harmonic, we will see the analog of aberrations and dispersion.
Steady State Geometries in 3D 661

Figure 17.6. Transfer modulus for f2AD model

It is worth emphasizing that the harmonics for each coordinate


direction can be chosen independently. This means that each can have
travelling or standing waves, can have their own phase shifts and their
own wave forms and subsequent dispersion effects. This increases the
complexity dramatically of what might be observed and defines the
geometry of the active space. The active geometry is characterized by the
transfer moduli and transfer phases.

17.3 Player Payoffs

The dynamics depend not only on the active geometry but on the active-
internal geometry, which is described by the player payoff fields. We
describe them in the co-moving orthonormal frame. For the two-person
games under consideration, these fields are defined in terms of the three-
dimensional field equations in Sec. 14.9. Complete solutions are obtained
from these equations by providing appropriate boundary conditions,
which we do by providing the initial values for the fields and their
gradients along z on the two-dimensional surface z  0 .
662 Geometry, Language and Strategy—Vol. 2

The geometry is defined by the boundary specifications and by the


field equations. We focus on these behaviors to learn the underlying
geometry. We start with the payoff fields.

Figure 17.7. Payoff field for “blue” player


The player payoffs provide the “vector” part of the active-internal
geometry. Each payoff can be written in terms of the “curl” of a vector
field, as is done explicitly in Sec. 14.16, Ex. 2. This is part of a more
general technique that we illustrated for decision process theory in Sec.
2.11 of breaking the full geometry into active components, inactive
components and their mixture. In that context, we are focusing on the
“mixture”. We deal separately with the pure inactive components later.
The geometry of the payoff field is similar to that of electromagnetic
field lines; the field equations are similar. We gain some insight from
this comparison. The payoff fields are also the core of game theory and
form the basis of our game equivalency, Sec. 15.1. However, we specify
the payoffs for each player not only at a point, but on the full boundary
surface. We use as an example the f2AD model, where player 1 is defense
Steady State Geometries in 3D 663

or “blue” team and player 2 is “red” or attack team. The model helps
identify the similarities with game theory and electromagnetic theory as
well as the differences.

Figure 17.8. Playoff field for “red” player

For example, the payoff field for “blue” is primarily along the z
direction, Fig. 17.7; similarly for “red”, Fig. 17.8. This direction is
orthogonal to the attack-defense strategies, so it is not a surprise based on
the game theory input. The behavior is present in the initial conditions at
z  0 : the largest initial vector components are shown in Figs. 17.9 and
17.10.
The player payoff can also be derived from a vector potential,
Bα  B    Aβ , Sec. 14.16, Ex. 2. Though not computed, we infer
from the electromagnetic field analogy that the vector potential fields
that generate these payoff fields will consist of circular orbits in the x, y
plane.
664 Geometry, Language and Strategy—Vol. 2

Though not a requirement of the


theory, we picked the initial
boundary conditions to match, as
closely as possible, to the ideal of the
Nash equilibrium by focusing the
initial payoffs to be maximal at
x  y  0 . In this way we hoped to
explore how the theory would
differentiate itself from game theory.
We found solutions for the partial
Figure 17.9. Initial boundary condition
for “blue” player differential equations with periodic
boundary conditions on x and y .
We chose a boundary size of 5 .
The effect of these conditions is
seen in the resultant payoffs for “blue”
Fig. 17.7 and “red” Fig. 17.8, where
we plot over the extended regions
10 to show the periodic behaviors.
We note that specification of these
boundary conditions is totally under
our control. In general, they need not
be periodic or have a simple
maximum value. Figure 17.10. Initial boundary for “red”
It is a property of partial player
differential equations that the solutions will change continuously as you
move away from a boundary, so the shape of the boundary will manifest
itself as a shape in the solution. We see this in Figs. 17.9-17.10.
The field equations, Sec. 14.9.6, provide the second aspect of the
geometry. These equations are generalized electromagnetic field
equations: there are both “curl” equations involving   B α and gradient
equations involving   B α . The comparison to electromagnetic fields
provides insight on the behavior of the field lines that we see in our
numerical figures. In particular, we are guaranteed solutions when both
sets of equations are provided.
Steady State Geometries in 3D 665

Figure 17.11. “Blue” player transverse payoff

The gradient equations prevent payoff field lines from being created
or destroyed. We expect field lines that are approximately straight, or
closed loops. For “blue” and “red” payoffs we see such lines.
We might see quite different behaviors if we change the boundary
conditions on z  0 , Figs. 17.9 and 17.10 to oscillate for example or if
we impose conditions on the outer boundaries as opposed to z  0 . Since
the equations are elliptic in nature, there exist solutions; they are likely to
be similar to those found in electromagnetic theory, where there is a rich
set of solutions based on different possible boundary conditions.
The “curl” field equation   B α , as the name suggests, determines
the rotation of the field lines. For our solutions, the rotations are small
compared to the strength of the field along the z axis.
We observe a small effect however, which we isolate by looking only
at the transverse components. For “blue” and “red” players we see the
effect by looking at representative slices. Now we see evidence that
the field lines can form closed loops. For “blue” player Fig. 17.11, we
see that there is a circle centered near x  5, y  0 . For “red” player
Fig. 17.12, there is a circle centered near x  0, y  2 .
666 Geometry, Language and Strategy—Vol. 2

Figure 17.12. “Red” player transverse payoff

It is worth noting that the relative size of the effects has everything to
do with the initial boundary conditions. What is important is that the
“curl” and “divergence” mechanisms are always in play and depending
on the initial conditions, may totally change the sizes of these effects. For
example, the effects that we see are quite different for the code of
conduct (player 3), Fig. 17.13. Here we see circulations around
x  0, y  0 , even on the initial boundary z  0 .
It also illustrates an attribute of equations in differential geometry that
some equations are constraint equations. Indeed the curl component
 x B y   y B x of Eq. (14.24) is just such a constraint, which must be
satisfied on the initial boundary. Although by construction, we start with
an initial boundary condition close to zero, Fig. 17.14 for B3 z , we can’t
consistently set the other components to zero.
Steady State Geometries in 3D 667

Figure 17.13. Code of conduct transverse payoffs

The total payoff field has lines that now flow in both directions,
Fig. 17.15. The field equations are satisfied: that lines disappear or
appear is due to the divergence terms in Eq. (14.24).
Thus in computing model payoffs, we have only scratched the surface
of what is possible based on the boundary conditions and the field
equations. If we move away from
periodic boundary conditions and
from the focus on Nash equilibrium
like solutions, then we will see a
much richer structure, such as what
is seen for electromagnetic static
fields.
The analogy is even closer since
it is possible to write the decision
Figure 17.14. Code of conduct initial
process theory equations for the
boundary condition
payoff in terms of a vector potential,
Sec. 6.10, Exs. 43-45 and Sec. 14.16,
668 Geometry, Language and Strategy—Vol. 2

Ex. 2: we can use the same intuition for the geometry of the
electromagnetic field lines in this theory. This approach generalizes to
higher dimensions.

Figure 17.15. Code of conduct payoff

17.4 Player Passion

The geometry for the player payoff field lines is not complete without a
discussion of external effects that generate currents and hence provide
sources for the payoffs.
The mixed stress components of the energy momentum tensor p in
the co-moving orthonormal frame define a source, the player passions
Pα , for the player payoffs. We identify such stresses as arising from
dynamic contributions that we have not explicitly included with the
active players and their strategies. From the field equation for the player
payoffs, Eq. (14.24), we see that the player passions do indeed provide a
source current for the player payoffs.
Steady State Geometries in 3D 669

Figure 17.16. “Blue” player passion density plot for f2AD model

For the examples in this chapter, we have assumed the conductivity


model, Sec. 6.3 and assumed that each player takes on a certain passion
Pα  P  P set by the scalar potential P  x, y, z  , Sec. 14.16, Ex. 1.
We assume that the initial passion components are:
2  x  2  y 
P1  x, y ,0   cos2    cos   
5  4 10   4 10 
2   x  2   y 
P2  x, y ,0    cos2    cos   
5  4 10   4 10 
2 x 2y
P3  x, y ,0   cos2 cos . (17.7)
5 10 10
P   Lx , y , z   P  Lx , y , z 
P  x,  Ly , z   P  x, Ly , z 

The form of the player passion potentials on the initial boundary has
been chosen arbitrarily. The model is periodic on the transverse
boundary, whose behaviors are given in Fig. 17.16 for “blue” and Fig.
670 Geometry, Language and Strategy—Vol. 2

17.17 for “red”. A similar density plot can be obtained for the code of
conduct player corresponding to Eq. (17.7). See Ex. 11 and Fig. 17.39.

Figure 17.17. “Red” player passion density plot for f2AD model

The geometry of the gradient passion field is determined by our initial


conditions, which for model consisted of a specification of values on the
boundary z  0 and periodic boundary conditions for the other two
dimensions. Different boundary conditions will yield different
geometries. The geometry determined by Eq. (17.7) shows major
gradient changes in the upper right quadrant for “blue” and in the lower
left quadrant for “red”.
The code of conduct payoff is also quite different from the “Blue” and
“Red” behaviors. We have set the code of conduct on the boundary to be
zero. That means the behavior is impacted more by the “current”
generated by the player passion.
Each of the player passion potentials generates a player passion
vector field. See Ex. 12. The fields are computed to be mostly transverse
to the z direction. Some of this can be anticipated from the shape of the
density plot. The contours drop off sharply away from the haystack
Steady State Geometries in 3D 671

structures in each plot. We expect vector fields to be normal to the


surfaces of constant potential.
Of course the situation here is complicated by the presence of the
periodic boundary conditions. That may skew our expectations, though
in this case the behavior seen is actually as expected. What is important
is to realize the effects boundary conditions can play on the geometry and
make sure those effects have been properly accounted for. In our case,
the proper accounting is done by virtue of having a complete solution of
the field equations.
As part of the complete solution, the player passion will depend on
the player engagement of the code of conduct player (Sec. 17.5). It will
also depend on the composite payoff vector, which is primarily along the
z direction (Sec. 17.9). This is the only direction that the two players
have in common. The vector field is not uniform however, indicating the
important effect of the various other contributions. The field equations
for the seasonal potential (Sec. 15.15, Ex. 7), will generate an influence
for example (Sec. 17.10). We turn now to these additional effects.

17.5 Player Engagement

The impact of the player payoffs on decision making is based on the


strength of the player engagement. This is because the field equations
demonstrate that the composite payoff determines the forces, Eqs. (2.49)
and (15.31); it is the weighted average of the player payoffs with the
player engagement factors that determines the seasonal payoffs in the
game equivalency approximation. Player engagement arises from the
flow and is a property of the stress tensor.
If the strength of the player engagement goes to zero, the player acts
without constraints by other players: the player is entitled. If the strength
is large, the player is either strongly attracted or repelled by the
behaviors of others. The geometry of the player engagement is therefore
of great significance. As a scalar field, its effects are made through the
size or density of the field as well as its sign. We expect forces to be
generated along the normal of the surfaces of constant engagement, so
the possible density plots and surfaces describe the geometry.
672 Geometry, Language and Strategy—Vol. 2

We start with the attack-defense (f2AD) models solutions. What is


interesting about these solutions is that they arise from an assumption
that there is some engagement initially and that furthermore, the initial
conditions are periodic on the x y boundary which determines the
preferences of each player to defend or attack, respectively. The simplest
boundary conditions are those in which there are replicated positive and
negative towers of engagement along the z direction, such as for “blue”,
Fig. 17.18.

Figure 17.18. “Blue” engagement e1   e1 density plot

The density plot for “blue” generates a vector field that will be
vertical, along the positive z direction for z  0 and opposite direction
for z  0 . There will be no substantial transverse flows to this direction.
This qualitative behavior is a consequence of the boundary conditions
chosen. We can imagine different boundary conditions. Indeed, we see
different behaviors for “red”, Fig. 17.19. We find the gradients for the
vector field for this case to be similar.
As a third example, consider the player engagement for the code of
conduct player, Fig. 17.20. This shape is quite different, though here the
field lines are vertical, along the z direction. We get vertical vector
Steady State Geometries in 3D 673

fields because of our starting boundary assumptions. We can choose to


add higher harmonics in the z  0 plane, or choose different boundary
conditions on the surface of the strategy box. We imagine cases in which
there is a uniform density ball of engagement at the center, in which case
the fields would all flow out from that like an electric charge. Or, we
could have a positive and a negative charged ball in the center, with
boundary conditions on the outer boundary set to zero. This will generate
transverse fields between the two charged balls. The possible geometries
are thus immense.

Figure 17.19. “Red” engagement density plot


The boundary conditions are set by considerations of the decision
process. In particular, the decision process provides candidates about
what is known and so sets the boundary on which we compute our
known starting point. As we have mentioned previously, weather is a
pretty good analogy. We know the weather at some initial point in time
from which we then extrapolate behaviors at later points based on the
field equations of fluid mechanics. The most interesting starting values
depend on the phenomena we wish to study. Hurricanes for example
depend on the storms that get created near the equator due to dramatic
674 Geometry, Language and Strategy—Vol. 2

heating conditions. For decision processes, we may also identify areas


that are the main sources of instability and use them as the initial values.

Figure 17.20. Code of conduct engagement density plot

For player engagement, the most interesting case might be one of two
extremes: region in which the player engagement is zero, in which case
the player payoffs are entitled and not constrained. The other case would
be when the player payoff is extremely large. In this case the players are
highly competitive or highly cooperative: they are definitely engaged.

17.6 Player Interest

Closely related to the player engagement scalar fields are the player
interest vector fields, Eq. (14.29):
1
Iα  e  qe  Θe  σα βe . (17.8)
ni

If we ignore the other fields temporarily, we see that the gradient of the
player engagement scalar field determines the player interest vector
Steady State Geometries in 3D 675

field. There are other fields and we indeed have equations of motion for
the player interest field in Sec. 14.9.5. These fields satisfy their own
boundary conditions.
From these fields we can in fact determine the player engagement
fields: the fields are thus strongly related. Conversely, we can use the
equations to derive the scalar engagement fields from which we compute
the vector player interest fields.

Figure 17.21. “Blue” player interest transverse vector field

In particular, we note that the field values in the initial z  0 plane


constrain the x y components of the player interest to be related to
the corresponding x y components of the scalar gradients of the
player engagement and the vector components of the acceleration, player
bonding and player shear components.
One of many ways to satisfy these constraints is to use Eq. (17.8) to
 
set the values of I x I y . We can supplement this with the
assumption x  y  z  0 of specific values for some of the variables at
the initial point, which we in fact do for the f2AD model:
676 Geometry, Language and Strategy—Vol. 2

q  0
I x3  I y 3  0 . (17.9)
z11  310 z 22  4 10 z 33  510 z12  0
Based on our analysis of the player engagement, we expect that the
largest components of the player interest will be along the vertical or
longitudinal z direction as opposed to the transverse x y
directions.
The transverse vector field for “blue”, Fig. 17.21, reflects the
boundary choices as can be seen in the z  0 slice. The behavior
propagates to the other slices without much change. We note that the
longitudinal contributions are much larger, as seen in Ex. 13.

Figure 17.22. “Red” player interest transverse vector field


Similarly, we see that the transverse vector field for “red”, Fig. 17.22,
reflects its boundary conditions. It is also true that the longitudinal
components are much larger, Ex. 13. In addition to inferring the
behaviors from the gradients of the player engagement, we can also look
directly at the differential equations for the player interest, Sec. 14.9.5.
We see that the source term is in the divergence of the player interest,
Steady State Geometries in 3D 677

which depends on the overlap of the composite payoff and the player
payoff, ω  B α . Since we start with game equivalency, this starts at zero.
However, this may not be the most interesting boundary on which to set
the initial conditions. Based on analogy with electric fields, a region of
high density overlap might generate quite different player interest vector
fields and player engagements.

17.7 Player Strategy Bias

Figure 17.23. “Blue” reduced player strategy bias field

The player interest is closely associated with the player strategy bias,
f j , Eq. (4.74), the analog of the electric field. See also the discussion
in Sec. 15.4. We can rewrite the player strategy bias in terms of a
reduced player strategy bias, fˆ  
, which extracts some of the frame
dependent behavior that might become singular:
678 Geometry, Language and Strategy—Vol. 2

E j  fˆ 

f j
 
1  e e . (17.10)
fˆ 

 2q e   2 I e e   2 I   2e  

The singularity occurs when the player engagement terms become


sufficiently large. We have seen that this corresponds to limits to choice,
Sec. 10.9 for Robinson Crusoe economics and generalizes here. We see
that the player strategy bias and player interest are closely related. The
forces will in fact be expressed more directly in terms of the player
strategy bias, which makes it a useful vector field to investigate.

Figure 17.24. “Red” reduced player strategy bias field

For “blue”, the reduced player strategy bias field, Fig. 17.23, is
primarily vertical. This is a force primarily along the  z direction, which
pushes towards “blue” based on the definitions, Eq. (15.24). The origin
of the direction of force is that for the boundary conditions under
consideration, the engagement terms are small and the player interest
term dominates. We can certainly envision situations where this might
not be the case.
Steady State Geometries in 3D 679

We see a similar behavior for “red”, Fig. 17.24, though in this case
the field goes in the other direction. We suggest that the strategy bias
fields are indications of the same concept as value in game theory. For
this reason we suggest the name “bias” as a force that pushes a player in
one direction or another. It suggests the same concept as game value in
game theory.

17.8 Decision Pressure, Player Bonding and Shear

The player strategy bias field depends not only on the player interest, but
on the acceleration, player bonding and player shear. Bonding and shear
define geometric properties of how a unit square changes along a path.
The player bonding and shear are consequences of the stresses in the
system, which are represented by the decision process pressure field,
Fig. 17.25.
We believe it is noteworthy that the decision process pressure density
has distinctive high pressure areas and low pressure areas. We find that
pressure plays a similar role to that played in weather: fluid flows from
high pressure areas to low pressure areas. We see that there will be
gradient forces here that do precisely that: the forces will move the
decision process from high pressure areas to low pressure areas. The
forces in the theory generate acceleration. The geometry that results will
be the geometry of the decision process flow.
The decision process pressure is proportional, by assumption, to the
energy density of the decision process. The high pressure areas indicate
areas where the process has accumulated, whereas low pressure areas
will be less populated. We indicate the steady state behavior after all
transient effects have died out; even though this is steady state, we
nevertheless allow for standing and travelling waves. It is steady state in
the same sense that AC circuits are steady state, yet describe time
dependent behavior.
680 Geometry, Language and Strategy—Vol. 2

Figure 17.25. Decision pressure for the f2AD model

Figure 17.26. Decision process acceleration potential


Steady State Geometries in 3D 681

The decision process pressure contributes to the acceleration vector


field q  q , which reflects some of its same characteristics. The
acceleration vector field can be expressed in terms of the gradient of an
acceleration potential field, q   q , Fig. 17.26. We see areas of high
acceleration and low acceleration.
These field behaviors are associated with the constraints not included
in the explicit set of players and payoffs. This is quite similar again to
weather where we include the pressure of the air, without including a
detailed analysis of the atoms that make up the air.
Stresses in general generate strains. If we consider the pressure as a
stress, what are the possible strains? We suggest that the strains are
described by the symmetric tensor  for each proper strategy  . We
decompose this tensor field into a player bonding scalar Θ   and a
traceless player shear matrix σαβ   , Eq. (14.18). These components
determine payoffs that we can identify with cooperation and opportunity.
To understand the behaviors, we need to set the initial conditions. We
use the known behavior payoffs to set some of these values in Eq.
(15.29). However, we have an insufficient set of data to complete the
determination.
We supplement our knowledge with the field equations from the
theory. In addition to setting the initial values of  , we must insure
that the values are derived from a potential and that for each active
proper strategy  , the matrices in the inactive index commute. Our
approach is to separate out the initial values we know. We then iterate
our process as follows.
The components of  are known with the exception of
 
x11 x12 x 22 y11 y 22 y12 . Given these values, all of the
components of  are specified and can be written in terms of a shear
potential scalar     . Transform this shear potential scalar to a
frame in which it is diagonal,    x , y , z  . Take for the diagonal shear
potential scalar    x , y ,0  on the boundary, a function that depends on
six variables, assuming two variables for each diagonal component with
each variable representing a degree of freedom along either x or y .
Similarly, take a form for  z  x, y,0  . It should be determined.
Compute  x and  y . This gives new values for the above unknowns,
possibly changing the value of  z . We iterate this process until closure;
682 Geometry, Language and Strategy—Vol. 2

we find it converges, giving us one of many possible solutions that fit the
known conditions and satisfies all of the field equations.
We obtained closure with these known behavior values for the shear
components:
 y   x   x   y 
 11  x, y ,0   0.0469857cos   sin    0.064015cos   sin  
 5   5   5   5 
 y   x   x   y 
 12  x, y ,0   0.20336cos   sin    0.343935cos   sin  
 5   5   5   5 
 y   x   x   y 
 13  x, y ,0   0.0563027cos   sin    0.151236cos   sin  
 5   5   5   5 
. (17.11)
 y   x   x   y 
 22  x, y ,0   0.0782177cos   sin    0.132287cos   sin  
 5   5   5   5 
 y   x   x   y 
 23  x, y ,0   0.000179326cos   sin    0.0562865cos   sin  
 5   5   5   5 
 y   x   x   y 
 33  x, y ,0   0.125203cos   sin    0.196302cos   sin  
 5   5   5   5 

These shear components are zero at the still point.


As part of this process we obtain the shear gradients along the z axis:

 0.3 0 0.362542 
 x  2 y 
 z  x, y ,0    0 0.4

0.362345  cos2   cos  
 0.362542 0.362345   10   10 
 0.5 
 x   y 
  x, y ,0   0.37561sin    0.588905sin   . (17.12)
 5   5 
6  x  2 y 
 z  x, y,0    cos 2   cos  
5  10   10 

These shear components then generate the complex interplay of effects


that we see in Figs. 17.27-17.32. What does shear mean?
An important aspect of geometry is how a unit cube changes shape as
it moves along a path, Eq. (8.3) and see Sec. 8.4. The volume changes
are determined by the player bonding scalar  , Fig. 17.27. There are
two regions where the density is high, which appear like two stacks. This
is a consequence of our boundary conditions. One consequence is a
bonding vector field which is primarily longitudinal. We can change the
assumed form of the boundary conditions to have higher order harmonics
in the transverse directions, in which case we expect to get more
complex bonding structures.
Steady State Geometries in 3D 683

Figure 17.27. Player bonding scalar field

We also note that the boundary conditions in part reflect the


simplifying assumption of the player fixed frame model. We can relax
that assumption, in which case we also relax the fairly restrictive
conditions imposed on the form of the player shear and bonding. We
think it feasible to solve such models, which we intend to do at a later
date.
In Fig. 17.28, we see the diagonal component for “Blue”. We
compare this with Fig. 17.29, the diagonal component for “Red”. They
are similar though the longitudinal flow is generally opposite.
684 Geometry, Language and Strategy—Vol. 2

Figure 17.28. Player shear vector field  11

Figure 17.29. Player shear vector field  22


Steady State Geometries in 3D 685

Figure 17.30. Player shear vector field  12

Figure 17.31. Player shear vector field  13


686 Geometry, Language and Strategy—Vol. 2

The shear components for off-diagonal components are quite


different. Between the two players, Fig. 17.30, the vector field is
reasonably well spread out, whereas between “Blue” and the code of
conduct player, Fig. 17.31, the field is localized near the origin. We see
something similar between the code of conduct player and “Red”, Fig.
17.32.

Figure 17.32. Player shear vector field  23


The player shear gives a description of how the geometry changes as
we move along an active strategy direction  . That we have a shear
indicates that the model we are discussing is responding to a variety of
stresses in order to maintain the player fixed frame model. More
generally, we expect strains whenever we impose boundary conditions.
One unexpected benefit of computing the player shear tensors is the
possibility of observing not only longitudinal vector fields, but transverse
vector fields as well. We observe this for the diagonal components of
shear for both “blue” and “red”, Figs. 17.28-17.29. We trace this
behavior back to the form on the boundary, Eq. (17.12), where we can
see that the longitudinal component vanishes whenever  x 10   2 or
 y 10   2 . Thus along the boundaries, the transverse components
Steady State Geometries in 3D 687

have the opportunity to compete. This again shows the importance of the
boundary conditions on the qualitative shape of the geometry of the
scalar and vector fields.

17.9 Composite Payoff

We used game equivalency to set the composite payoff to zero on the


initial boundary. This can’t be maintained however because of the field
equations. The composite payoff represents the rotational aspect of the
geometry as seen in the orthonormal co-moving coordinate basis. If in
this basis the frame does not appear to rotate, then the frames are fixed.

Figure 17.33. Composite payoff


However we saw that this does not prevent a seasonal payoff from
being present, which will be the weighted average of the player payoffs
with each of the player engagement couplings. The general case is more
complicated. We compute the composite payoff for the f2AD model,
Fig. 17.33. We note that the overall value of the vector field is in fact
small compared to the player payoff fields, such as the one for “blue”,
Fig. 17.7: the ratio of scales is 0.10: 0.0125, or about 8:1.
688 Geometry, Language and Strategy—Vol. 2

According to Sec. 14.16, Ex. 2, we can find an integrating factor and


write the composite payoff in terms of the composite scalar field,
ω    , at least in three dimension. It suggests that the composite
field is in some ways simpler than the other payoff fields. The geometry
is determined by the composite scalar field. The source of this gradient is
I α  B α . Cf. Exs. 15-16.

17.10 Seasonal Payoff

The composite payoff is determined not just by the player engagements


and player payoffs, but by the seasonal payoff, Eq. (14.30). The seasonal
payoff is a tensor field determined by the seasonal vector potential, Eq.
(14.31), the gauge condition Eq. (6.64) and an additional condition on
the boundary, Sec. 15.15, Ex. 7 and Eq. (15.36). The seasonal payoff
values that we obtain, Fig. 17.34 turn out to be quite different from the
player payoffs and comparable scales: the ratio of the maximum scale of
one to the other is 0.10:0.05.
The player payoffs exist because of an isometry associated with each
player. The isometry is an internal strategy not visible to any other
player, which generates a symmetry transformation in the theory,
associated conservation laws and a player vector field 1-form whose 2-
form is the player payoff. There is in fact a similar mechanism behind the
seasonal payoff.
In the fixed frame model, we have assumed a time isometry: there is a
symmetry transformation associated with time. In the central co-moving
frame, this isometry is manifest in that the metric elements are
independent of time. There is a vector field 1-form generated from the
characteristic potential, which in turn generates the seasonal payoff 2-
form.
Steady State Geometries in 3D 689

Figure 17.34. Seasonal payoff vector field

As a side note, the equations that generate the seasonal payoff have a
great many terms so it is helpful to identify ways that insure the
calculations are correct; using the fact that the gauge condition should
remain satisfied at all values of z is one helpful check. Our
computations meet this minimal test.

17.11 Outcomes

We have explored the geometry of the theory using the f2AD model as a
guide, along with multiple appeals to properties of partial differential
equations. The geometry of the theory consists not just of the space in
which actions occur, it consists of the fields. There are both scalar and
tensor fields, each with their own special set of equations. However, the
equations separate out into divergence equations for the scalar fields and
for the vector fields, divergence equations and “curl” equations.
In this chapter, we have observed the effects of several mechanisms
such as player interest, engagement, entitlement, passion, and strategy
690 Geometry, Language and Strategy—Vol. 2

bias. We observe that they play as big a role in decision process theory
as do the concepts of game values, payoffs and Nash equilibrium that we
have taken from game theory. These insights contribute to our
understanding of decision processes.
An important key to understand how to apply decision process theory
to realistic situations is not only to understand the causal equations, but
also to understand the boundary conditions that are part of the problem
statement of any realistic problem.
With some caveats, the approach taken here can be modified and
adapted to incorporate any two-person game with two-strategies per
player. Even though relatively simple, such an extension significantly
advances our understanding. We now deal with a very large number of
games that have already been dealt with in the game theory literature.
With somewhat more effort, we can extend the discussion to games
of three or more players as well as to games in which each player has
more than two strategies. Though decision process theory deals with any
number of players who have any number of strategies, the relevant
question is one of practicality. We are greatly helped by the state of
current technology and its ability to solve partial differential equations.
We have used Mathematica’s numerical method of lines to provide
practical solutions and suggest it can handle even more players and more
strategies. There are currently limitations however, though we
understand work is underway to remove many of them. For example, it
will be of great help to have a general technique to numerically solve the
non-linear elliptic partial differential equations for more general
boundary conditions than purely periodic ones. One promising approach
is the use of finite element method, which has been implemented in
Mathematica for linear partial differential equations.
If we extend our numerical efforts to differing numbers of players or
strategies, what might we expect? As in game theory, the possibilities
become more interesting as we increase the number of players. For
example, it is possible with current computer technology to extend our
analysis to three-person processes as well as four player processes with
active strategies u1  r1  r2 u2  r2  r3 u3  r3  r4  . We expect a
more extensive discussion of which strategies are active, which are
inactive and which are parts of a code of conduct. We also anticipate an
Steady State Geometries in 3D 691

even more complex set of boundary possibilities, making it even more


important that a strong numerical tool be available.
As another example, in dealing with decision processes having three
or more players, we have in common with game theory the attribute that
is related to forming coalitions, [Von Neumann & Morgenstern, 1944, p.
220]. Our discussion and conclusions are quite different however. For
example with a decision process having three agents, assume that the
inactive strategies are the player relative preferences s jk along with the
sum of the player preference scales r1  r2  r3 .
The active strategies u1  r1  r2 u2  r2  r3  provide non-zero co-
moving player strategy bias fields f j ,rm  rn between players that are not
possible with only two players. As in game theory, this appears to reflect
a payoff based on a coalition between players as opposed to the skill
based payoffs. However, we need not introduce mechanisms such as
imputations and dominance that have now fallen out of favor (Cf. Sec.
12.1 and the quote from [Aumann, 1989, p. 13]): here, the field equations
of decision process theory suffice. The only changes are those specific to
identifying which strategies are active and which are inactive.
Imputations and dominance are replaced by the action principle of the
causal theory.
The attainment of the outcomes of this chapter is facilitated by doing
the exercises in the following section. Though not essential for a general
understanding, there is benefit to replicate the numerical examples given.
Based on this investment, the student should achieve the more detailed
outcomes below. Furthermore, though again not required, it is suggested
that the reader carry out as many calculations as they feel comfortable
with to build up an intuition of the interactivity of the many effects
incorporated into the theory; as with engineering disciplines, knowledge
is gained by doing not just by reading or listening.
In addition to these general considerations, the student should have
learned the following outcomes:
 In Sec. 17.1, the student will have learned to identify which regions of
space time are valid and which are not valid or can’t be reached from
a given starting point.
 In Sec. 17.2, the student will have learned the concept of the transfer
functions for the f2AD model, which describes the active geometry.
692 Geometry, Language and Strategy—Vol. 2

 In Sec. 17.3, the student will build on game theory by looking at the
payoffs as fields that describe the active-internal geometry. The
student should be able to understand what boundary conditions are
needed to determine the payoff fields and how they might be set in
any particular model.
 In Sec. 17.5, the student will learn about a component of decision
process theory that acts like an external current, the player passion
field. The student should understand how the vector field can be
obtained from a scalar field and under what assumptions this makes
sense.
 In Sec. 17.5, after considering the geometry of player payoffs, the
student should understand how payoff fields couple to other fields and
create forces. This coupling, the player engagement, has a geometry
that is determined by a scalar field, whose density distribution
determines the surfaces of constant engagement and thus determines
the gradient vector field that results.
 The gradient of the player engagement field is a key contributor to the
player interest. In Sec. 17.6, the student will learn about the shape of
this vector field, both from the model considerations as well as from
more general considerations based on possible and different boundary
conditions.
 In Sec. 17.7, the student will have learned about the player strategy
bias vector field, which generalizes the concept of game value in
game theory. The geometry of this field is determined in large part,
but not entirely, by the player interest in the f2AD model. The danger
of looking at a specific model is that it depends on relative strengths
of competing mechanisms, which with different boundary conditions
will yield totally different results.
 In Sec. 17.8, the student will have learned the internal geometry
behind the decision pressure process, player bonding and player
shear. These are the stresses and strains of the decision process
system. The scalars are described by their surfaces of constant
potential. This leads to an understanding of the associated vector
fields.
Steady State Geometries in 3D 693

 In Sec. 17.9, the student will have learned that the composite payoff,
though small in the model under consideration, is nevertheless not
zero.
 In Sec. 17.10, the student will have learned that the seasonal payoff,
which is an attribute of the geometry considering time as active, is not
necessarily small. Not only is it the source of the harmonic behaviors,
it can also be thought of as being made up of the weighted
contributions of the player payoffs and their engagements.

17.12 Exercises

(1) Show that the determinant of the metric tensor g ab is given in


terms of the frame transformation as indicated in Eq. (17.3). Show
that the transformation to the reduced frame transformation E u n
m


provides the determinant of the metric tensor proportional to
m 2
E u n , as shown in Eq. (17.4). The reduced frame transformation

transforms a “spatial” direction  n in the co-moving orthonormal


frame to a “spatial” direction u m in the normal coordinate frame.
This latter direction corresponds to one of the active strategies.
Show that the inverse transformation E um can be used to determine
n

the inverse of the frame transformations Ea Ea shown below:

E u n E u m   ml ,
l n

 E ul E u  
m l

   nm
E  
 E n E t  
m m t
u
E 
 u  . (17.13)
E n E  t   E t l E  n E   E  l E ui E u 
l l i
t t
 u   u 
 E t E t E t 
 

(2) From Sec. 6.10, Ex. 23, the time component of the metric in the
central frame is
694 Geometry, Language and Strategy—Vol. 2

e2 q
g  . (17.14)
1  e e

Can you therefore conclude that a necessary condition is that the


factor 1  e e  0 ?
(3) Show that the velocity of the normal-form coordinate basis as seen
in the co-moving coordinate basis is characterized as the βn:
E t
  . (17.15)
E t  1  e e

This is the velocity of the coordinate time surface, which defines the

Figure 17.35. Allowed region for βn, in the f2AD model

“now” or “current” time in the normal coordinate frame basis.


Show that this is obtained from the normal-form coordinate time
surface t  x  whose normal to the surface is   t   t . Transform
this covariant vector to the co-moving coordinate basis. This
velocity must be strictly less than unity, since we assume that the
coordinate surfaces travel strictly less than the speed of
communication (“light”). The two frames are communication
Steady State Geometries in 3D 695

compatible if we can move from one frame to another at less than


“light” speed.
(4) As a continuation of the investigation of the requirement that
g tt  0 in Sec. 17.1, show that this is a requirement on the normal to
the “time” coordinate plane. In particular, in the orthonormal rest
 
frame, show that the normal is the vector E t  E t  E t and the
velocity vector can be thought of as the βn, Eq. (17.15). The
constraint is that β  1 and is illustrated in Fig. 17.35. The gaps in
the streamlines also indicate when physical constraints are not met.
(5) The βn of Ex. 3 is determined by the covariant component of the 1-
form X  X  du  for the special case in which X    t is along the
time direction. The 1-form is the line element that moves along the
normal to the time surface: we have normal time. The use of the
word “normal” defines a meaning of what it means to be at right
angles. We have other ways to define “normal”: the metric and
forms. Consider the vector Y    t . We construct the n  1 form
from all the space components with the fully antisymmetric tensor
 1na , which is a hypersurface and is also a candidate for “normal”:

Y  Y  1na du1    du na . (17.16)

The component vector Y  represents a vector that is along the


“normal” to the surface defined by the antisymmetric hypersurface
form. It is orthogonal to the space components: we have orthogonal
time. In the co-moving orthonormal basis, the components are
 
Et Et Et and define a dual velocity *βn.

E t
*   . (17.17)
Et 1  e e

Using the condition g tt  0 and the definition of the metric in terms


of the frame transformation, show that this means *β  1 . This says
that both definitions of “normal” make sense and that each vector,
βn and its dual *βn, must be less than unity. If we start with that
principle, show that we arrive again at the conditions g tt  0 and
g tt  0 respectively. Show that the two velocities, which may be
696 Geometry, Language and Strategy—Vol. 2

equal at one point, are not in general equal. Show that the n-form
X  Y , which is the wedge product of the two forms, does not
vanish. Show that the “dot” product of the components is
X Y   1 . Show that the “dot” product of the two velocities is:

E t E t 1
*      t 1. (17.18)
Et E  1  e e  Et E  1  e e 
 t 

(6) It may seem strange to associate a velocity with an n-form.


However, recall that fluid flow consists of the amount of energy in a
differential volume: see for example Eq. (3.7). Thus the flow-β,
defined as
m

um Eu
 flow  
, (17.19)
E t Et  1  e e

should be related to the dual-*βn. To support this definition, show


that the flow-β is a rotation of the dual-*βn:
m

  um Eu
Et 1  e e  flow  t
E
um
E
Et E u n   t    Et 1  e e  u flow .
n m m
(17.20)
E
  *   E u n
m n m
u flow

2
(7) Show that the flow-β is bounded: βflow   . Use Eq. (17.4) and
2

the fact that the dual *βn is bounded by unity. Can you
determine  ?
(8) Based on Sec. 17.1 and the Exs. 1-7, show that the frame
transformation is characterized by the βn, the dual *βn, the time flow
E t  and the reduced frame transformation matrix E u n , Eq. (17.4).
m
Steady State Geometries in 3D 697

Figure 17.36. βn for the f2AD model.

(9) As shown in these figures, show that βn (Fig. 17.36) and its dual *βn
(Fig. 17.37), may be equal at one point, but are not in general equal
everywhere. Why are the shapes so different?

Figure 17.37. The dual-*βn for the f2AD model


698 Geometry, Language and Strategy—Vol. 2

Figure 17.38. Coordinate streamlines for the f2AD model

(10) To give a picture to the possible topological shapes, consider


Fig. 17.38. We are getting close to the limit to choice. As the
streamlines go around the bend, the outside is getting close to being
the inside and the inside the outside. What is the physical meaning
of having the streamlines touch? Does this correspond to a
singularity (or zero) of g tt ?
(11) The density plot for the code of conduct for the passion field, Fig.
17.39, is essentially centered based on the initial conditions. Are
there any constraints on these initial conditions based on the field
equations, Sec. 14.9.7?
(12) Based on the passion contour plots, Figs. 17.16- 17.17, what will the
vector components look like? Use your understanding to describe
the numerical computations for the f2AD model, Figs. 17.40-17.42.
The figures are not streamline plots but vector field plots. To
simplify the understanding, in each figure, the magnitude of the
vector is restricted to be greater than the indicated minimum. This
should help focus attention on the parts of the field that are large.
You should use Sec. 14.16, Ex. 1 to express the vector field in terms
of the gradient of the passion potential.
Steady State Geometries in 3D 699

Figure 17.39. Code of conduct passion density plot for the f2AD
model

(13) The player interest vector fields, Figs. 17.43-17.45, demonstrate


that the longitudinal components are much larger than the transverse
components. Based on the arguments in the text, what would
generate large transverse components?
(14) Show that the internal geometry is characterized by the player
bonding and player shear. In particular, use Eq. (4.85) for the
gradient of the inactive metric to make your case. What role does
the frame transformation E j play in the geometry?
(15) The composite payoff satisfies its own pair of “curl” and “gradient”
equations, Sec. 14.9.8:

 ω  I α  B α  q  ω
. (17.21)
  ω   Θ  2q   ω
700 Geometry, Language and Strategy—Vol. 2

Figure 17.40. “Blue” passion vector field f2AD model

Figure 17.41. “Red” passion vector field f2AD model


Steady State Geometries in 3D 701

Figure 17.42. Code of conduct passion vector field f2AD model

We suggested that this will have a simplified solution with


ω    in Sec. 14.16, Ex. 2. Show that this is the case. Show
that the integrating factor and equation for the scalar field are given
below:

 ln   Θ  2q
. (17.22)
 2    Θ  3q   Iα Bα

Based on the numerical work done for the f2AD model, in what
region will the source I α B α in Eq. (17.22) be large? Cf.
Fig. 17.46.
702 Geometry, Language and Strategy—Vol. 2

Figure 17.43. “Blue” player interest vector field f2AD model

Figure 17.44. “Red” player interest vector field f2AD model


Steady State Geometries in 3D 703

Figure 17.45. Code of conduct player interest vector field f2AD


model

(16) The form of the equation for the composite payoff in Eq. (17.22) is
valid only for the specialized set of dimensions dealt with in the last
few chapters. The more general form of the equations has been
provided in Chap. 5. One of the equations is the Codacci Eq. (5.36):

          2q    0 . (17.23)

The second Eq. (5.49) is:

     2      q   
    2     q   .
    2      q    0. (17.24)

Show that these two equations can be written in terms of geometric


forms, Sec. 2.2:
704 Geometry, Language and Strategy—Vol. 2

d * ω  *j
. (17.25)
dω  Iα  Fα  q  ω  0

What is the current j in these equations? Prove or disprove the


following conjecture that the equations that generalizes Eq. (17.22)
will be in terms of an n  3 form G :

ω  *dG
d     Θ  2q  . (17.26)
 d * dG    Θ  3q   *dG  Iα  Fα

Figure 17.46. Density plot of the source Iα B α for composite payoff


f2AD model
Chapter 18

Structural Distinctions for


Decision Processes

In the next two chapters, we identify distinctions in decision process


theory that we believe provide increased understanding of the decision
making process. In this chapter we focus on structural distinctions and
in the next social distinctions. Structural distinctions arise in the theory
because of the non-linear nature of the differential geometry, extending
the linear ideas that provided the initial foundation. Because phenomena
are coupled in time and space, they provide feedback loops that form the
basis of the non-linear modeling in the theory. Social distinctions arise
because decisions usually occur between more than just two individuals,
which generate forces that are more social in nature than individual.
The structural distinctions are attributes that are directly reflected in
the differential geometry. An underlying assumption of decision process
theory is that it applies to processes which are both causal in time and
connected in the space of strategic possibilities. These processes have a
degree of continuity similar to other physical processes, which has led us
to apply the tools of topology and geometry.
The structural distinctions that we have in mind are typically not
locally visible. For example the earth is not flat, yet its roundness is not
locally visible. Roundness is the structural distinction. Planets circle the
Sun in elliptical orbits. However, due to the interactions of planets, orbits
can exhibit chaotic behaviors. Chaotic behaviors are structural
distinctions that are not visible in a single orbit.
Decision process theory allows for both curvature and chaotic
behaviors, so it is of interest to identify the structural distinctions that
might apply. As in other physical processes, there are several distinct

705
706 Geometry, Language and Strategy—Vol. 2

mechanisms that mutually interact and are expected to provide insights


that are new and distinct.
We organize the chapter as follows. We start with a discussion of
morphogenesis and structural stability, Sec. 18.1, which gives a general
discussion of what one might expect in a theory built from coupled
differential processes. Events are coupled in time. In particular, events
are influenced by the past, not the future: we have causality, Sec. 18.2.
We use harmonics to characterize the dynamics associated with time,
Sec. 18.3. By working in the frequency space, we treat the system as if it
were stationary with harmonic frequency  that can be varied. Of
particular interest will be events associated with high frequency
harmonics, Sec. 18.4.
We maintain that events are connected in space, which is addressed
by their network connections, Sec. 18.5. An important structural
phenomenon for any physical process is its adherence to symmetry. We
have identified three distinct symmetries in decision processes: the
individual, the code of conduct and the time isometry. We examine the
related notions of individuality, honoring contracts and sustainability in
Sec. 18.6.
A subtle aspect of differential geometry is the notion that the
geometry is characterized by fields, which generate the distinctive shape
we associate with geometry. The field storage of energy in physical
processes, such as electric, magnetic and gravitational field energy,
provides elaborate ways to engineer devices from batteries that do useful
work to circuits that generate resonances. We examine this in Sec. 18.7.
In Sec. 18.8, we return to a notion discussed in earlier chapters, limits to
choice. When couplings are sufficiently large, we suggest that the system
moves from a “flat” behavior to a curved or chaotic behavior. We look
for such behaviors in the theory.
The opposite extreme is examined in Sec. 18.9 where we look at the
system when flat: game equivalency and free fall. The ingredient that
distinguishes these behaviors is the acceleration of the system, Sec.
18.10.
We turn first to the morphogenesis and structural stability of the
theory.
Structural Distinctions for Decision Processes 707

18.1 Morphogenesis and Structural Stability

We assert that decision processes are not stochastic, but connected both
in space and in time: strategic choices made by one group impact choices
another group makes. Despite the obvious uncertainties about what will
happen in the future, there are aspects of the decision process that are
deterministic.
We have pointed out that the resultant equations, which are
hyperbolic partial differential equations, exhibit a mathematical stability
that small changes in the initial conditions result in small changes in the
solutions [Wald, 1984]. We expect that such solutions are characterized
by a structural stability or locality property that the structural character
of a solution does not change dramatically as long as we stay in the near
vicinity of that solution.
The expectations of structural stability are local, not global attributes
of nature or of decision process theory. In nature we have phenomena on
a lake or an ocean that on a calm day, exhibit no wave motion. The
components of the lake are connected. With a small breeze, the lake is
changed only slightly with the addition of small ripples. This is an
example of the expected structural stability.
As the breeze picks up however, the ripples turn into waves that at
some point exhibit breakers, which are not simple extrapolations of the
small ripples. This morphogenesis follows from rather general
arguments, is observed in nature and is expected whenever we have a
theory based on topology and geometry of the type we have been
considering [Thom, 1975].
The observed phenomena of ripples can be used to derive the fluid
equations of fluid dynamics. The equations can be used to extrapolate
behaviors beyond their local origins. The structural anomalies we
observe in hurricanes and tornadoes in meteorology can be simulated
based on these considerations [Friedman, 1989].
We have shown in Secs. 16.9-16.10 that such structural changes also
occur in decision process theory. For example, with only seasonal
harmonics, we get slow rolling standing waves, Fig. 16.22 as seen for the
frame transformation in the f1WWW model. The ripples that we see are
part of an added harmonic that is part of the external force or initial
708 Geometry, Language and Strategy—Vol. 2

condition of the problem. We see hints therefore of the morphogenesis


analogous to fluid mechanics and ripples on water.
We expect that if we add additional harmonics with sufficient
amplitudes, we will generate breakers. We see hints of this in Fig. 16.46
where we have added a second harmonic in the f2WWW model. We have
zoomed into a much shorter time scale and in fact do see the increased
strength of the ripples. We have not yet been able to create ocean like
breaker waves, but believe it may be possible.
Structural instability in decision processes was raised in the classic
work that investigated the limits of global growth [Meadows, Meadows,
Randers, & Behrens, 1972]. However, it is not often the focal point of
economic theories. For example game theory is concerned with the
equilibrium state and so is concerned at most with the small ripples that
occur nearby. It is not concerned with the breakers that occur when the
wind picks up.
We are concerned with such phenomena for a couple of reasons.
First, such phenomena could provide confirmation or condemnation for
the theory. Second, such phenomena may be of practical interest to either
identify or potentially avoid. Like hurricanes, such phenomena may be
destructive so knowing when they are imminent is useful information.
In addition to the arguments noted above based on models from
Chap. 16, we have additional evidence of morphogenesis that was
presented in Sec. 9.5, where we demonstrated that if the stakes in the
prisoner’s game are sufficiently high, we obtain a dramatic change in the
behavior of the decision process value, Fig. 9.15. In Sec. 10.12, for the
Robinson Crusoe model, we found a similar result if we introduced a
large strain, Figs. 10.21 and 10.22.
In all of these examples, when we have a very small disturbance, the
resultant structure appears to be the same: we have structural stability.
However if we increase the size of the disturbance sufficiently, the
behaviors cease to be characterized by a scaled up version of the same
structure. The behaviors are morphogenic. We see new structures that
were not present with only small disturbance to some base starting point.
The study of morphogenic behaviors brings in other attributes of the
theory that themselves represent simple concepts but are now interacting
together in a complex way.
Structural Distinctions for Decision Processes 709

We see in a dramatic way that mechanisms in decision process theory


are connected in space as well as causally connected in time. What are
these connections that change structure? Do we have additional evidence
to support the existence of these connections? To answer these questions,
we inquire further into the decision process theory’s mechanisms that
change structure. We find a set of new standalone features predicted by
decision process theory. We believe a very important one is related to
causality, which we turn to next.

18.2 Causality and Determinism

In Sec. 12.1 we argued that decision processes are causal not stochastic,
despite the fact that decisions involve uncertainty. We argued against the
Bayesian approach [Bayes, 1764] and in favor of an approach looking at
possible decision choices as a frequency distribution that evolves in time
in a causal and deterministic manner. In Secs. 9.9 and 16.11 we
identified recurrence plots as a direct way to identify whether processes
are causal and deterministic. We found examples of what we expect to
see in time sequence plots: for the prisoner’s dilemma, Figs. 9.23 and
9.24; and for the f1WWW model, Fig. 16.27 and the f2WWW model, Fig.
16.51.
We find examples of this
behavior in a variety of
places. First, in the literature,
such recurrence plots have
been used by a number of
authors to investigate chaotic
behaviors. For example
Sabelli [2005] has used such
plots to study heart beats,
whose time sequences might
Figure 18.1. Time sequence for a typical stock be thought stochastic. He
Jan-Dec 2015 showed that the behaviors are
causal and argued that the behaviors indicated chaotic (or what he termed
biotic) behaviors. Second, in stock market data, we find that stock prices
710 Geometry, Language and Strategy—Vol. 2

fluctuate daily. You might argue that the prices are random and unrelated
to each other from day to day. To test this idea consider a stock, where
we have used Wolfram|Alpha® to obtain stock market data [Stock
Market, 2015], Fig. 18.1.
The behavior of the stock price appears to be random elements.
However, that is not necessarily the same as actually being random. We
see causal structure in the recurrence plot for this data, Fig. 18.2, which
has much in common with our model plots. There is a strong correlation
in time: if the data were random, then we would not see any structure in
the recurrence plot. The plots are not however periodic with a simple
period. It is for this reason we see the choppy behavior in the time
sequence plot. We have not attempted to model the stock prices using
decision process theory, but we believe these preliminary inquiries are
encouraging evidence that time is causally connected.

Figure 18.2. Recurrence plot for typical stock time sequence


If we were to model the stock market, the models we have considered
up till now might have to change. We have modeled a single type of
Structural Distinctions for Decision Processes 711

market, whereas the stock market reflects many markets and many
products. We imagine that the stock market reflects processes that go
beyond our simple attack-defense example.
In the real world, a market is not isolated but carries on in a larger
environment. One way to model that environment is to ignore its effects
as reflecting longer time scale phenomena. That is, we replace the real
strategic variables by ones in which the environment effects are absent.
For the stock market we would look at specific stocks and over limited
periods of time.
Only then might we see the free fall behaviors we are interested in.
We would proceed somewhat like scientists investigating weather
patterns by ignoring the spin of the earth (Coriolis forces) as a first
approximation. At some point however we would return to the problem
and add back such effects.
We examine causality with a second example, this time from the
business domain. A common problem in software development is
delivering a complex project on time. We fabricate a scenario below,
though from our experience, it is a scenario that we believe does in fact
happen quite often. We give a self-contained description here of a more
detailed description framed in systems dynamics language, Fig. 13.4.
The delivery of a complex project, one that involves perhaps 500
developers, reflects a large number of decision processes. It is attractive
for management to think of such processes as occurring stochastically
and to believe that the delivery of a complex project will follow the same
pattern of previous projects of the same scale and complexity. We
imagine that company X has gone through 4 or 5 delivery cycles, each
taking 2 years and each delivered (more or less) on time. The projection
of the next delivery is thus expected by management to take the same
time, 2 years. But an unexpected variation occurs; after development has
started, the customer asks for substantial new functionality, increasing
the size of the project.
Let’s say both the customer and company X agree that the increased
size is 20% and that the company agrees to hire 100 more people and the
customer is willing to pay for the increased cost. Management predicts
that the original schedule will thus be maintained. The underlying
thinking is that the delivery process is stochastic in nature.
712 Geometry, Language and Strategy—Vol. 2

The people are hired but to the dismay of the customer and
management, the project ends up costing significantly more than the 100
people and does not come in on time, even though the additional staff
were hired with sufficient time, based on past data, to get the job done.
Both sides looked at the problem from the same worldview, that of a
complex stochastic process. Neither side anticipated failure.
How would we look at this process and would the expected results be
predicted to be different? A causal approach (such as decision process
theory or systems dynamics) would look for key dynamic mechanisms
that govern the delivery of the software product. Using such an approach,
we would agree that hiring the additional people is a necessary
ingredient. However, we would find on analysis and by interviewing
members of the technical staff that not all employees will be equally
productive on the new project without sufficient training.
The previous releases were staffed by essentially the same people
with only a small turnover. Over time, they built up a high skill level
based on the known requirements. The skill level of that staff was
constant: new hires that were needed as replacements were always added
to the project only after adequate training.
To satisfy the new customer request however, staff were added after
the overall project was well underway. The additional 100 people were
employees new to the project and were not yet trained and so had a much
lower skill level. Though many of the employees and middle level
managers complained that they would not be able to deliver based on the
addition of untrained staff, their voices were overruled since the
management model had no way to use the skill level to modify their
schedule prediction. They interpreted the employee complaints as normal
grousing about being asked to go for a stretch goal.
The consequence of the lower skill level was in fact not felt
immediately. The lower skill level introduced a higher level of defects
into the product, but these were not detected early on. When the product
was moved along to the testing phase, the higher number of defects per
unit of code was detected. The defect insertion rate could have been
predicted based on existing data on new hires from past releases.
Because of the worldview of management, however, this was not done.
Structural Distinctions for Decision Processes 713

The company of course was now in the testing phase and had to hire
even more staff to find and fix the defects. They were also not able to roll
off existing staff as planned to other releases. They hired an additional
unplanned 50 people, though in fact that was enough only to fix the first
round of defects. They spent 6 months additional because of this testing.
The testing time could not be shortened since it depended on going
through a fixed number of tests that each took a fixed amount to execute.
Of course the additional people hired were again unfamiliar with the
product and many of their fixes also broke the release. The failure rate of
their fixes was about the same as the other new hires.
The consequence of this problem was the fix-break time cycle. Under
normal circumstances a release was expected to pass all but 5% of its
tests by the time it entered the final test cycle. The test cycle of 6 months
was part of the expected 2 year delivery. That final cycle would reduce
the number of escaped defects to less than 5%, an acceptable escaped
defect rate. The reduction of defects was in fact a direct measure of the
skill level and training of the employees. The new circumstance dropped
that defect reduction percentage dramatically. They saw only a 50%
reduction in defect rate.
For a 50% reduction, as they went into the final cycle, 50% of the
tests would not pass. If they did a second cycle, they would have 25%
that would not pass; a third cycle and 12% and a fourth cycle would be
6%. To get to where they should have been in past releases they would
need a fifth cycle, which is four extra cycles. Each extra cycle cost them
6 months and 50 people. This extended the project by two years and
added an unplanned 50 extra people on the staff, as well as keeping on
the project staff that had been expected to shift to other releases. The
total project cost should have been 500 people for 2 years: 1000 staff
years. The budget office showed that the overall project cost was 2000
staff years and a 2-year delay in schedule.
Now the example above is fictitious but illustrates how even a very
simple exercise of adding a single causal mechanism (staff expertise) has
a profound impact on the outcome. Feedback loops can be very
expensive in time and money. They are also relatively invisible to the
participants of the decision process. The decision about what features to
deliver to the customer followed standard form, albeit late in time. The
714 Geometry, Language and Strategy—Vol. 2

decision about when the project was complete did not change. From a
stochastic view the late project and the normal projects are similar.
Indeed, it is easy to make the assumption that all behaviors are
stochastic. The alternative approach of modeling the behavior as a causal
processes seems harder as it requires identifying the causal mechanisms.
Nevertheless, understanding the causal mechanisms is beneficial to the
company.
Despite being complicated, it has been observed that such causal
mechanisms are well understood by organizations (Cf. Sec. 13.2). In this
example, employees and managers knew that a lower skill level would
lead to a higher defect rate, even in the highly subjective world of
software development. In a stochastic model, there is no place to put that
information. The real difficulty is that the global consequence of such
mechanisms is not easily deduced without the causal framework. We
believe that decision process theory provides just such a framework, not
only for computations but for providing an appropriate worldview of the
problem.
We believe the approach is also superior to a purely systems dynamics
perspective because the approach takes into account not just the effects
of causality and determinism, but also relates these effects to the
continuous network connection aspects of strategic behaviors. We will
return to this concept in Sec. 18.5, but first turn to an example based on
the above example.
For the software problem above, such effects might also be evidenced
by critical exponents analogous to fluids, which indicate internal scaling
symmetries. For example, there is the well-known mythical staff-month
that says if a team of N people can build a small project; then 10N can
build a project 10 times larger. The fact is that it takes more than 10
times the number of staff: you need N 1 . The number  is the critical
exponent, which goes up because in projects, the more people you have
the more managers you need to facilitate communication. Such managers
however don’t build code so the productivity goes down. A theory that
includes the strategic aspects of decision making will include that
[Thomas G. H., 2006]. Such critical exponents reflect the network
connectivity character of fluids.
Structural Distinctions for Decision Processes 715

Our approach is to incorporate sudden changes in time behavior, or


“jolts” in the field behaviors themselves, such as the strategy for use of
resources or the strategy for delivery of product. Changes in time
behavior of the fields can be studied by looking at the frequency or
harmonic steady state behaviors. We turn to this next.

18.3 Harmonic Steady State Fields

Morphogenesis need not just involve causality, but may reflect changes
in shapes as functions of strategic positions. A typical system behavior is
a stable behavior until an external event jolts it. The jolt can be
represented by a characteristic superposition of harmonic steady state
waves. We provided an example of this in our software project example
(Sec. 18.2). The jolt in this case was the addition of new software
requirements. Another example of a jolt might be the meltdown of the
market due to toxic securities.
It is convenient to know the base line behavior of the system, such as
it is in free fall. This does not reflect all aspects of the system however.
Realistic systems are usually more complicated because they have been
jolted. We must include the ability to jolt the system or set up steady
state waves in the system to reflect realistic behaviors. Harmonic steady
state waves provide the dynamic mechanism for studying such behavior.
As the software project example illustrates, a jolt to the system may
not generate the result expected by the players at the time. We must
include other dynamic mechanisms to accurately assess the outcome. In
decision process theory, we take into account these other dynamic
mechanisms. We assign a particular harmonic behavior along an initial
streamline and use the theory to compute the behaviors along all other
streamlines. In this way we consider the network connectivity effects; in
the case of the software project, such other effects were the learning
curve of new hires. This corresponds to a streamline that was not
associated with free fall behavior.
Technically, we are moving away from a description in space and
time to a description in space and frequency. Consider the following. At
each point in space, the flows (and more generally the frames) can be
716 Geometry, Language and Strategy—Vol. 2

represented as a superposition of harmonics. We can decompose the


strategic flow at the still point in terms of free fall harmonics in which
the frequencies are fixed in terms of the payoffs, and more general
harmonic steady state waves in which the frequencies are arbitrary. For
example, to the attack-defense model Sec. 16.4, we add harmonic steady
state waves to create the f2AD model, Sec. 17.2, with two harmonics.

Figure 18.3. f2AD model flow vectors

The initial model represents the system prior to adding a “jolt”. In the
f0AD model, we define the still point behavior as that which occurs when
there is no active acceleration Q  0 , Eq. (15.31): collectively the
absolute forces, competitive forces and cooperative forces cancel. We
now add the jolt, so that the coordinate flows, Fig. 18.3, show the effect
of the added harmonics. The seasonal harmonic is not visible on the scale
of this figure as the frequency is small compared to the added harmonics.
The coordinate behavior along a streamline, Figs. 17.38 and 18.4,
reflects a helix-like behavior.
We convert this time picture of the dynamics to the frequency picture.
For each frequency, we have a phase shift, Fig. 17.5 and modulus, Fig.
17.6. These pictures contain the same information as the time domain
picture. The advantage of looking at the frequency domain is that the
entire problem appears steady state; of course this is deceptive because
the actual behavior is obtained by adding the frequency components,
Structural Distinctions for Decision Processes 717

with their phases, moduli and sinusoidal wave factors, to obtain the time
dependence.
The “jolt” in time is translated into a set of frequencies that must be
included, which in turn generate for each frequency the amplitudes and
phases. However, the time picture is also deceptive since it is not always
easy to compute or envision what happens after a given “jolt”. The
frequency picture provides a complementary approach to the problem.

Figure 18.4. Coordinate plot u a for f2AD model


For example, we can categorize effects depending on their time scale.
If we consider effects that have a very long time scale compared to the
seasonal payoff then we have a harmonic frequency that is small
compared to the seasonal payoff frequency. Such low frequency waves
might be things to ignore or might be significant hints of things to come.
On the ocean, low frequency rollers can indicate oncoming severe
storms. We would need to know the amplitude of the effect as well as the
frequency. Low frequency effects in the market might be long lead time
effects; if they reflected boom-bust cycles they could represent stormy
financial times.
718 Geometry, Language and Strategy—Vol. 2

Of course, low frequency phenomena might be something we ignore.


Do we have good evidence that that represents a good decision policy?
Global warming is an example of something that has a long cycle time.
Those that argue that global warming is significant may still believe that
the effects are a long way off. They may also agree the effects are
currently small but growing. The issues that are still in dispute are
whether the amplitude will continue to grow and whether that growth is
caused by humans.
Whether significant growth is caused by humans or not, observed
large amplitude indicates changes in policy that will be needed to avoid
serious consequences to the global economy; so one criterion for low
frequency phenomena to be important is that their amplitude is
sufficiently large to be of significance. In comparing different
phenomena we can compare the product of the amplitude and the
frequency. As the frequency goes down the amplitude must go up to
make an equivalent effect.
Global warming is an effect that grows over time. There are effects
however that are entirely cyclical and long term. An example is cyclical
or seasonal market demands. At the end of the year, sales go up
dramatically because of holiday buying. This skews the monthly market
figures for members of the retail industry. It obviously changes their
business strategies because the demand (amplitude) during the holiday is
large. Such cyclical changes are important even though they don’t
generate exponential growth.
Shorter term fluctuations would correspond to high frequency
harmonic steady state waves. These would be frequencies large to the
seasonal payoff frequency.

18.4 High Frequency Harmonics

In this section, we address some examples of high frequency harmonics.


Since we are equally able to discuss the dynamics in the time domain or
the frequency domain, the challenge will be to identify dramatic form
changes such as chaos and turbulence in the frequency domain.
Structural Distinctions for Decision Processes 719

In the market place, high frequency harmonics arise for decision


process “jolts” that are short lived such as fads. It is clear that a robust
dynamic picture requires inclusion of such effects. Such effects can
dominate or mask other behaviors. Often, the effects are large. If
sufficiently large, they may generate morphogenic changes.
For the travelling waves in Robinson Crusoe economics, Sec. 10.10,
we investigated the consequences of large frequency, Figs. 10.9 and
10.10. We found that high frequencies generate harmonic behaviors in
space, as long as the limits to choice are not too severe. We can see this
effect in the attack-defense model.

Figure 18.5. Initial flow for f1-hf AD model


For the f2AD model, see evidence of the travelling waves in the phase
gradient of the transfer function, Fig. 17.5. However, these waves don’t
manifest in the actual flows because the limits to choice are too
restrictive. We consider a modified f1-hfAD model with a single high
frequency component   10 . It still has a restrictive limit to choice, but
clearly demonstrates a travelling wave Figs. 18.5 with a wave number
near 10, Fig. 18.6.
The operative measure of importance for short term fluctuations is the
product of the frequency and the amplitude. Since the frequency is large,
even a small amplitude fluctuation can make an effect. For example
consider the potential impact of advertising. If ads are run, they can
change attitudes and market, even when short. It is the frequency of the
ads that makes an impact. This fact is observable during presidential
720 Geometry, Language and Strategy—Vol. 2

elections. The ads need not address the issues so in that sense their
frequency is much larger than the seasonal payoff. Other dynamic
mechanisms will determine how these effects propagate. Our view is that
advertising changes the narrative.

Figure 18.6. Travelling wave phase gradient along z for


f1-hf AD model
In general, we expect to see a variety of effects: seasonal frequency
due to game equivalency, short term fluctuations and long term
fluctuations. Returning to the stock market behavior example (Fig. 18.1),
we imagine that the short term fluctuations reflect response to daily news
events and information about individual stocks. The long term
fluctuations might reflect trends in foreign markets. On top of this the
general market behavior would reflect growth or decline based on the
success of the market: the seasonal behavior.
We suggest that the addition of forced oscillations does not change
the definition of the still point based on game equivalency, only the
frame of reference. We view the harmonic steady state travelling wave
change of reference as changes to the frame transformation. The frame
transformation generates fictional forces that modify the free fall
behavior.
For the f1-hfAD model, the forced oscillation might represent cyclical
behaviors outside of the context of the time frame of the attack-defense
scenario. In decision process theory, forced oscillations drive the system
Structural Distinctions for Decision Processes 721

and reflect a behavior with modified oscillations that characterize the


underlying system behaviors. This has been exploited by others as well
and is the basis for the Systems Dynamics analysis by Meadows,
Meadows, Randers, & Behrens [1972] of the limits to growth in the
global economy.
As a musical example of forced oscillations, consider the motion of a
bow over the strings of a violin. The sound that results is more than the
oscillations of the strings. The sound reflects the character of the violin
itself, so that one gets a sound from a Stradivarius that is more
extraordinary than from an average violin. Thus starting from an input
set of harmonic flows, we get Figs. 18.3 and 18.4, which is more than the
input. It is the superposition of the resultant amplitudes along with their
phase shifts. This is the “rich” tone that is produced from the input.
To verify that we are on the right track, we look for empirical
periodic behaviors. We know that there are numerous cycles in business.
There are bust and boom cycles that have historical time frames
measured in decades [Wikipedia, 2012] down to corporate project release
cycles measured in weeks or months. In considering free fall behaviors,
we should consider such behaviors as key characteristics of the decision
process.
Certainly one aspect of these different cycles is their length. The
length is inversely proportional to the frequency and the frequency is
proportional to the payoff. Assuming the same proportionality constant,
high payoffs would correlate with short cycle time lengths. Low payoffs
would correlate with long cycle time lengths. We would be forced to
consider their relative cycle lengths a dynamic feature of decision
behaviors and correlate this with their payoffs along with the appropriate
proportionality constants.
From the last model, we see that the dynamics is influenced by such
things as the limits to choice, which is determined by the size of the
payoffs and the size of the bonding strengths. The dynamics is thus
dependent on the stresses and strains of the system and how they
interrelate.
A focus on causal mechanisms leads us to look not just at decisions
as events in the time domain, (or equivalently frequency domain), but to
look at the stresses and strains that underlie such mechanisms. We look
722 Geometry, Language and Strategy—Vol. 2

at effects that may slow down or speed up decision processes. In the


software example, Sec. 18.2, the skill level of new hires slowed the
decision process. An outcome of our inquiry into decision processes
using decision process theory is the identification of many such
mechanisms and their corresponding distinctions. We turn next to those
that are associated with network connectivity.

18.5 Network Connections

Decision processes occur in a societal context, so that decisions are


connected to each other not only across time but across strategic
distance. The limit to choice (Sec. 18.8) is an example. We have
envisioned the societal context as that of multiple groups doing similar
things at about the same time. In this way the network is similar to a fluid
in physics. However, one significant difference between human
behaviors and classical fluids is that humans remember.
Thus the same individual may see similar situations a variety of times
and respond not just on present circumstances but on those events that
have happened in the past. In this way the individual sees a network
based on his or her experience as well as on a network based on the
observations of what others are doing currently. We allow for both
meanings of network in what follows. Both contribute to the societal or
network connectivity.
The significance of network connectivity [Barabási, 2003] is that
such connectivity plays a role in behavior. It is deterministic not random.
The internet provides an example of the importance of network
connectivity. What we do or say influences what others do or say. When
the effects are small, the impacts are local; as the effects increase we
expect the impacts to increase proportionally. We expect this linear
relationship to break once the increase is sufficiently large. We expect
that the network distinctions give rise to morphogenic changes such as
phase transitions for sufficiently large changes.
We have found some examples of network connectivity. Diverse local
behaviors can combine to form new and unexpected global structures
such as Figs. 18.9 and 18.10. We have computed other examples of this
Structural Distinctions for Decision Processes 723

in Sec. 9.5 where high stake and low velocity behaviors generated
trapped behaviors, Fig. 9.15. These network distinctions are new and
don’t arise in game theory. They are specific to theories such as decision
process theory. They are also expected to be observed in the market
place.
As a business example, when Japanese practices dominated
manufacturing because of adoption of Deming’s management method
[Walton, 1986], manufacturing in the United States was not unaffected.
Indeed those practices returned to the United States and caused an
effective revolution in quality for US made products. Adopting
techniques currently being used by someone else is a network
connection. We are not waiting for another decision to be made by
ourselves; we are mimicking what someone else is doing because they
are in a similar space at a neighboring point in time.
These are gradient effects over the network, which propagate
causally (Cf. Exs. 9, 10, 16, 17 and Figs. 18.31-18.33). An analogy is the
pressure gradients in weather. The flow of air reflects the causal nature
of the weather. We effectively describe the weather however by
considering not the flow of air but rather the areas of high pressure and
low pressure. The pressure gradients (as seen over distance) from high
pressure to low pressure generate the forces, which help us understand
the causal motion of the air. These gradients reflect the elastic network
connectivity of the atmosphere.
In a similar fashion we look for pressure gradients in decision
processes (e.g. Ex 15). They measure the network connectivity forces.
We look for evidence of network player bonding of strategic areas.
Areas that compress should be under higher pressure than those that
don’t. The ability of a strategic area to resist player bonding, its bounce,
is a measure of the elasticity of the decision process in that area.
Decision processes reflect a complex scenario of network interactions.
The measure of network connectivity will be based on the change
between players executing a particular strategy and other players
executing a neighboring strategy.
Again as a business example, when a company comes out with a
revolutionary and successful product, we might well believe that it is the
result of causal activities that happen solely within that company. That
724 Geometry, Language and Strategy—Vol. 2

product might reflect new technologies or new uses of existing


technologies developed by that company. The product is revolutionary in
that it is not a mirror of any existing product. Once released however, we
often see many companies simultaneously imitating that product. This is
reflective of the network connectivity rather than a causal train of events
in each of these companies based on their own innovation and design.
Our use of the distinction of pressure gradient fits in with our general
framework of units, Sec. 2.1. We think of strategies as the effort that
goes into making a choice and utility as the value of that choice. Utility
and effort are not equivalent in decision process theory, analogous to
stress and strain, respectively, in physics; each generates the other
through dynamic mechanisms. We have identified three mechanisms that
change utilities (Secs. 12.3, 12.4 and 12.5): pair-wise competition, pair-
wise cooperation and organizing sources. Each are associated with
forces that change behaviors and create high density and low density
behaviors: capital accumulation and capital loss. Network connectivity
is a mechanism for change that is a consequence of these
capital fluctuations. Accumulated capital tends to be spent and lost
capital tends to be replenished.
Capital, including experience, therefore represents stress that is either
constructive or destructive. For example, constraining the public’s
liberties generates stress, which we normally consider destructive. A
company’s creative ability generates wealth, which is a stress that we
normally consider constructive. Stress causes things to happen. Capital
must be created or spent. Deprivation of freedoms can stimulate
revolutions. Failures can stimulate growth. Creativity creates new
opportunities. If there are stresses we should ask what efforts will now be
undertaken and what strains will now occur. The loop must be closed.
Stresses cause strain just as strains generate stresses. This leads us to the
following hypothesis about stable structures.
Successful decision structures are built on three key attributes:
effort, utility and capital. They correspond to the definition of success as
the three W’s: work, wisdom and wealth [Wikipedia; Henry Wriston,
2012]. We characterize successful decision structures as those that are
sustained and steady state, as those in which each player’s efforts realize
their vision (value or utility) and those in which the network flows result
Structural Distinctions for Decision Processes 725

in locked behaviors with capital accumulation corresponding to the


collective visions. These structures can be identified in the frequency
space and define the successful dynamic structures.

18.6 Individuality, Honoring Contracts and Sustainability

Symmetry is an essential structural element in nature. We see examples


of symmetry wherever we turn, be it a pretty gem or the structure of a
flower. Symmetries are observable and don’t always have an easy
underlying explanation. To observe symmetries, we observe the lack of
change or variation where one might be expected. In geometry, the
surface of a sphere is characterized by a radius that does not change as
we move on the surface. Away from equilibrium, we expect strategy
choices to fluctuate; if we see no change this may reflect symmetry. We
believe this may be seen in manufacturing where, for example the quality
delivered (a strategy) is apparently immune to improvement despite
management’s best efforts.
We have relied on the concept of symmetry in decision process
theory for the distinction of an individual. The individual’s internal
strategies and utilities have no direct impact on the strategic outcomes. In
other words, the payoffs and strategic flows don’t depend on an
individual’s internal preferences, only on the actions by the individual.
Such actions are not internal but external.
A collection of individuals may act as an individual if their
preferences and strategies are internal, not external. We make these
concepts precise by identifying the strategy of an individual  j as an
isometry of the theory; the strategy of a collection of individuals y m in
like fashion can also be an isometry. We can say that a collection of
individuals with an isometry strategy are honoring a contract. This
works exactly like an individual, albeit with a social aspect (Sec. 19.1).
A third example of symmetry in decision process theory is time
isometry. This works just like honoring a contract, only over periods of
time rather than along a specific strategic direction. It is a choice: we
relate it to the idea of sustainability.
726 Geometry, Language and Strategy—Vol. 2

The idea of travelling and


standing waves are specific
implementations of this: they are
used in electrical engineering to
describe the behavior of
transmission lines, for example.
This is a structural distinction that
persists in time, not unlike the
Persistence of Memory Fig. 18.7. It
Figure 18.7. Persistence of Memory by
Salvador Dali [2015] is not a requirement of the theory
any more than a code of conduct is
a requirement of the theory; yet it produces desirable structures that
organize decisions.
We have defined an isometry as a transformation that leaves the
metric field invariant: there is a frame of reference in which that field is
independent of the corresponding strategy (Sec. 4.2), though this is not
true in all frames.
To each isometry, there are three major consequences:
 To each isometry there will be a corresponding payoff field F j ab that
depends on the isometry direction  j and the active (external)
strategies a b . Such fields act like Coriolis forces and generate
rotations.
 To the isometry directions, there will be scalar fields  jk whose
gradients act like centripetal (bonding) forces that are repulsive or
attractive.
 To each isometry there will be a conserved charge V j that acts like a
conserved flow. Its value does not change over time.
Thus individuals, collections of individuals that honor a contract and
collections of individuals that pursue sustainable contracts have the
same structural behaviors. We expect to see evidence of such behaviors
by seeing the existence of their corresponding fields: their payoffs,
bonding fields and conserved charges. Currently, we don’t have any
reason to expect that such isometries exist, only to observe that when
they do exist they have these properties.
We have already pointed out some of the payoff behaviors that we
expect (Secs. 17.3 and 17.10), bonding behaviors we expect (Sec. 17.8)
Structural Distinctions for Decision Processes 727

and the conserved charges (player engagements, Sec. 17.5). We believe


that it is noteworthy that in decision process theory, the distinctions of
honoring a contract and sustainability are at the same level as an
individual. They play identical roles and so in some sense have the same
importance. This is different from game theory for example, where the
individual and competition play a more central role than say cooperation
and sustainability.
Of course we have the flexibility to create model differences.
Individuals need not be the same; different collections of individuals
need not be the same; and individuals can be different from a collection.
We saw examples of some of this in Sec. 17.5 with the f2AD model. Note
the striking difference here between the collection Fig. 17.20 and either
of the individuals “blue”, Fig. 17.18 or “red” Fig. 17.19. Furthermore,
note that “blue” and “red” are quite distinct.
To understand any decision process, we believe it is essential to first
identify these symmetries. Who are the individuals? What contracts are
in force and who controls the terms of the contract? Are the structures
sustainable? Just as individuals control certain strategies, collections of
individuals control certain strategies. There will certainly be overlaps.
The symmetries though hidden are not without consequences. We see
their consequences if we follow the energy of the system.

18.7 Energy Storage—Capacitance, Inductance and


Gravitas

One consequence of sustainability in decision process theory is the


existence of a conserved charge, which we call the energy of the system.
Though the distinction is an analogy from physical theories, the concept
is also a distinction of decision process theory. Since the energy has
many possible forms, the consequence of the conservation law is that
energy will appear to move from one form to another. We make use of
the underlying common mathematics to use concepts from electrical
engineering and physics to help our understanding of how this energy
moves from one form to another in decision process theory.
728 Geometry, Language and Strategy—Vol. 2

Based on the presence of strategy bias fields and player payoff fields
in decision process theory, energy storage is related to the ideas in
electrical engineering circuits. In that domain, engineers have accepted
the idea that electric and magnetic fields are real and can be used to build
useful circuits. We can’t take these ideas at face value however. We must
modify these ideas somewhat since our decision systems don’t have all
the attributes of a circuit.
Fortunately, not all systems that are studied in electrical engineering
are circuits either. Engineers do study transmission lines and view them
as distributed element systems [Wikipedia, Distributed Element Model,
2015]. Maxwell’s partial differential equations define such systems.
They lead to the types of travelling and standing waves we have been
discussing in decision process theory. These equations are based on the
existence of electric and magnetic fields, whose behaviors are not always
easy to deduce.
Circuits, by contrast are lumped element systems. For such systems
[Wikipedia, Lumped Element Model, 2015], it is assumed that the
energy and momentum as well as all of the electrical properties are part
of a lumped component such as a resistor, capacitor or inductor. These
components are connected by idealized wires with no property other than
they carry current. More complicated components can be added without
changing this basic assumption. There are two advantages: the
approximation accurately reflects what engineers want to build and the
equations are ordinary differential equations that are significantly more
tractable. These advantages provide a means for understanding the
distributed element systems.
In some sense, the distinction between lumped elements systems and
distributed element systems goes back to Newton and Huygens on the
nature of light: is it particle or wave. The current version of this
argument is whether matter is particle or wave. Each description has
advantages giving insight into the nature of things. We have taken one
point of view that the field (wave) is the more basic.
This discussion is relevant to decision process theory in that we too
are solving complicated partial differential equations. Moreover, the
mathematics of the equations generates Maxwell like equations. We have
electric and magnetic fields. We have in addition, metric components
Structural Distinctions for Decision Processes 729

that generate gravitational fields. Though we are not proposing to use the
lumped element approximation, the concepts may give useful insights.
The lumped element approximation provides a physical picture of the
objects that store energy. For example, resistors dissipate energy and so
relate to heat and loss; capacitors store electrical energy; and inductors
store magnetic energy. Each of these has an analogy in decision process
theory. In a dynamic system, energy can flow from the strategy bias
fields to the player payoff fields, which corresponds to storage of energy
in a capacitor like object to an inductor like object. Energy can be lost in
the stress fields (pressure fields), corresponding to resistance.
The additional possibilities however are not part of the electrical
engineering analogy: there can be energy stored in the gravitational and
metric fields. This includes a type of magnetism (Coriolis force) as well
as a scalar gravitational field. These are related to the way in which
utility is ascribed to different strategy possibilities.
This approach has been used by others to understand the behaviors of
fluids [Wikipedia, Mobility Analogy, 2015]. The lumped element system
effectively becomes a systems dynamics problem. Not surprisingly, the
stocks that store things in systems dynamics are ideal distinctions for how
energy is stored. In addition to finding analogs of the electrical
engineering concepts of heat, electrical and magnetic energy in fluids,
these studies find properties that are specific to fluids: one such concept
is inertance: it stores the inertia of the fluid based on the fluid’s
compressibility.
Based on such general considerations, we expect that energy in
decision process theory will be stored in specific subsystems as follows.
 Energy will be stored as electric energy and there will be identifiable
capacitive subsystems that accomplish this storage. We expect energy
in this subsystem to be identified with the player strategy bias fields,
or game value.
 Energy will be stored as magnetic energy and there will be inductive
subsystems that accomplish this storage. We expect energy in this
subsystem to be related to the player payoff fields.
 Energy can be stored as metric field energy and there will be gravitas
subsystems that accomplish this storage.
730 Geometry, Language and Strategy—Vol. 2

 Finally, energy will be stored as fluid compressibility energy and


there will be inertance subsystems that accomplish this storage. We
expect energy in this subsystem to be related to the energy density of
the decision process.
Though these distinctions merely restate concepts from the full
theory, they may be valuable in building decision structures as one does
in systems dynamics [Richmond, 2001]. There are structures in business
and economics that recur and can be used in any system to accomplish
specific goals.

18.8 Limits to Choice—High Bias Example

Players in a decision process may engage more or less actively as a


function of dynamics. For example, they may engage strongly based on
their player engagement. They may act more often than other players as
evidence by the relative player scale, Sec. 14.2. They may see more
value in the game than others as evidenced by their player strategy bias.
Any of these factors may influence the outcome of the decision process.
Any of these changes may make a structural change in the behavior if the
changes are sufficiently large. We shall investigate a few possibilities. To
act at all requires an act of will.
Let’s consider the relative player scale: if one player acts more often
as indicated by frequency of making decisions, we say that that player is
acting aggressively. The opposite of this behavior is that the player is
accommodating: they are acting with the same frequency as the other
player(s). To act at all requires an act of will and so must require some
aggressive or accommodating behavior. If there is no act of will, the
player is acting passively.
There is more to dynamics than this of course. In the theory, a player
is engaged if their player engagement is not zero. Their player
engagement can be positive or negative, weak or strong. If the player
engagement is zero, we say that the player is entitled: whatever they do
has no influence on any possible action.
For the attack-defense model, the player engagements are almost
identical for “blue” and “red”, Fig. 16.4. In order to obtain locked
Structural Distinctions for Decision Processes 731

behavior, Sec. 16.3, we increased the player strategy bias for honoring
the contract to f 3  5 , which has the effect of separating the “blue” and
“red” player engagements. When “blue” acts more aggressively,
Fig. 18.8, we see a charge structure showing morphogenetic changes.

Figure 18.8. Player charge for “blue” when “blue” is


aggressor for f0AD model
We see similar effects by looking at the phase space plot of the three
player engagements. For “blue” being the aggressor Fig. 18.9, we see a
shape that is very similar to the one for “red” being the aggressor, Fig.
18.10. In the f0AD model, the major difference is in the sign of the player
engagement for the code of conduct.
We see somewhat different behavior for the prisoner’s dilemma,
Secs. 8.8.2 and 8.8.3. In that model, we assume that initially, the two
prisoners have zero engagement: they are entitled. As a consequence of
the dynamics, if one player is the aggressor, then the charge for that
player is negative and the charge for the other positive. This is not the
case for the attack-defense model. We expect that in general, the details
will depend on the specific model chosen.
We expect that the results reflect common sense. Not everyone can be
the aggressor and not everyone can be the accommodator. The world
does not naturally separate into groups of people who will always be
732 Geometry, Language and Strategy—Vol. 2

accommodating and groups who will always be greedy. Because we live


in a social network, there are limits to how greedy or how
accommodating we act in a given situation. That should be reflected in
decision process theory.

Figure 18.9. Phase space plot of player engagements


for “blue” as aggressor in f0AD model
There will be limitations in the harmonic solutions where growth is
limited to be linear in proper-time. There will also be limits in the
stationary aspects of the solutions, as illustrated here. The limitation can
be seen in the elasticity of the decision process system. In an elastic
system, stressing the components by displacing them (generating strain)
produces a bounce of the system as a response; the system changes other
parts in a continuous manner in such a way as to balance out the applied
force. We have applied an input force as a consequence of setting the
boundary conditions and now observe the consequences to the system in
these figures. The elasticity provides a limitation to greedy behavior or
accommodating behavior by either party.
The limits to greed or accommodation appear model dependent,
though there may be a neutrality principle: there are limits to the amount
of polarization. In decision process theory, strong player engagement is
associated with greed or accommodation depending on the numerical
Structural Distinctions for Decision Processes 733

sign of the player engagement. An element of the distinction of the limits


of greed (accommodation) would be the existence of a neutral point of
zero player engagement that attracts other behaviors.

Figure 18.10. Phase space plot of player engagements


for “red” as aggressor in f0AD model
The physical analogy is that in general, matter that occurs in nature is
electrically neutral. Though it is possible for matter to be strongly
polarized, that is not its “natural” state. On the other hand, even though
matter is neutral, we can’t ignore the effects of polarization. We see these
effects as radiation, magnetization, electrical permittivity etc. We don’t
confuse limits on polarization and ground state neutrality as a statement
of the dynamic condition. The existence of a neutral point is the source
of the elastic dynamic behaviors.
We have demonstrated the consequences of locked behavior (Sec.
16.3), which supports this neutrality principle. We have solutions that
have zero acceleration and player engagements at one point. At this
point the solution concentrates the energy density supporting the notion
that the preferred place to be is at this neutral point. So for example if
each player has a positive (negative) player bias, the generalization of
game value, then when that player is more aggressive, then the player
engagement is negative (positive).
734 Geometry, Language and Strategy—Vol. 2

18.9 Game Equivalency and Free Fall

A general property of differential geometries is that locally, they


represent a space that is flat. This is an important structural distinction
for a general class of theories that includes decision process theory as
well as many current theories of the physical world.
In these theories, at each point we can always find a frame of
reference in which all the curvature effects vanish. For a dynamic theory,
this means there is a frame of reference in which there will appear to be
no acceleration and the space will be flat at that point.
Morphogenic changes can be expected whenever the curvature at a
point is sufficiently large. Even if we go to a frame in which space is
locally flat, curvature will be evident in the second order effects: at least
one of the second order gradients of the metric field will be large.
Decision process theory is also built on the principle that at each
point, the strategy space is locally flat: locally, one can find a coordinate
system in which we appear to be at rest and there appears to be no
acceleration. We distinguish this case by saying that in that specific
coordinate system, we are in free fall at that point. However, this is only
true to first order in the coordinates.
The acceleration effects we are discussing are frame rotation effects
and are of second order. The space is in fact curved, not flat, just as the
earth appears locally flat but is a globe. In decisions, we see a
connectivity that is not only local but global so that we use differential
geometry to stitch together the local pictures into a consistent global
picture. The orientation flux fields accomplish this by stitching together
adjacent regions. As a consequence, our central holonomic frame, which
provides a global view, in general is not locally flat everywhere, Ex. 29,
Sec. 6.10. In that frame, we are not in free fall at every point.
We have seen that decision process theory admits solutions that are
perturbations of game theory equilibrium behavior. The game theory
distinctions are worth repeating as they are not exactly the same as the
decision process theory distinctions.
In game theory, decisions are characterized by pure strategy choices
or mixtures of pure strategy choices reflecting some frequency
distribution. In contrast, the still point and free fall behavior is
Structural Distinctions for Decision Processes 735

determined by the way in which one sets the initial conditions. Those
conditions could be to require certain equilibrium values or could be to
require certain initial player values (Chap. 14).
What we can say is that for free fall, there will be a mixed strategy
choice that sets the rate at which each strategy changes. There will also
be a seasonal payoff matrix that will determine the resonant frequencies.
Additional dynamic mechanisms from the theory determine the detailed
relationship of these free fall values and initial boundary condition values
for the players.
For the player fixed frame models, using the definitions and Eq.
(15.30), Ex. 1 from Sec. 15.15, we can write the frame equations Eq.
(4.32) as Eq. (15.31). From these equations, we glean that for behavior at
the still point, we have zero active geometry acceleration (Fig. 18.11):
the still point behavior is then set by   :

q  2 e   e e 
Q 
1  e e 

 2

1
  1  E k  Ek 
1  e e
      1 2 Ek f k 
    1 2 e  q f   e    (18.1)
      e        1 2 e  q  f 
 E a   0
 E a    E a 

This represents the simple structure analogous to ripples on a lake.


Specifically, when there are no tidal forces (free fall) and no charges,
the flows are constant and the spatial components of the frame rotate
with components  1 2 e  q  f  , the seasonal payoff. The seasonal payoff
generates a Coriolis force that reflects a rotating central co-moving frame
(Cf. the form of the metric, Ex. 23, Sec. 6.10). Moreover we see that the
absolute force is also zero. If the composite payoff is zero, the seasonal
736 Geometry, Language and Strategy—Vol. 2

payoff is the weighted average of the player payoffs and corresponds to


the figures in Vol. 1.
The assumption of no tidal forces can be imposed at any point by
choice of gauge: it reflects seasonal payoff acceleration or Coriolis
forces. Due to second order effects, this will not be true at all points as
viewed in the same reference frame. For our numerical examples, it is
approximately true on the initial boundary, not just at the still point,
Fig. 18.11.

Figure 18.11. Active acceleration Q  x, y ,0  for f0AD model

We don’t expect this to hold away from the initial boundary. For
example it might not be true along an initial streamline, even though
along that streamline we can take the metric to be Minkowski and make
simple model choices for the orientation potentials, so that they are small
or correspond to a static gravity field. Indeed, these choices describe the
basic assumptions or models used in Vol. 1. They provide a baseline
structure on which we can look for new behaviors.
For example, in Vol. 1 we ignored the fact that the inactive scalar
fields and the payoff fields may vary along the streamline (Cf. Ex. 30,
Sec. 6.10). We ignored the second order effects. If we include these
Structural Distinctions for Decision Processes 737

effects, the free fall harmonic behavior is a good approximation only on


the boundary. Away from the boundary, the additional effects will
impact the non-zero harmonic behaviors (Fig. 18.12). When the second
order effects are large, then “ripples” can become “breakers”. We obtain
the possibility of “breakers” from the acceleration effects. A complete
description requires the full set of equations with the appropriate
boundary conditions.

Figure 18.12. Active acceleration Q  x, y , 0.1 for the f0AD


model
We have noted that decision processes behave in a way analogous to
electrical engineering systems. We suggest the analogy that there is a
seasonal impedance and a load impedance on which one imposes a
forced behavior from a source: the seasonal payoff reflects the network
attributes of the system; the harmonic steady state waves (Sec. 18.3)
reflect the source; the load is determined by the initial conditions. To
fully understand the system we need to look at both behaviors.
Let us start with the boundary conditions (on the initial surface z  0 ,
for example), which provide one of the behaviors. The theory will have
solutions in which there is a still point (Sec. 15.1), which will appear as
738 Geometry, Language and Strategy—Vol. 2

having no active geometry acceleration forces: these are the sum of the
absolute force, competitive or game theory force and the cooperative
force. Game theory identifies such behavior as Nash equilibrium.
Though there are differences in how game theory and decision
process computes the individual player behavior, both theories provide
essentially the same distinction. We use the insights learned from game
theory to set the initial conditions in decision process theory. We do this
while remaining consistent with the ideas of causality, without resorting
to Bayesian probability as a predictor for future behavior, which is also
consistent with at least one view of game theory [Von Neumann &
Morgenstern, 1944]. The important difference between game theory and
decision process theory is that the latter has second order effects, which
are expected to generate morphogenic changes. These changes result
from the field equations and the specific initial conditions. Different
initial conditions in general lead to structures that are not the same; they
are not diffeomorphic.
Now let us deal with the second behavior, which is provided by the
initial conditions (which in the time domain would be at t  0 ). These
conditions provide the applied force, which generates the acceleration
effects that change the geometric structure. The acceleration effects may
manifest themselves as flows that change in time or as Coriolis effects in
which the flows appear fixed, along with fictitious forces that appear,
reflecting the fact that the description is in a frame that is rotating.
To take into account both behaviors, our strategy has been to choose a
frame of reference so that at the still point there is no acceleration. Away
from that point there may be acceleration. The numerical results and the
mechanisms away from the still point depend on the details of the
acceleration.
The details of the acceleration are generated by the forced harmonics.
We look at the harmonics (Sec. 18.4) in the central holonomic frame,
Sec. 4.6, which is related by (a non-diffeomorphic) linear transformation
to the normal-form coordinate basis. We examine the model in this
frame and then transform the results back to the normal-form coordinate
basis.
In the central frame, the active strategy flows are independent of time
(indeed, in the central holonomic frame for the fixed frame model, the
Structural Distinctions for Decision Processes 739

space components are zero and the time component is unity) while by
assumption, the covariant rate of change of the flow with time is not
constant. This is consistent with the perception that such effects appear to
be absent.
To summarize, there are two mechanisms, each leading to distinct
harmonic behaviors: the free fall harmonics and the forced harmonics.
The latter are harmonic steady state wave behaviors (forced behaviors),
with arbitrary frequencies, and replace a time domain picture of the force
with a frequency domain picture (Sec. 18.3). The former free fall
harmonic structure generates resonant harmonics based on the seasonal
payoff. We may get morphogenic changes when these behaviors are
combined. Therefore, it is useful to separate out the equilibrium
behaviors as an important structural distinction.

18.10 Acceleration

Our goal is to compare decision process theory structures to real world


processes such as the stock market behavior, Fig. 18.1, which have lots
of fluctuations. If we interpret them as random or stochastic, we deny the
possibility that the behavior is causal. If we argue that the behavior is
causal, then an alternate interpretation is that we are seeing multiple
harmonics. We are seeing evidence of forced harmonics and
acceleration, which we take to be a structural distinction.
Though the stock market behavior is complicated, it is encouraging
that it appears to indicate the presence of forced harmonics. The market
responds as if it was an elastic-object that has been struck. Specific
events do appear to cause the market to oscillate and reverberate long
after the event. That we can imagine that the market can be “struck” is
evidence that we are receptive to the notion that there are strategic flows
that change in time.
The outcome of the decision process consists of the actual choices
made. The rates at which these choices are made represent the decision
flow and are something that we directly measure. For example, we might
survey each player ahead of making their decision. The survey questions
would provide the current state of each player prior to making the
740 Geometry, Language and Strategy—Vol. 2

decision. We could ask in terms of frequency, how they would rate their
choices. The survey is not predictive in the probability sense but provides
information about the initial state; decision process theory makes
assertions about how this state evolves in time: how a later survey might
be answered.
In this way the theory generalizes the static flow considerations of
Game theory to steady state flow and dynamic flow. Since the flows
need not be constant, there will be acceleration defined as the rate of
change of each strategy flow. We expect the flow of decisions to change
in direction and magnitude. These changes provide a direct means to
measure the new distinctions in decision process theory.
We distinguish between the player’s worldview of payoffs that they
use to make their decisions and the seasonal worldview that reflects the
decisions that actually have been made. In free fall when there are no
forces other than this seasonal worldview, the strategy choices reflect
rotations. Indeed we argue more generally that the concept of player
payoffs and seasonal payoffs reflect some type of rotation.
Qualitatively, we generalize Nash equilibrium. If players act on the
basis of their individual payoffs, then Nash argues that there will be an
equilibrium choice that is in some sense best in that no player can do
better because of what other players would do in response. The Nash
equilibrium is thus an optimal strategy for the seasonal payoff.
Conversely, if we know the Nash equilibrium we imagine a variety of
rotations that leave this direction unchanged, any one of which is a
possible seasonal payoff. We say that the collective behavior will always
be a rotation, so there will always be a seasonal payoff that measures that
rotation. For any rotation, there will always be directions that are left
fixed. Decision choices along such a direction correspond to Nash
equilibriums.
Acceleration is what takes us away from this equilibrium.
Acceleration behavior is converted from the time domain to the
frequency domain with the steady state harmonics (Sec. 18.3). These
harmonics will exhibit either travelling wave or standing wave behavior,
subject to limits to choice. We expect the acceleration effects to be
captured by the transfer functions. The transfer function behaviors obey
Eq. (14.36) and so depend on only a few of the co-moving frame tensors:
Structural Distinctions for Decision Processes 741

the time metric, the acceleration vector q , the player bonding Θ , the
time metric e 2 q 1  e e  and the seasonal vector potential a .
1

Figure 18.13. f0AD model of the time metric factor

The time metric factor, Fig. 18.13, vanishes near the boundary. The
coefficient of the linear term is proportional to  2 e 2 q 1  e e  4aa  .
Since the zero frequency solutions are unbounded, we expect that we
may get unbounded solutions near the boundary, which is whenever this
factor changes sign. This is what we find from the numerical
calculations. A large negative acceleration potential accomplishes the
same thing.
We have computed the acceleration potential previously, Fig. 17.26,
where we see that there are no areas where the potential is large and
negative. The other possibility would be singular points due to the player
bonding, whose bonding potential is Fig. 17.27, which is well-behaved
for this example. Finally, the seasonal vector potential is very small,
corresponding to a small seasonal payoff, so is also expected to play a
small role. The main variations that remain are due to the frequency
variations. We expect that very high harmonics lead to high frequency
spatial behaviors.
742 Geometry, Language and Strategy—Vol. 2

The acceleration in any frame is based on transforming the


acceleration q from the co-moving orthonormal frame: E a q . For a
q
single (phasor) sinusoid e i e , the acceleration in the frequency domain
is determined in terms of the transfer function, Sec. 17.2:
i e q   H a q  H a q q  . A complete analysis in the frequency domain
would then require we study the behavior of the transfer function as a
function of frequency and position, which would then determine the
acceleration. We note that a sudden impulse in time typically translates
into a frequency dependence H a   1 , which needs to be verified.

Figure 18.14. Initial co-moving frame acceleration q  x, y ,0  for


f0AD model

To get the acceleration in any frame we need the acceleration in the


co-moving orthonormal frame q . It will be based on the assumed
acceleration on the initial boundary surface, Fig. 18.14. We have
computed q z  0,0, z  previously, Fig. 16.1. To understand what we might
expect, we argue as follows.
As decisions are made, we expect that choices in some strategic areas
will be denser than in other areas, Fig. 17.26. This generates pressure,
Fig. 17.25 and player bonding Fig. 17.27 that pushes future choices away
Structural Distinctions for Decision Processes 743

from those areas. See the player bonding  z  0,0, z  , Fig. 16.3 as an
example.
The behavior of those choices would appear to accelerate from the
high pressure areas to the low pressure areas. Such effects generate
changes in momentum, a term often used in sports and in the stock
market. We are familiar with team behavior improving suddenly as if
they pick up outside support. We see their fortunes accelerate.
The opposite of acceleration is inertia. In the f0AD model, we have
assumed a fairly high decision mass, md  40 which corresponds to
significant inertia. Being slow to change reflects large inertial effects. In
fact there might be confusion over behavior that has settled into
equilibrium versus behavior that is not changing because of inertia. We
think of companies that are not making a profit but are set in their ways.
There may in fact be winning strategies that they could adopt if it were
not for the inertia and lethargy of the organization.

18.11 Outcomes

We have examined possible structural distinctions in this chapter that are


typically not visible when the focus is local. Our understanding of
physical processes as well as business and economic processes relies on
our ability to see and agree on local behaviors. However, we believe that
it is difficult to put disparate local facts together into a consistent global
whole without a unifying theoretical framework.
Based on this chapter, the student should have learned a number of
structural distinctions that might arise as we move from local behaviors
to global behaviors. Some of these behaviors might be related to chaotic
structures. Other behaviors relate to other types of breakdowns that
would not be expected from a linear extrapolation of local structures.
The attainment of the outcomes of this chapter is facilitated by doing
the exercises in the following section. Though not essential for a general
understanding, there is benefit to replicate the numerical examples given.
Based on this investment, the student should achieve the more detailed
outcomes below. Furthermore, though again not required, it is suggested
that the reader carry out as many calculations as they feel comfortable
744 Geometry, Language and Strategy—Vol. 2

with to build up an intuition of the interactivity of the many effects


incorporated into the theory; as with engineering disciplines, knowledge
is gained by doing not just by reading or listening.
In addition to these general considerations, the student should have
learned the following outcomes:
 From Sec. 18.1 the student will have learned the concepts of
morphogenesis and structural stability and how they provide a
general framework for what to expect in theories based on differential
geometry [Thom, 1975].
 From Sec. 18.2, the student will have focused on the causality aspects
of decision process theory. The student will have learned that
causality itself may be an unexpected structure for some when applied
to decision processes. Certainly a more striking possibility is the
existence of chaotic behaviors.
 From Sec. 18.3, the student will have learned that many structural
features can be observed in the frequency domain as opposed to the
time domain. The two domains work collaboratively in providing
insights. The student should be able to distinguish seasonal
harmonics from forced harmonics. Forced harmonics can be used to
drive the system and helps illuminate the free fall harmonics. A
general solution is a superposition of harmonic steady state waves of
both types, and is set by the appropriate initial conditions at t  0 .
This sets the relative weights of the superposition. A system can be
studied by driving it with forced steady state wave harmonics, which
will illuminate the underlying free fall harmonics. The student will
have learned that in going from the time domain to the frequency
domain, one needs to be able to determine the frequency behavior of
the transfer functions.
 From Sec. 18.4, the student will have learned to identify high
frequency harmonic structures and their relationship to short time
frame events.
 From Sec. 18.5, the student will have learned the implications that
events are not only connected in time but in space. This gives rise to
network connections that are as important as those in time [Barabási,
2003]. These distinctions are associated with gradient effects:
pressure, network player bonding, bounce and elasticity. The student
Structural Distinctions for Decision Processes 745

will have found support for the three W’s: work, wisdom and wealth.
These could be the basis for successful decision structures.
 From Sec. 18.6, the student will have learned that symmetries
underlie a key set of structures. In decision process theory they
determine what it means to be an individual, what it means to honor
a contract, and what sustainability might involve. In general,
symmetries are not derived from a theory but may exist in the theory
as aspects that are part of the empirical framework. Once observed, it
is the job of the theory to represent them.
 From Sec. 18.7, the student will have learned that the fields we have
in mind are real and carry energy that can be stored and can be part of
replicable components of any process. In a lumped element
approximation in analogy to electrical engineering, one can imagine
capacitors that store electrical energy (capacitance, which in decision
process theory would correspond to value), inductors that can store
magnetic energy (inductance, corresponding to utility through
payoffs), gravitational energy (gravitas, cooperation and player
bonding, as resonant energy) and fluid mechanical energy (intertance,
stresses due to players that are not explicitly included).
 From Sec. 18.8, the student will have learned that there are not only
limits to growth, but limits to choice. This distinction places a
structural constraint on the size of the spatial geometry. It helps
distinguish how a system moves from a “flat” behavior to a curved or
chaotic behavior. The student will have learned some of the
mechanisms that cause this to happen.
 In Sec. 18.9, the student will have learned the importance of
identifying the areas that are flat, in which the behavior is free fall
and most closely aligned with game theory. Decision process theory
will have solutions in which behavior along a path will have a still
point (Sec. 15.1) at which the active geometry acceleration is zero.
This generalizes the notion of Nash equilibrium from game theory
and allows the use of game theory as a baseline for characteristic
decision behavior. A special case of still point behavior is one in
which the strategies appear to be in a flat space. Free fall behavior is
determined by the seasonal harmonic behaviors with frequencies set
746 Geometry, Language and Strategy—Vol. 2

by the seasonal payoff. Such frequencies are properties of the specific


decision process.
 From Sec. 18.10, the student will have learned the opposite of free
fall, which is when acceleration effects are dominant. The
acceleration is the rate of change of the strategy flow. The student
should have learned that we expect to see acceleration in decision
processes; we expect the flow of decisions to change in direction or in
magnitude. These changes may provide the most direct means to
measure the new distinctions in decision process theory.

18.12 Exercises

(1) Modify the attack-defense model f2AD (Sec. 16.13) to the f2AʹD
model in which the payoffs will be changed for “red” to Eq. (16.3)
as suggested in Sec. 16.3:

 0 0  2 5  310 813160 
 0 0  110  2 5 813160 

Fab     2 5  813160  ,
1 blue 1
10 0 0
 3 
 10
2
5 0 0  813160 
  813  813160 813160 813160 0 
 160

 0 0 0 1
10  301 60 
 0 0 1 0  301 60 
 5

Fab     0 60  .
2 red
 15 0 0 301 (18.2)
 1 
  10 0 0 0 301
60 
 301 0 
 60 301 60  301 60  301 60

Show that these changes introduce strong strategy biases. Include


the following changes to Eq. (16.20) at z  0 :
Structural Distinctions for Decision Processes 747

md  4
B 3   110 0 1

10

f 3
z  10
. (18.3)
e1  e2  0
1  2  2  e
H a k  0,0,0   2 10 3
10
4
10
5

10

Show that collectively, these changes generate limits to choice.


What is the Nash equilibrium vector corresponding to this model?
(2) For the f2AʹD model for Ex. 1, both “blue” and “red” are entitled in
the sense that their initial charges are set to zero in Eq. (18.3).
Explain the behavior of the player engagements in Fig. 18.15. Note
that we have not set the engagement to zero for the code of conduct.
What causes the charges to be opposite in sign away from the initial
boundary? For positive aggression, “red” is acting as expected but
“blue” has an engagement that is opposite in sign.

Figure 18.15. Player engagements for the f2AʹD model

(3) Structural stability and morphogenesis can be found in a number of


different areas. One cause of morphogenesis might be a local
behavior accommodating to a boundary condition, as shown in
Fig. 18.16. Here we have the motion of a cylindrical shape moving
near the square boundary. What would you expect if the boundary
were circular?
748 Geometry, Language and Strategy—Vol. 2

Figure 18.16. Coordinate phase space for the f2AʹD model

Figure 18.17. Frequency dependence of the f2AʹD model transfer function


Structural Distinctions for Decision Processes 749

Figure 18.18. Phase behaviors for the f2AʹD model

Figure 18.19. Recurrence plot for the f2AʹD model relative strategy
component
750 Geometry, Language and Strategy—Vol. 2

(4) Based on the initial conditions, Eq. (18.3), the player aggression
components f2AʹD model have a common factor of 4 10 and two
additional frequency components. We see in Fig. 18.17, the
frequency dependence of the transfer function for both the seasonal
resonance frequency and the forced frequencies. Plot the modulus
as a function of frequency. Add more points to this plot to obtain a
feeling of the overall frequency dependence. What does it tell you
about the time dependence?
(5) For the f2AʹD model, the forced harmonics are travelling waves, Fig.
18.18. Because of the limits to choice, the attenuation factor
 p  1.801 , Eq. (16.17) is close to the forced harmonic frequencies.
What happens if we choose this as the first harmonic  1  1 2  as in
the f2AD model?
(6) For the f2AʹD model, the recurrence plot, Fig. 18.19 demonstrates
causality and determinism. Discuss the differences of this plot to
that of the f2WWW model, Fig. 16.51.
(7) For the f2AʹD model, think of the co-moving frame transverse active
strategy z as a measure of player aggression. The gradient term is
seen to decrease the flow V u as z increases, Fig. 18.20. What
actions do you expect “blue” to take as a consequence? Compare
your answer to the model results, Fig. 18.21.

Figure 18.20. f2AʹD model initial strategy flows


Structural Distinctions for Decision Processes 751

Figure 18.21. Time dependence of the f2AʹD model strategy flows

(8) For the f2AʹD model, we have computed the contour flows,
Fig. 18.22. Because of network connections, argue that the
recurrence plots for the player aggression will change as you
increase z , the relative player aggression. Based on the contour plot
provided, how will the recurrence plots change?

Figure 18.22. Player aggression flows for the f2AʹD model


752 Geometry, Language and Strategy—Vol. 2

Figure 18.23. Gradient aggression strategy flow for the


f2AʹD model

Figure 18.24. Network connections recurrence plot for aggression


preference strategy for the f2AʹD model
Structural Distinctions for Decision Processes 753

(9) Another way to analyze the network connections is to look at the


gradient plot, Fig. 18.23. How is this related to the contour plot of
the flow, Fig. 18.21? Hint: refer to Sec. 6.10, Ex. 8.
(10) Argue that you get information about the network connections from
the recurrence plot such as E u z , Fig. 18.24. For example, argue the
change in the aggression flow due to changes in the aggression
preferences must be continuous in time. What else can you learn
from this figure?

Figure 18.25. Time metric factor for the f2AʹD model


754 Geometry, Language and Strategy—Vol. 2

Figure 18.26. Initial acceleration gradient for the f2AʹD model

Figure 18.27. Acceleration potential for the f2AʹD model


Structural Distinctions for Decision Processes 755

(11) To honor a contract or to engage in a sustainable strategy, we


expect that the corresponding strategies will not change in time:
they are conserved. This remains true even if the system is subjected
to a shock. An example might be the quality of a product produced
by an organization. It is a recognized principle that such
organizations maintain their quality (or lack of quality) despite
shocks to the system [Walton, 1986]. Can you find examples of
this? Is a denial of global warming an example, irrespective of
whether global warming is true? Does this also apply to what it
means to be an individual?

Figure 18.28. Initial active acceleration vector for the f2AʹD model

(12) The f2AʹD model is an example of high bias that generates limits to
choice. An example of the impact of high bias is provided in Ex. 3.
Another example is provided by the density distribution of the time
metric factor 1  e e , Fig. 18.25. Compare the structure to
Fig. 18.13.
756 Geometry, Language and Strategy—Vol. 2

Figure 18.29. Active acceleration slices for the f2AʹD model

Figure 18.30. Pressure for the f2AʹD model


Structural Distinctions for Decision Processes 757

(13) Compare the initial values of the acceleration gradient for the f2AʹD
model, Fig. 18.26 to that of the f0AD model, Fig. 18.14. Why is the
former so much larger? As a second comparison, look at the
acceleration potential for the f2AʹD model, Fig. 18.27. Why is the
active acceleration so much smaller for this model, Fig. 18.28?
(14) For the f2AʹD model, the active acceleration becomes large near the
boundaries, Fig. 18.29. Explain why this comes about.

Figure 18.31. Recurrence plot for stock price ratio (dB) between
two similar companies
(15) There are strong player biases for the f2AʹD model, leading to a
change in the pressure, Fig. 18.30 compared to the f2AD model from
Sec. 17.8, Fig. 17.25. Explain the quantitative increase in pressure
with the new model and the change in structure.
(16) In Sec. 18.5, the argument was made that not only should we find
causal relationships for fields, but also for neighboring values of
fields, such as Fig. 18.24 and Exs. 9-10. Consider two similar
companies such as Nokia and Alcatel. In our view, the difference of
their stock prices reflects a price gradient over nearby strategies. We
758 Geometry, Language and Strategy—Vol. 2

would expect to see correlations in the recurrence plot for such a


difference. Is there evidence for this in the data [Stock Market,
2015] Fig. 18.31? Here we plot the recurrence plot of the ratio
expressed in units of dB, which is 10 Log10 priceNOK priceALU . Is
there a “jolt” in the system around the start of the second quarter of
the year?
(17) Using the same stocks as in Ex. 16, consider a restricted time
interval, the mean price priceNOK price ALU , Fig. 18.32 and the
price ratio priceNOK priceALU , Fig. 18.33. The price ratio appears
to oscillate after the price shock in November. Is this statistically
significant evidence for there being a “jolt” to the system causing
harmonic oscillations as suggested in Sec. 18.3?

Figure 18.32. Mean price of two similar Figure 18.33. Price ratio of two similar
companies companies
Chapter 19

Social Distinctions for Decision


Processes

In this chapter we continue our identification of distinctions in decision


process theory that increase our understanding of decision making
processes. We focus here on social distinctions as opposed to purely
structural distinctions. We again use specific numerical models to
understand these new distinctions, returning to the work-wealth-wisdom
models from Sec. 16.8 as well as the attack-defense models from
Sec. 16.2.
We set the framework for identifying these distinctions in society. We
do more, since many of these distinctions are known; we show how they
fit together as part of a unified whole. They don’t operate independently
but together.
We start with code of conduct, Sec. 19.1. This covers behaviors that
are preserved over time that are both cooperative and sustainable. In Sec.
19.2, we explore the concept of entitlement, in which a player acts alone,
empowered without regard to other player behaviors. Though it is
egoistical it is necessary since it allows for such things as creativity. This
is the first component that governs a player’s choice. In Sec. 19.3, we
explore what happens when a player influences or is influenced by other
players through player engagement. This coupling to the composite
payoff is a new dynamic distinction. The player engagement payoff is
the product of the player engagement and the composite payoff. It is the
second component that governs a player’s choice.
In Sec. 19.4, we explore social forces that cause the player
engagement to change: the player interest, which generalizes the notion
of value from game theory. Changes to the player payoff are governed by

759
760 Geometry, Language and Strategy—Vol. 2

a variety of forces, one of which is the player passion, Sec. 19.5: A


player’s passion for a particular strategy may overpower objective
choice and determine entitlement.
The theory allows and requires cooperative behaviors, starting with
mutual player support, Sec. 19.6: decision processes are governed in
part by player cooperation driven by mutual player validation forces,
which play as important a role in governing behaviors as do competitive
valuation forces. This depends in part on a matrix of potentials which
describe the strength of bonding between individuals. Such bonds
provide gradient forces dictating how players respond to cooperative
forces. Gravity and centripetal acceleration are both examples of bond
forces. We suggest there are similar forces in decision processes.
Finally, in Sec. 19.7, we consider a dynamic distinction: hidden-in-
plain-sight. In decision process theory, we expect that some harmonic
steady state waves may be aspects of cycles that are hidden-in-plain-
sight, cycles that need to be incorporated systematically into the solution.
Until now, such effects are unseen; distractions we remove when we
attempt to see the big picture. For example, they may be long cycles,
which help to hide them.

19.1 Code of Conduct

Strategy is closely associated with the notion of something that is


executed or owned by a specific individual. We break from that idea
when we allow more than one individual to determine the action for a
particular strategy. An extreme version of this is when that choice no
longer influences any of the values of the fields that determine the
player’s behaviors. When this happens we have a code of conduct
symmetry (Sec. 18.6).
The technical aspects of this symmetry provide additional insights. In
the centrally co-moving holonomic frame, Sec. 6.7, the central time is
inactive and for code of conduct directions m , n , the competitive force
along them vanishes for each player j : F j mnV n  F j m V   0 . This
follows from the fact that in this frame, the metric elements are
Social Distinctions for Decision Processes 761

independent of time, the code of conduct strategies and the player


internal strategies, Sec. 15.6. Hence the player payoff F j mn between any
two codes of conduct is zero and the player bias F j m for a code of
conduct in this frame is zero. These characterize the code of conduct
since in this frame, the active flows are zero, V   0 .
Having the code of conduct distinction as an embedded component of
the theory is a significant departure from game theory and current
economic theories (Sec. 11.2). In decision process theory, a code of
conduct provides preserved and conserved cooperative behaviors. It is a
possible symmetry of solutions; any decision process may have such
solutions. We use symmetries to define what a player is and extend the
concept to define cooperative behavior where certain strategies are
removed from the competitive context by agreement of the players (Sec.
18.6).
Though not an embedded concept in game theory, this idea was
envisioned as necessary for a complete theory of games by Von
Neumann & Morgenstern [1944, p. 511ff.]. They identify discriminatory
solutions, which impose many of the same constraints as we have in
mind for the code of conduct, though not limited to the negative
connotations of their word.
Their idea has not been generally adopted as a necessary part of game
theory and does not seem to have been exploited in the literature. We
think the concept is worth revisiting since we think decision process
theory resolves some of the drawbacks that were found with the game
theory approach. In particular, the current theory is by construction self-
consistent, which addresses one of the concerns of the previous approach
[Luce & Raiffa, 1957, p. 206].
The existence of a code of conduct changes the essentially greedy
nature of game theory moving the focus away from the behavior of an
individual’s decisions. It puts the agreements made between players,
honoring a contract, on the same footing as other components of the
theory. It is a possible and stable way to play the game. As noted by
Adam Smith [1776], it is a requirement for the invisible hand (Sec.
11.2.1).
The type of cooperation we have in mind is more along the line of
coalition rather than consensus (Sec. 12.4). All players benefit in a
762 Geometry, Language and Strategy—Vol. 2

solution in which contractual agreements are honored, despite the


possibility that individual gains might be larger without such agreements.
The contract must be honored even for those choices in which a
significant advantage would be obtained if the contract were violated.
Therefore, every player builds the symmetry into their idiosyncratic view
of payoffs.
However, it is not a requirement in the real world that contracts be
honored and it is not a requirement in decision process theory for there to
be symmetries. Nevertheless, there are stable and perfectly acceptable
solutions in which symmetries exist and in which there are codes of
conduct that match real world behaviors.
We have explored in detail the prisoner’s dilemma (Chap. 7), which
provides an example of a code of conduct: in this case a code of silence.
We think the possibility of such coalitions occur frequently. In Chap. 14,
we used the overall player preference scale as the code of conduct:
behaviors are driven by skill not by the sum of resources expended by all
participants.
This code of conduct generalizes the notion from game theory of a
zero-sum game. The players agree that the overall effort played is not to
be considered as part of strategic or cooperative payoffs. It may make
some difference in terms of hidden degrees of freedom that carry energy
and momentum, but is not to be explicitly considered. For each player,
their player payoff matrix will reflect this symmetry: the total preference
scale is a conserved charge, yet there will be gradient forces and payoff
forces associated with its direction.
The total preference scale behaves as a fictitious player as in Figs.
18.8-18.10. In the f2WWW model, Fig. 19.1, we see that the total
preference scale appears to have a dependence on both strategy and time:
the total preference scale charge however is constant along each
streamline as required by the conservation law.
Our view is that decisions occur not only in a survival of the fittest
context but in a societal context. Human beings are social beings as well
as competitive beings. That means they are as likely to follow ethical
norms as to work towards maximizing personal gain. The code of
conduct symmetry provides an important dynamic mechanism for
Social Distinctions for Decision Processes 763

incorporating this societal context. It exists in the real world and so it


should be part of any theory describing that world.
We suggest that solutions not only be utilitarian but just [Tavani,
2011]. Of course this is not a requirement. Just as in engineering, we can
build structures that serve human needs and those that don’t; decision
process solutions may equally meet or fail to meet human needs. The
choice of solutions remains subjective.
As an example, consider two global companies. Though not a
solution in our sense and not an ethical rule [Tavani, 2011] since all
players have not adopted the same code of conduct, one company might
create and adhere to a code of conduct in which bribery is expressly
forbidden, even when that company operates in countries where it might
be common practice. The other company might not have such a code of
conduct, or at least its code of conduct remains silent on this issue.
Both companies might prosper, though possibly in different ways.
The first company would seek revenue opportunities in countries where
taking bribes is not practiced, or might if practiced succeed because of a
superior product. The second company might seek revenues everywhere
and turn a blind eye if bribes are required and employees comply.
As stated, the proposed codes of conduct are not in fact symmetries.
We imagine however that there may be forces that encourage the players
to adopt the same set of codes. Once this happens we have an ethical
standard, which is what all players have bought into.
In realistic solutions, it is hard to imagine that there won’t be some
set of strategies that will be excluded from the competitive arena. In the
market place or even in war there always appear to be rules of the game
that have been adopted and are followed. A code of conduct is not just an
interesting dynamic possibility but a required dynamic mechanism for
describing realistic decision processes.
An important step in studying any decision process is to identify the
operative codes of conduct, even though it may be hard to always
identify such strategies. The strategies may be hard to identify because
they are uninteresting precisely because by agreement, they are not part
of the strategic choice process. This partitions the strategies into those
that are active from those that are part of the code of conduct and
inactive.
764 Geometry, Language and Strategy—Vol. 2

A criticism of this approach might be the commonly held view that


few if any codes of conduct are adhered to all of the time. You may argue
that any set of ethical rules, regardless of how noble, are usually violated
from time to time. Depending on the rules, the violations may be small to
large. The attempt to curb alcohol consumption during prohibition was
met with scant success.
However, deviations from the expected mechanism in fact provide a
strong support for the dynamic point of view that we take. Whenever
there are large deviations, the strategy in question is in fact not part of
the code of conduct. There may be other behaviors that are in fact obeyed
but not generally recognized as conserved. The code of conduct is a
symmetry mechanism. It provides reasonable solutions only when it is a
good approximation.
We can always remove the symmetry constraint and study the
behavior of the system over time assuming the system is adjusted to
perfect symmetry on some boundary condition. We are then in a position
to study whether the symmetry is reasonably stable or dramatically
unstable. In effect we study the effect of the other dynamic mechanisms
on our initial symmetry.
More generally, it is our view that each dynamic mechanism has an
area of applicability in which the mechanism is clearly visible. All
mechanisms interact through the theory, which predicts breakdowns. Our
example of the late project (Sec. 18.2) was an example of both a
mechanism for delivering the project on time and the impact of late
requirements that generated a significant delay in delivery.
Thus a given strategy can be a candidate for symmetry, but when
realistic boundary conditions are applied, we may see that the boundary
conditions fail the symmetry or that there are forces preventing the
symmetry on the boundary to be maintained. The theory allows us to
simulate possible behaviors to learn how one might engineer the decision
process so that desired codes of conduct are enforced and sustainable.
Finally, in looking at various ways in which players may cooperate
and hold certain strategies apart as codes of conduct, we imagine that
they can cooperate to build societal behavioral structures that are
sustainable, Sec. 18.6. They may accomplish this in practice by simply
letting all transient behaviors die out. Such behaviors will be dynamic
Social Distinctions for Decision Processes 765

but steady-state. Time is then the symmetry variable. For such behaviors,
we expect to find as evidence, field behaviors that are constant when
viewed from the appropriate point of view: namely when viewed in a co-
moving frame in which the active strategy flows are zero.

19.2 Player Entitlement

We use player payoff in different contexts, depending on the frame of


reference. We make an invariant distinction for the player payoff based
on the centrally co-moving orthonormal frame, Sec. 6.10, Ex. 19. We
explored this idea in Sec. 15.4 in the context of entitlement in game
equivalency, Eq. (15.8), where we showed that the co-moving player
payoff has two components:

f     2     2 e   . (19.1)

Our interpretation is that each player has two views of the decision
process corresponding to the two terms in the equation.
In one view the player makes choices depending on the collective
behavior represented by the composite payoff,   . In the other view, the
player makes choices depending purely on their own worldview,     ;
they are empowered without regard to other player behaviors.
In a positive sense of the word, they act with entitlement (Sec. 15.4).
They believe they have a right to make their own decisions and manage
their own interactions. They are focused and forcefully state their
position in making their decision. They are in the “zone”. Entitlement
might be viewed as a self-centered and egocentric component, though
without the intent to necessarily harm others. Entitlement just doesn’t
take into account the other point of view; they only see the other through
their private payoff map.
In our models of a societal framework, the entitlement effects are
steady-state, though as with the charge, Figure 19.1 and other steady
state fields, the entitlement effects will be seen as time and space
dependent in the normal form coordinate basis. The private payoff map
is the entitlement payoff in Eq. (19.1), 2  , which in the 3D notation
is 2Bα .
766 Geometry, Language and Strategy—Vol. 2

The co-moving player payoff f    consists of this contribution and


the player engagement payoff 2e   , which in the 3D notation is
2ωe . On the initial boundary, this latter contribution is zero or small:
zero at the still point and small on the remaining part of the boundary.
This is because at the still point, we set the composite payoff to zero,
ω  0 , Sec. 15.11.

Figure 19.1. Code of Conduct charge behavior for the


f2WWW model

Away from the still point, we don’t expect the composite payoff
to vanish. We close the loop and express the composite payoff in terms of
the seasonal payoff and the entitlement payoffs of all of the players,
Eq. (14.30):

    1 2 e  q f   e     . (19.2)

It is reasonable to consider the composite payoff as a social distinction as


it depends on all of the players. It does so only if some of the player
engagements are not zero. Therefore it is not a stretch to see that the
engagement payoff is also a new social distinction, depending as it does
on the composite payoff. We discuss this further in Sec. 19.3.
So why do we think of the player entitlement as a social distinction?
First, the player entitlement is the private map of what each player thinks
Social Distinctions for Decision Processes 767

each of the other players will do: it is that player’s view of the payoffs.
As an example, for the f2WWW model, the player payoffs for “work”
Fig. 19.2, and “wealth” Fig. 19.3, represent and extend the game
equivalency behaviors.

Figure 19.2. “Work” entitlement payoff in the f2WWW model


They represent each player’s private map on the initial boundary;
they extend that private map as a steady-state map over all of the space
based on the theory’s feedback loops. Of course these feedback loops are
not entirely private, but depend on the other player behaviors as well.
Based on the model calculations, we also observe that the square
boundary is reflected back onto the player payoff behaviors. Since the
boundary and the behavior of others are social, we conclude that the
player payoffs are in part social as well.
Second, as an example, one of the players in the f2WWW model is in
fact the total player effort coalition, representing a code of conduct. This
highlights the point that the effect of the code of conduct is not just a
conservation law, but may include payoffs that are on the same footing
as the other player payoffs.
768 Geometry, Language and Strategy—Vol. 2

Figure 19.3. “Wealth” entitlement payoff in the f2WWW model

We expect that in every decision process, each person puts forward


their best argument in support of their personal view, which is based on
how they expect others to act as a consequence of their own actions.
Each player’s personal view is their worldview as expressed in the
payoffs they expect. Because we take a dynamic view, that payoff may
change with time or across the network. In the steady-state view in the
co-moving frame, we are left with the network behavior only. The
possibility of a learning curve is a necessary attribute of entitlement.
In some extreme cases only the entitlement view is present. Consider
political parties as players. It is not uncommon for a political party to
portray only their view and make decisions solely on that basis. In other
extreme cases, only the other view is present. Consider a family. The
“ideal” parents take only the “other” view. They try to do what is best for
the children, sometimes at the expense of their own welfare. In general, it
may take effort to identify the component of the payoff that truly
represents the player’s current best interest; or it may be impossible if
both are equally present.
Social Distinctions for Decision Processes 769

This type of split of the player payoffs seems to have no parallel in


game theory. The split is blurred by the process of looking for an
equilibrium state. It is clear that a dynamic split must exist, that the split
benefits from the learning curve. If we return to the example of the
political parties, there is no long-term future for either party to
stubbornly maintain their entitled position and hope at the same time to
pass legislation if they exist in a two-party system. Moreover, even if
they have a stubborn ideological split, it may not be on all issues. So
either their entitled position evolves, they adopt a position (or work in a
strategic area) that has some characteristic support or some combination
of both.
Let’s return to the example of the software project in Sec. 18.2. The
customer and the company start with a view that is probably aligned.
They both expect the delivery of the project in two years and they both
agree on the price of the project (customer) and the cost (company).
They have differing entitled views however on things that don’t directly
concern the other. The company expects to roll off its developers onto a
new release that will be for a new customer once the project is complete.
The customer has hired marketing and a rollout crew to introduce the
new product to its customers. This rollout is predicated on the delivery
date. If the delivery date is met or is close to being met, neither party has
an issue or concern about the other’s entitled view, since they have built
in a little slack in their schedules based on their own past experience.
The introduction of the new features and the addition of new costs
and pricing were based on their past experiences. Neither party changed
their fundamental notion of what would happen. The company was
probably the first to realize that its view of the world was wrong. After
all it was not going to deliver on time. It was hiring more people than
anticipated and was forced to rearrange its development schedule for
other customers. It had learned the error of its ways; the new hires were
not prepared to deliver as planned and the number of hires made was
insufficient for the task.
The customer however was going on without much warning that
things were not proceeding as planned. Its worldview remained
unchanged. It was getting reports that there were a few glitches but the
company was doing everything it could to deliver. So the customer
770 Geometry, Language and Strategy—Vol. 2

proceeded to do its hiring and getting its marketing campaign in place.


At some point however the customer was told that delivery was not just a
little late; it was very late. The customer’s view was now badly out of
whack with reality. The learning curve of exactly what had transpired
was now fully operative.
Both the customer and the company would now have to address their
entitled worldviews and continue to make changes until a new working
arrangement could be forged. What is clear is that at each stage they had
a selfish view of what they wanted and a less than realistic view of the
obstacles in the way of achieving their goal. If they had used a causal
model, they could have anticipated all of these problems. We believe
decision process theory provides just such a causal model of this
transitional behavior.

19.3 Player Engagement

In addition to the entitled behaviors, such as for the f2WWW model,


Figs. 19.2 and 19.3, there is an equally important and related social
distinction: the coupling (charge) or player engagement to the composite
payoff. We considered the structural aspects in Sec. 18.8 in our
discussion of the active strategy of player aggression, Figs. 18.9 and
18.10. Each player engagement corresponds to an inactive strategy and
represents a conserved charge. Here we consider the product of the
player engagement and the composite payoff: this product and the
entitlement component add to produce the co-moving player payoff,
Eq. (19.1).
There is a player engagement for each player and for each code of
conduct (which we treat as players). The code of conduct, as in the
f2WWW model, Fig. 19.4, may look nothing like the payoffs of the
players or the composite payoff Fig. 19.5. It may also have initially a
large coupling as in Fig. 18.9 or Fig. 18.10 (see Ex. 5). For a different
model, it may have zero coupling as in Fig. 19.6 or Fig. 19.7.
Each player payoff in the centrally co-moving frame is the sum of an
engagement term and an entitlement term, Eq. (19.1). See also, Ex. 6.
Depending on the relative sizes of the two terms, the actual player payoff
Social Distinctions for Decision Processes 771

may favor one, the other or neither of these two terms. If there is no
player engagement then the player payoff is all entitlement. If there is no
entitlement then the player payoff is determined by the player
engagement payoff, e ω . The theory produces this model with two
mechanisms and allows us to study their mutual interaction. The
seasonal payoff and the entitlement payoffs of all the players determine
the composite payoff, Eq. (19.2), such as Fig. 19.5 for the f2WWW
model. We hope to find empirical evidence for this decomposition.

Figure 19.4. Code of conduct entitlement payoff in the f2WWW model


We argued above that we can identify when player engagement is
zero and so identify when the player behaves in the extreme case of
acting entitled. These cases establish that there is an effective dynamic
mechanism of entitlement behavior.
We now want to establish the other case: a player recognizes that
there are payoffs agreed to by all parties and buys into that by aligning
their worldview to that payoff scheme, to the composite payoff (such as
Fig. 19.5). This is not the same as a code of conduct. The player accepts
what appear to be the rules of the game in terms of who wins and who
772 Geometry, Language and Strategy—Vol. 2

loses and what payoffs occur in each case. The strength of compliance is
set by the player engagement. This is the opposite case of the player
paying no attention to what others appear to be accepting as the rules of
the game and acting based on their own sense of entitlement.

Figure 19.5. Composite payoff in the f2WWW model


There are many examples in which player engagement appears to be
the primary dynamic mechanism. We provided an example of why
people queue up for tickets in Sec. 11.2 as an example of a particular
code of conduct.
Here is a slightly different story of queuing that illustrates player
engagement rather than code of conduct. Some years ago I was in
Leningrad, as it was called then as it was part of the Soviet Union.
Trading on the black market was apparently very common. I observed
that no matter the time of day, when walking down main streets I came
across people standing in long lines. I asked someone about them. He
told me the following story. A man saw one of these long lines and
immediately joined it at the end of the line. After standing there for a few
minutes another man did the same thing and then after a few minutes,
asked the first man, “what are we in line for?” The first man answered, “I
Social Distinctions for Decision Processes 773

don’t know, but whatever it is it must be great because the line is really
long.”

Figure 19.6. Phase space plot of player engagements in


the f2AʹD model on initial boundary

We think of player engagement as providing that coupling to the


collective worldview of what constitutes payoffs for the decision. That
collective worldview is likely to differ from the individual’s entitlement
worldview. The more engaged we are, the higher we value the composite
payoff. We have focused in this book on processes in which the player
engagement and composite payoff are harmonic steady state with non-
trivial network connectivity. We implicitly allow transients to die out so
that we have a pure travelling or standing wave. Our solution of the
equations may not correspond to the “story” one normally tells about a
situation.
Thus in the example of the software development project, the initial
composite payoff reflected the successful delivery of past software
releases to the customer and so also reflected the successful sales of the
company’s product to this customer.
It is not clear however that this represents strong player engagement.
Indeed, we think it may represent the opposite. We argued that both the
774 Geometry, Language and Strategy—Vol. 2

customer and the company internalized behaviors so that their composite


payoff view represented how they viewed the transactions. Fortunately
for both sides these two views were the same. The player engagement we
argued was probably zero and their entitled payoffs proportional to the
composite payoff.

Figure 19.7. Phase space plot of player engagements


in the f2AʹD model on z  0.02 boundary

This is supported by the fact that as soon as the situation changed,


neither side anticipated any change to the basic rules and payoffs. Their
player engagements were initially zero. They made their plans based on
their internal entitled view, which ceased to match the composite payoff.
Only after some time did it become apparent that there was a problem.
The player engagement of each was no longer zero and as a consequence
both sides adapted to the change based on their level of player
engagement to the problem.
I came across the concept of player engagement as a dynamic
mechanism in a management seminar that focused on Lincoln [Phillips,
1992, p. 121]. The seminar recounted that during the Civil War, Lincoln
believed that his generals were not sufficiently engaged. In one example,
General McClellan, with a large army, met an advance by Robert E. Lee
Social Distinctions for Decision Processes 775

in 1862. Although Lee was outnumbered two to one, he inflicted heavy


losses on McClellan’s troops after which he then retreated across the
Potomac River pursued by McClellan’s superior numbers. McClellan did
not pursue, claiming victory. However, Lincoln believed that McClellan
had only defended his ground and had failed to engage. Lincoln replaced
him.
Player engagement is important because it provides a mechanism for
the decision maker to react to the other decisions that are being made.
There will always be a composite payoff based on the decisions that have
already been made. To fail to respond to those can at times be disastrous,
possibly to one’s career.

19.4 Player Interest

The player engagement is based on the network gradient of the player


interest, Eq. (14.29). A player may not be initially engaged. For example,
the boundary condition can be that the player engagement is zero, as it is
for Fig. 19.6. In this example, the initial player interest is large, which
generates a player engagement away from the initial boundary. When the
initial player interest is small, little change is seen in the player
engagement, such as Fig. 16.4 for “blue” and “red”.
We believe the player interest is a social distinction. The player
interest for each particular strategy changes a player’s engagement and
generalizes the notion of value from game theory, Sec. 14.9.5. Value, as
usually understood in game theory, is the same regardless of strategy: for
game equivalency, Sec. 15.8, there is the same player interest for any
strategy chosen. For zero sum games, we imagine what is gained by one
player is lost by the other, so the sum of the values is zero.
This does not reflect the general case in decision process theory in
which the player interest may differ from one strategy to the next.
Instead, we have a symmetry that replaces the zero-sum concept that
leaves some aspect of this sharing intact.
The source for player interest is the buy-in expressed by the overlap
between the player’s entitlement payoff and the common composite
payoff, ωBα (Sec. 14.9.5). This source depends on all of the players
776 Geometry, Language and Strategy—Vol. 2

through the composite payoff. There is also a matching feedback loop.


Not only does player interest change a player’s engagement, the product
of the entitlement payoff and player interest I α  B α (Sec. 14.9.8)
contributes the player impact to the composite payoff. Player interest is
in a pivotal position.
The player interest reflects a player’s disposition to a particular
strategy at the moment. For example in our example of a queue (Sec.
11.5.1), a person stands in line because of interest. It is not based on
knowledge of the consequences. Standing in line does not reflect a
decision to purchase something that might be sold to someone who gets
to the front of the queue. It reflects a spatial gradient: an interest to adapt
to social pressure.
Interest often occurs when we are on autopilot. We are content with a
particular set of strategies that have been successful in the past, as in our
software development example (Sec. 18.2). New circumstances may
remind us of companies that have been in similar circumstances. We
become interested in particular strategies as a consequence. Thus our
interest is encouraged or fed by the fact that our decision processes occur
not in a vacuum but in a society of other similar processes. We respond
not only to the causal nature of our own experiences but to nearby
societal or network experiences that are occurring simultaneous with
ours. These are new effects not ordinarily considered in economic or
game theory deliberations. They have been discussed in the literature for
example by Gladwell [2005].
Being on autopilot relative to a particular strategic direction suggests
that along that strategy and across the network, there are no variations of
any quantity we choose to look at so that there is no reason to pick one
region of the network over another. This is the requirement for a code of
conduct. It is variations in at least one quantity that break a code of
conduct. One possibility is that that quantity is the player interest. Thus
the situations in which a contract is not honored (Sec. 11.7, Ex. 1) or a
case in which resources of the commons (Sec. 11.2) are being exploited
or the prisoner’s dilemma paradox (Chap. 7) are all cases that may bring
forth a player interest.
Like the composite payoff, player interest reflects the decisions and
payoffs at that point. The player interest can be characterized as a payoff
Social Distinctions for Decision Processes 777

between the idiosyncratic inactive worldview strategy and an active


strategy. Like payoffs in general, player interest reflects the potential for
change. For this reason it is an important attribute to identify and
quantify. We expect to measure it based on its effects on the player
engagement coupling. It is significant that player interest is associated
with player engagement and with adaptive change to social pressures.
The interest in a strategy is based on the social attraction.
As a social attraction, we expect that player interest to be a source for
change to the player entitlement. One measure, ω  Iα , might be whether
the direction of player interest is parallel to an equilibrium direction of
the composite payoff (Sec. 14.9.6). If not, then the player interest
challenges the entitlement payoffs to change. Such challenges generate
network changes to the entitlement. This is in addition to more selfish
forces that change entitlement.

19.5 Player Passion

Besides being attracted to a particular strategy for a social reason, a


player may be attracted for selfish reasons. An example is a player’s
passion for a particular strategy, which may overpower objective choice
and also determine entitlement. The concept here is analogous to stress in
physics as opposed to strain. It represents a force. We expect a strong
player passion to influence initially the entitlement payoff.
We view player passion as a stress on the system associated with a
player and pointed in a particular strategic direction. This is similar to the
challenge generated by the player interest. The difference is that the
player passion is self-centered and the player interest is socially
centered. Both contribute to the entitlement payoffs and can be different.
Stresses however are not purely private, but can also be social: they
represent constraints not taken into account by the active strategies, Sec.
6.3. They are like electric currents that come into existence because of an
external electric field: we observe the effect though we may not always
see the source. The player passion is generated from a potential field in
our models. See for example Figs. 17.40 and 17.17. The player passion is
computed from the gradient of this potential.
778 Geometry, Language and Strategy—Vol. 2

We think that there are many examples of player passion. A strong


belief in a cause can be characterized as a player passion for a specific
strategy. It makes perfect sense that the entitlement payoffs seen by a
person with such a cause will be determined by that passion. We suggest
that charismatic leaders generate such player passion.
Though we expect player passion to significantly impact the
entitlement payoff, in general it will not impact the composite payoff
unless there is a strong player engagement. We suggested this earlier by
noting our expectation that impact on composite payoff would be made
when the product of the player interest and entitlement payoff was
significant. We think an equivalent measure is the product of the player
engagement and player passion (Sec. 14.9.11), which will impact the
seasonal vector potential and indirectly the seasonal payoff; this will
impact the composite payoff. A strong impact is made if a player is both
passionate and engaged; in other words if the player pays attention both
to those things they care most about and at the same time pay attention to
those decisions that have the most social interest.

19.6 Mutual Player Support

The idea that players may interact with each other outside the confines of
the competitive decision process was envisioned by Von Neumann &
Morgenstern [1944]. However, parts of that proposal were flawed (Cf.
the discussion in Sec. 12.4). We believe we have a replacement of that
concept without the flaws that were identified. We replace their proposal
with the idea of cooperation potentials, Sec. 12.4, which we refine here
as mutual player support potentials. These are social constructs within
decision process theory.
Though the mutual player support potentials are specific to decision
process theory, we see some similarity between them and the
imputations and side payments introduced by Von Neumann &
Morgenstern [1944]. They are payments outside the framework of the
competitive valuations of the forces as measured by the payoffs. The
difference is that in decision process theory, these side payments are not
outside of the theory, which as constructed is a self-consistent
Social Distinctions for Decision Processes 779

mathematical framework. The side payments reflect what can be viewed


as psychological inputs to the decision process.
We recall some aspects of decision process theory that are relevant.
We have three components of the active geometry acceleration Q in
Eq. (18.1) that may force strategic behavior: an absolute force q , a
competitive valuation force I e and a cooperative support force
 e e  (Cf. Sec. 12.4). Game theory focuses on the competitive forces
and considers behaviors that are in equilibrium under them. However, we
think that decisions require knowing all three components for a given
choice. Decisions require knowing that a given set of choices has mutual
player support as measured by the cooperative support force above. We
suggested that a measure of the mutual support potential is the co-
moving frame potential  . This is a quantitative test for whether
competitive forces are sufficient. If the mutual support potential is zero
between two players then there are no cooperative support forces
accelerating the strategy choices.
In any frame, the mutual support player potentials are necessary
attributes of the theory that follow directly from our formal
characterization of active and inactive strategies. In a decision process,
they reflect the fact that players may change their payoff strategies as a
consequence of the support potentials, which act as mixing terms. For
example, the mixing changes the payoffs and the passions for one player
as a function of payoffs and passions of another player by Eq. (14.24)
and Eq. (14.25) respectively. Similar results hold for player engagement,
Eq. (14.29), and interest, Eq. (14.23).
We compute the attributes of the mutual support potentials in the co-
moving orthonormal frame, Secs. 14.9.3-14.9.4. In any frame, the mutual
support player potentials are determined from the shear components by
the inactive frame transformations, Sec. 14.9.10. Thus the player
engagements, strains and bonding determine the mutual player support
potentials.
These are special cases of the general results for the fixed frame
model solutions that are summarized in Tables 5.1, 5.4, 5.5 and 5.6 in
Chap. 5. If there is no mixing, then each player makes choices that
depend only on information available to all, such as the acceleration: the
choices are idiosyncratic. Mixing is the essence of cooperation in that it
780 Geometry, Language and Strategy—Vol. 2

is not idiosyncratic; it requires the players to make an agreement. The


agreements occur between each pair of players. The mutual player
support potentials determine the fractional change to what would
otherwise be idiosyncratic attributes.
It therefore becomes important to identify mutual player support in
practical situations in order to set the amount of mixing that occurs at an
initial point. In general, we expect the mutual support potential to be
non-zero. In particular there are player self-support forces possible,
which will be non-zero only when the player engagement is non-zero. At
the still point these forces may vanish, but not otherwise.
When these forces are non-zero, they may effectively balance the
absolute forces. In particular we think of the sum of the self-support
forces 1 ni  h e e  as playing this role. The sum contains the self-
support for each player. The remaining contribution to the cooperative
support force is the shear term,   e e  . The mutual support
potentials are therefore the player bonding potential,  and the player
shear potential   .
For the fixed frame model solutions, the shear tensor is a diagonal
matrix or a constant rotation of a diagonal matrix. If diagonal, then only
self-support terms exist. If there is a constant rotation of a diagonal
matrix, then by a suitable redefinition of player, we have only self-
support contributions from characteristic players. However in the latter
case, from the “real” player’s perspective, there will be cooperative
forces. This process was carried out for the models we have employed
and we do find cooperative effects. More generally however, we argue
that decision process theory need not make the fixed frame model
approximations and that for more general models we will get richer
cooperative behaviors. However, we think it significant that cooperative
behavior is already a feature of the solution even in our fixed frame
model approximation.
Note that the mutual player support distinction is different from the
code of conduct distinction. Both have the same origin: the existence of
an isometry or inactive strategy. However, players are not making a pact
about adhering to a particular set of strategies. The mutual player support
potentials are payoffs that drive actions by changing payoffs and
passions. For example, in the f2WWW model Sec. 17.8, the payoffs of
Social Distinctions for Decision Processes 781

the original model determined the mutual player support potentials


between the code of conduct player and each of the players. The
corresponding potentials for the shear are in Ex. 7. Note that each player
can in fact determine the support potential between himself and the code
of conduct player.
We also note that the mutual player support is determined by the
player bonding. There is a force due to the matrix of potentials whose
surface normal vectors describe the strength of bonding between
individuals. Such bonds provide gradient forces dictating how players
respond to cooperative forces. Gravity and centripetal acceleration are
both examples from physics of bond forces.

19.7 Hidden-in-Plain-Sight

Ordinarily, we smooth out time fluctuations such as business cycles,


seasonal cycles, etc. We quote company profits by month, taking out the
fluctuations that we know are there because of seasonal buying. In this
way we claim to get an accurate picture. Such an approach is effective
with statistical models, but not in causal models.
In a causal model, such cyclic time dependent effects create
qualitatively new phenomenal, Sec. 18.3. For example, the rotation of the
earth is a cyclical effect, which creates weather patterns as a consequence
of Coriolis forces. In decision process theory, we also expect that cycles
are harmonic steady state waves that can’t be factored out and must be
incorporated systematically into the solution. An important subclass of
such cycles consists of those that are hidden-in-plain-sight, cycles we
don’t take into account yet ones that play an important dynamic role.
These are cycles that are actually not seen by us, or if seen, are not felt to
be of significance and ignored.
For example, in our daily life, there is one obvious hidden-in-plain-
sight cycle: the “daily” part of “daily life”. We make decisions each day
assuming that the “daily” part of that cycle plays no critical role. In a
public corporation, there is a quarterly cycle that is the hidden-in-plain-
sight cycle. Each quarter, profits or losses must be announced. This
drives behavior and can be taken into account in decision process theory.
782 Geometry, Language and Strategy—Vol. 2

These hidden-in-plain-sight cycles are not always taken into account;


they are part of the background noise not because they are small or
unimportant, but because they are so pervasive they are not noticed.
The earth’s rotation is a hidden-in-plain-sight cycle. Sunrise is in
plain sight every day yet the relationship between sunrise and the earth’s
motion is hidden. Once the relationship has been pointed out, it becomes
totally obvious. Even though obvious, its role in weather may not be
noticed: for example it causes breezes near large bodies of water to flow
towards land or away from land based on the relative heating of the
water and land, an effect which changes on a daily cycle.
We expect that the addition of hidden-in-plain-sight cycles may
amplify effects that are already present, Sec. 18.3. We do expect such
effects to be present in the real world. Consider the effect of an
advertising campaign. We run an ad daily for a particular product. The
target audience for that ad listens to the ad daily and also makes buying
decisions. We suggest that the effect of the ad campaign is subtle. The
target audience has items it might buy and will do so according to
preferences. The ads however may impact the timing of the buying,
which may change the observed payoffs. The ads also skew the actual
buying power. The ads are an external harmonic force that modifies
behavior to create a new steady-state behavior.
In special cases, we suggest that such small skewing behaviors
generate effects analogous to tornadoes and hurricanes if left in place for
long periods of time, analogous to the earth’s rotational role in
generating prevailing winds and storms. We suggest the need for
different ways of looking at decisions to take into account such
behaviors. In Exs. 8-9, we suggest the notion of decision isobars,
analogous to weather isobars, as one new way of looking at decision
processes. The exercises illustrate the possibility for additional hidden
cycles.

19.8 Outcomes

Based on this chapter, the student should have learned a number of social
distinctions that might arise as we move from local behaviors to global
Social Distinctions for Decision Processes 783

behaviors. The student should be able to identify these distinctions in real


world decision processes and qualitatively see how such distinctions
would apply in the theory.
The attainment of the outcomes of this chapter is facilitated by doing
the exercises in the following section. Though not essential for a general
understanding, there is benefit to replicate the numerical examples given.
Based on this investment, the student should achieve the more detailed
outcomes below. Furthermore, though again not required, it is suggested
that the reader carry out as many calculations as they feel comfortable
with to build up an intuition of the interactivity of the many effects
incorporated into the theory; as with engineering disciplines, knowledge
is gained by doing not just by reading or listening.
In addition to these general considerations, the student should have
learned the following outcomes as social distinctions:
 Code of conduct: in Sec. 19.1, the student will have learned that
certain behaviors preserve and conserve cooperative behaviors.
 Entitlement: in Sec. 19.2, the student will have learned the idea of
entitlement: it is a player acting alone, empowered without regard to
other player behaviors; egoistical behavior. This is the first
component that governs a player’s choice.
 Player engagement: in Sec. 19.3, the student will have learned a
second component of payoffs, which is based on the coupling or
player engagement to the composite payoff. This is a new dynamic
distinction. It is the product of the player engagement and the
composite payoff and is a component that governs a player’s choice.
 Player interest: in Sec. 19.4, the student will have learned what
generates a player’s interest. The player interest for each particular
strategy changes a player’s engagement and generalizes the notion of
value from game theory.
 Player passion: in Sec. 19.5, the student will have learned that there
may be stresses outside of game equivalency that may drive the player
payoffs. A player’s passion for a particular strategy may overpower
objective choice and determine entitlement. Though idiosyncratic, in
nature, player passion may arise from social stresses.
 Mutual player support: in Sec. 19.6, the student will have learned
that decision processes are governed in part by player cooperation
784 Geometry, Language and Strategy—Vol. 2

driven by mutual player validation forces. These play as important a


role in governing behaviors as do competitive valuation forces. This
depends in part on a matrix of potentials whose surface normal
vectors describe the strength of bonding between individuals. Such
bonds provide gradient forces dictating how players respond to
cooperative forces. Gravity and centripetal acceleration are both
examples from physics of bond forces.
 Hidden-in-plain-sight: in Sec. 19.7, the student will have learned that
some harmonic steady state waves may be aspects of cycles that are
hidden-in-plain-sight, cycles that need to be incorporated
systematically into the solution. Until now, such effects are unseen or
ignored; they are distractions. They may however be important as part
of the bigger picture.

19.9 Exercises

(1) Discuss the differences between the code of conduct charge for the
f2WWW model, Fig. 19.1 and the corresponding plot for the f2AʹD
model, Fig. 19.8 in terms of the social aspects of each model. Note
the former is designed as a social model and the latter an attack-
defense model.

Figure 19.8. Code of conduct charge for f2AʹD


model
Social Distinctions for Decision Processes 785

Figure 19.9. Code of conduct charge for f2WWW


model with extended time interval

Figure 19.10. Entitled payoff for “blue” in the f2A´D


model

(2) Discuss the differences between the code of conduct charge for the
f2WWW model, Fig. 19.9 and the corresponding plot for the f2AʹD
model, Fig. 19.8 in terms of the social aspects of each model. In this
case the f2WWW model plot is for an extended time scale. How is
this different from Ex. 1?
786 Geometry, Language and Strategy—Vol. 2

Figure 19.11. Entitled payoff for “red” in the f2A´D


model

(3) For the f2AʹD model, for the entitled payoffs for “blue” and “red”,
we have Figs. 19.10 and 19.11. Compare this with the f2WWW
model, Figs. 19.2 and 19.3. Note that in both cases, the behaviors
follow the transverse boundaries for points sufficiently near.
Discuss the implications of these behaviors for the two models.

Figure 19.12. Entitled payoff for code of conduct in the


f2A´D model
Social Distinctions for Decision Processes 787

Figure 19.13. Composite payoff in the f2A´D model

(4) Compare the f2AʹD model for the entitled code of conduct payoff,
Fig. 19.12 with the corresponding plot Fig. 19.4 for the f2WWW
model. In this case, how do the boundary conditions influence the
behaviors? Include in the definition of the boundary, the behavior
on the surface z  0 . Similarly compare the composite payoff
Figs. 19.13 with 19.5.

Figure 19.14. Charge contours z  0.2 for the


f2WWW model
788 Geometry, Language and Strategy—Vol. 2

Figure 19.15. Charge contours z  0.2 for


the f2WWW model

(5) Charge contours for the f2WWW model are shown in Figs. 19.14
and 19.15. Compare and contrast this with the corresponding plot
for the f2AD model, Figs. 18.9 and 18.10.

Figure 19.16. Engagement payoff for “work” in the


f2WWW model
Social Distinctions for Decision Processes 789

Figure 19.17. Engagement payoff for “wealth” in the


f2WWW model

Figure 19.18. Engagement payoff for code of conduct in


the f2WWW model

(6) The player engagement payoffs, e ω , for the f2WWW model are
plotted in Figs. 19.16-19.18. Compare these payoffs with the player
entitlement payoffs, Figs. 19.2-19.4. Based on the sizes as provided
in the legends, describe the resultant player payoff in the central
co-moving frame, Eq. (15.8).
790 Geometry, Language and Strategy—Vol. 2

(7) The player shear gradients   , Figs. 17.28-17.32 are each


derived from a corresponding shear potential   , Figs. 19.19-
19.23. Based on the potentials, what would you expect the gradients
to be and does that agree with the numerical calculations provided?
(8) Hidden-in-plain-sight cycles: In meteorology, hidden cycles are
made visible by “abstract” contour plots. One such popular plot
consists of pressure contours or isobar plots. They help vision the
behavior of the weather which would otherwise be hidden
phenomena, based on cycles that are ordinarily ignored such as the
rotation of the earth. For example, consider the f2AD model for
pressure, Fig. 17.25. This is a density plot, which can also be
viewed as a plot of strategy contours. The situation is more
complex than the usual meteorological plots since we want to see
the full 3D behavior. Which would be better and why: a contour plot
or a density plot as the one that is shown? If we provide contours,
these then would be decision isobars.

Figure 19.19. Player shear  11 in f2WWW model


Social Distinctions for Decision Processes 791

Figure 19.20. Player shear  22 in f2WWW model

Figure 19.21. Player shear  12 in f2WWW model


792 Geometry, Language and Strategy—Vol. 2

Figure 19.22. Player shear  13 in f2WWW model

Figure 19.23. Player shear  23 in f2WWW model

(9) We continue the inquiry of Ex. 8 for additional hidden-in-plain-


sight cycles. We have considered solutions to the player fixed frame
model in which certain co-moving frame coordinates y are cyclic
and exact. One consequence of this is that in the normal-form
coordinate basis, the coordinate behaviors are constrained and cyclic
Social Distinctions for Decision Processes 793

(e.g. Fig. 16.15). We can investigate this behavior in more detail by


considering the surfaces in which one of these coordinates y is
constant. This will be a 3-surface in the 4-dimensional space-time of
the model. If we choose two such coordinates to be constant, we
will get a 2-surface, which is something that we can plot and hence
visualize. What would you learn from such a plot? As an example
consider and discuss Fig. 16.26. Is this an example of a hidden-in-
plain-sight cycle?
(10) The code of conduct distinction results from a group theoretic
notion of local symmetry. As articulated in Sec. 4.5, the isometries
reflect properties of the internal local group, which we take to have
the form U1    U1  G , in which the players form a mutually
commuting group formed from the one-dimensional group U1 . The
codes of conduct are represented by the remaining group G .
Certainly a unique aspect of our approach is that the local group
structure also generates global structures (Cf. Chap. 13) that reflect
the internal structures.

So far, we have assumed that this internal symmetry group G has


the same structure as the players; we have imagined that there are
fictional players representing each of the codes of conduct. This
assumption is not essential; it was not assumed in Vol. 1. The code
of conduct group G could in fact be more complicated. This
complication requires two or more codes of conduct strategies, so
was not possible in our numerical examples with three active
dimensions. An interesting example might be a three person
decision process.

With these considerations in mind, discuss what form the symmetry


group G might take and how behaviors would be changed. It might
be worth noting that physical bodies such as planets pick a spherical
shape (the internal symmetry group). Such shapes are created during
planet formation where viscosity might play a role. We have
speculated that viscosity might also play a role in decisions (Sec.
6.3). Might there be a connection?
794 Geometry, Language and Strategy—Vol. 2

(11) Even when the player engagement is small, there is nevertheless a


current (player passion) associated with it. We have argued that the
current generates the player’s payoff field in analogy to the way in
which an electrical current generates the magnetic field (Ampere’s
law, Sec. 2.4). How might this effect be measured? Note there may
be a relationship between the anxieties one experiences when
“things are not going right”. We could understand the concept of
“going right” as related to the internal view of the payoffs, and the
anxiety as resulting from the real world observation of the way
things are actually going. Discuss this also for the code of conduct
player.
(12) The opposite effect to Ex. 11 is that changes in the player payoff
over time generate a player bias (Faraday’s law, Sec. 2.3). How
might this be measured for players and for code of conduct players?
Note here that there might be an opposite feeling, one of
exhilaration to see how something could be approached differently.
This might be an indication that the player’s world map (payoff) is
changing. The world map also describes a flow of payoff flux lines.
This flow can be stationary or change in time. The exhilaration
might indicate that the payoff flux is changing in time, setting off a
strong player bias that creates a current (player passion) around this
flux direction.
(13) One way to identify codes of conduct is by examining the language
used to describe the decision process. For example, we talk about
the “rules of the game”, which suggests that there is a fictional
player representing what it means to be a “game”. Cultures have
also identified names for forces that were not understood at the
time, such as “wind”, “fire” and “water”. These concepts were
assigned personalities which acted as if they were indeed real
players. What examples can you come up with to support or
disprove this approach?
Chapter 20

Afterword

This is phase II of a project that grew out of a desire to create a new way
of talking about decision processes. In phase I, Vol. 1, I provided an
introduction to the ideas and qualitative reasons and simplified
calculations for seeing how one might address decision processes using
the language of physics and differential geometry. The idea was to use a
language that was created to describe and understand processes that
evolve in time according to specified dynamic mechanisms. In the first
phase, no attempt was made to provide a comprehensive solution to the
equations, nor was there an attempt to educate the readers to the level
where they could solve the equations themselves.
This is phase II of the project. I have identified the types of solutions
found in phase I as the analog of AC resonant solutions; game theory is
analogous to DC circuits. In addition to extending the coverage to AC
steady state solutions, in this volume, I attempt to make this
identification mathematically sound. The goal has been to provide the
necessary background to bring anyone with basic mathematics and
engineering foundations up to the level of a new way of talking about
decision processes so that they can apply the theory to real world
problems. The goal is that the reader should gain sufficient expertise to
understand how the ideas apply. That has entailed providing essential
background for the mathematics, the physics and the economic theory.
The next phase is the interesting challenge to apply the tools from
phase II and the ideas from phase I to practical decision processes. It is
also the opportunity to broaden the models considered and investigate
transient solutions more thoroughly. I envision that the way of looking at
decisions that has been elaborated here will involve looking at empirical

795
796 Geometry, Language and Strategy—Vol. 2

data in ways that will generate significant insights. It may also require
new ways to look at data. I intend to chart progress for this phase on the
following website: http://decisionprocesstheory.com. This site will also
have helpful information on the first two phases of work, including
Mathematica notebooks that are not provided in this volume.
It is of course difficult to predict how future studies may progress.
Our current view is as follows. So far, we have used the player fixed
frame model as our guide. We believe our next guide will be the player
ownership model, which extends what we have learned with an energy
momentum tensor that reflects the holonomic constraints.
(1) Pick a problem for analysis and start with a game theoretic view.
(2) Assume there is a stationary holonomic frame in which time is
inactive and mutually commutes with the player inactive strategies,
including any code of conduct strategies. Identify the players,
including the codes of conduct players.
(3) Choose the independent metric components to be the gauge

invariant quantities  jk g
ab

F j ab , (which in a centrally
co-moving frame determine player engagements V j and seasonal
flows V ). Choose the gauge independent holonomic variables to
be time and the active strategies.
(4) Transform to the frame in which the flows V  ,V j and charges
V ,V j are zero.
(5) Set the values of the energy momentum tensor using the player
ownership structures in this frame (Sec. 6.10, Ex. 40):

 j j, k  o  J     ,   o  J 
J jk   k , J       ,
 0 otherwise  0 otherwise
 J jk 0 0 
 
QJ    0 0 0 . (20.1)
 0 0 J   

(6) Transform the ownership structures back to the gauge independent


holonomic frame noting that ownership applies only to the active
and inactive strategies, not to time (where V o Vo  1) :
Afterword 797

J i  l   j  V V j   J jl ,

J  a      V V   J   ,

J  

 
  V  j  V  V j  J j kV k    V V  J  V   , 
 J i  j J   
j
J a  
j

 k

QJ    J i  k J   J a   .
   
 
(20.2)
  
 J i  k J  

J a  
  

(7) Impose the constraint of ownership through the energy momentum


tensor, not unlike constraints for immiscible fluids (Cf. Sec. 3.11,
Ex. 17 and Eq. (3.37)). Use the symmetrized form that implements
the idea of player from Vol. 1, Sec. 7.3:

p   J Q  J  g  Q  J    J Q  J     Q  J   . (20.3)
   
 
J J

(8) Set the gauge; e.g. use the seasonal gauge, Sec. 6, Eq. (6.98).
(9) Solve the resultant elliptic partial differential field equations in the
central stationary frame. The equations are now all elliptic since
they are independent of time. Use a general purpose solver to focus
on solutions rather than techniques.
(10) Match the boundary condition of the problem under investigation
using the harmonic wave equation.
(11) Compare the results of the simulation against observed behaviors to
establish the relevance of the model.

The context for decision process theory solutions is that individuals


control the possible outcomes of any decision process. There is an aspect
of decisions that can be predicted. We shape the future through the
normal cause and effect of any physical process. This is based on the
observational experience that there are causes that lead to effects and
these relationships follow knowable laws.
One way to understand that is to recognize that our behaviors are both
creative and restrictive. There are rules that constrain our behaviors. Part
of those constraints limits what we see as possible. In decision process
798 Geometry, Language and Strategy—Vol. 2

theory, these constraints are the idiosyncratic payoffs associated with


each person about how they see the world and how they think others see
the world. In addition, there are cooperative views about how pairs of us
think we can work together. There are constraints due to ownership as
described above.
The future behaviors depend in part on the snapshots of how we
currently see the world. Our current view is the world that occurs “now”.
This view becomes part of the cause and effect of how the future will
evolve; they are not the sole determinants of the future but are important
contributors. The collection of all of these views, as well as the effects of
the physical world, then determines what happens. The future does not
spring fully formed out of nothing, but has its origins in the past. We
suggest this reflects the reality of how past experiences impact future
behaviors. Certainly one high level test is to find evidence that
contradicts this hypothesis.
In adopting this hypothesis, we have access to creating the future by
directly working on changing our idiosyncratic views and our
cooperative views of the past. It is not changing the past but more
accurately articulating the past into our views. As we connect more
strongly to the way things are we allow more possibilities for the future.
If we don’t our behaviors may in fact be overly and unnecessarily
restrictive. For example, we may add attributes to the “now” based on
belief structures that don’t represent what is happening. In the future, this
may lead to actions that fail to play out as we hope.
We come to the view that we create the future not by causing things
to happen as prediction, but we create the future by allowing things to
happen as cause and effect based on a more realistic assessment of what
is possible. We base our view of what is possible not only by our learned
belief systems but by a serious inquiry into what happened in the past
and what is happening now, removing as much of our personal beliefs as
possible that are in conflict with observation. This leaves us free to
create the future based on our desires that are more achievable. In this
way we create who we are and to some extent change those around us.
What can now happen, will reflect these new viewpoints.
Bibliography

Aumann, R. J. (1989). Game Theory. In J. Eatwell, M. Milgate, & P.


Newman (Eds.), The New Palgrave: Game Theory (pp. 1-53).
New York: W W Norton.
Barabási, A.-L. (2003). Linked. NY: Plume, a member of the Penguin
Group.
Bayes, T. (1764). An essay in solving a problem in the doctrine of
chances. Philosophical Transactions of the Royal Society of
London, 53, 370-418.
Bentham, J. (1829). Fragm. on Govt. Wks., X, p. 142.
Bernoulli, C. (1738). Exposition of a new theory of the measurement of
risk. Econometrika (English translation 1954), 23-36.
Bhatti, M. A. (2005). Fundamental Finite Element Analysis and
Applications: With Mathematica and Matlab Computations.
John Wiley and Sons, Inc.
Borel, E. (1921). La Théorie du Jeu et les Ěquations Intégrale a Noyau
Symétrique. Compte Rendus de l'Académie des Sciences, 173,
1304-1308.
Bourbaki, N. (1968). Elements of Mathematics. Paris: Herman,
Publishers in Arts and Sciences (translated English edition
Addison-Wesley Publishing Company).
Cameron, G. (2008). Oikos and economy. PhaenEx, 3(1), 112-133.
Chandrasekhar, S. (1961). Hydrodynamic and hydromagnetic stability.
New York: Dover Publications.

799
800 Geometry, Language and Strategy—Vol. 2

Clemhout, S., & Wan, J. H. (1989). Differential Games. In J. Eatwell, M.


Milgate, & P. Newman (Eds.), Game Theory: The New Palgrave
(pp. 129-132). New York: W. W. Norton.
Condorcet. (1793). Plan de Constitution, presenté a la convention.
Oeuvres, 12, 333-415.
Courant, R., & Hilbert, D. (1962). Methods of Mathematical Physics
(Wiley Classics Edition ed., Vol. II). New York: John Wiley &
Sons.
Crawford, W. P. (1978). Mariner's Weather. New York: W. W. Norton
& Co., Inc.
Dali, S. (2015). Wolfram|Alpha. Retrieved November 24, 2015, from
Wolfram Alpha: http://www.wolframalpha.com/
Dresher, M. (1981). The Mathematics of Games of Strategy: Theory and
Applications. New York: Dover.
Dufwenberg, M., & Kirchsteiger, G. (1998). A theory of sequential
reciprocity. Tilberg Center for Economic Research, (discussion
paper 9837).
Edgeworth, F. Y. (1881). Mathematical Psychics: An Essay on the
Application of Mathematics to the Moral Sciences. London,
England: C. Kegan Paul & Co.
Eilenberg, S., & Steenrod, N. (1952). Foundations of Algebraic
Topology. Princeton: Princeton University Press.
Eshel, I., Samuelson, L., & Shaked, A. (1998, March). Altruists, egoists,
and hooligans in a local interaction model. The American
Economic Review, 157-179.
Feynman, R. P., & Hibbs, A. R. (1965). Quantum Mechanics and Path
Integrals. New York, McGraw-Hill.
Feynman, R. P., Leighton, R. B., & Sands, M. (1963). The Feynman
Lectures on Physics. Reading: Addison-Wesley Publishing
Company, Inc.
Fisher, R. A. (1925). Statistical Methods for Research Workers. New
York: Hafner Publishing Company.
Forrester, J. W. (1961). Industrial Dynamics. Waltham, MA: Pegasus
Communications.
Bibliography 801

Friedman, R. M. (1989). Appropriating the Weather, Vilhelm Bjerknes


and the Construction of Modern Meteorology. Ithaca: Cornell
University Press.
Gardenfors, P., & Sahlin, N.-E. (Eds.). (1988). Decision, Probability,
and Utility: Selected Readings. Cambridge: Cambridge
University Press.
Gladwell, M. (2005). Blink: The Power of Thinking without Thinking.
New York: Little, Brown and Company.
Gladwell, M. (2008). Outliers: The Story of Success. New York: Little,
Brown and Company.
Gockeler, M., & Schucker, T. (1987). Differential Geometry, Gauge
Theories, and Gravity. New York: Cambridge University Press.
Goldstein, R. (1959). Classical Mechanics. Reading: Addison-Wesley
Publishing Company, Inc.
Hansson, S. O. (1994, 8). Decision theory: a brief introduction, written
for the participants of a course on risk analysis at Uppsala
University in 1994. Uppsala, Sweden. Retrieved 4 12, 2010
Harsanyi, J. C. (1967-1968). Games with incomplete information played
by "Bayesian" players, I-III. Management Science, 14(3,5, 7),
159-182, 320-344, and 486-502 respectively.
Harsanyi, J. C. (1989). Bargaining. In J. Eatwell, M. Milgate, & P.
Newman (Eds.), Game Theory: The New Palgrave (pp. 54-67).
New York: W. W. Norton.
Hawking, S. W., & Ellis, G. F. (1973). The Large Scale Structure of
space-time. Cambridge: Cambridge University Press.
Howard, R. A. (1964). Dynamic Programming and Markov Processes.
MIT Press.
Howard, R. A. (2010). http://en.wikipedia.org/wiki/Ronald_A._Howard.
Retrieved May 10, 2010, from Wikipedia.
Jackson, J. D. (1963). Classical Electrodynamics. New York: John Wiley
& Sons, Inc.
Joyce, H. (2001, March). Adam Smith and the invisible hand. +Plus
magazine. Cambridge, UK.
Kaluza, T. (1921). On the problem of unity in physics. Sitz. Preuss.
Akad. Wiss., K1, 966.
802 Geometry, Language and Strategy—Vol. 2

Klein, O. (1956). Generalization of Einstein's theory of gravitation


considered from the point of view of quantum field theory. Helv.
Phys. Acta. Suppl., IV, 58.
Laves, T. (1994, February 14). Conversation on coalitions. (G. H.
Thomas, Interviewer)
Lucas, W. F. (1969). The proof that a game may not have a solution.
Transactions of the American Mathematical Socety, 137, 219-29.
Lucas, W. F., & Rabie, M. (1982). Games with no solutions and empty
core. Mathematics of Operations Research, 7, 491-500.
Luce, R. D., & Raiffa, H. (1957). Games and Decisions. New York:
Dover Publications, Inc.
Markus, H. R., & Kitayama, S. (1991). Culture and the self: Implications
for cognition, emotion and motivation. Psychological Review,
98(2), 224-253.
Mas-Colell, A., Whinston, M. D., & Green, J. R. (1995). Microeconomic
Theory. New York: Oxford University Press.
Meadows, D. H. (2008). Thinking in Systems-A Primer. (D. Wright, Ed.)
White River Junction, Vermont: Chelsea Green Publishing.
Meadows, D. H., Meadows, D. L., Randers, J., & Behrens, W. W.
(1972). The Limits to Growth (The Club of Rome). New York:
Signet Book.
Mill, J. S. (1848). Principles of Political Economy with Some of Their
Applications to Social Economy. (W. J. Ashley, Ed.) London:
Longmans, Green and Company.
Mill, J. S. (1947). On Liberty. (A. Castell, Ed.) New York: Appleton-
Century-Crofts, Inc.
Murphy, R. E. (1965). Adaptive Processes in Economic Systems. New
York: Academic Press.
Myerson, R. B. (1991). Game Theory: Analysis of Conflict. Cambridge,
MA: Harvard University Press.
Nash, J. (1951). Non-cooperative games. The Annals of Mathematics,
54(2), 286-295.
Ordeshook, P. C. (1986). Game Theory and Political Theory: An
Introduction. Cambridge, UK: Cambridge University Press.
Osborne, M. J., & Rubinstein, A. (1994). A Course in Game Theory.
Cambridge: MIT Press.
Bibliography 803

Pascal, B. (1670). Pensee. (W. F. Trotter, Trans.) Oregon State


University.
Phillips, D. T. (1992). Lincoln on Leadership. New York: Warner Books.
Piketty, T. (2014). CAPITAL in the Twenty-First Century. (A.
Goldhammer, Trans.) Cambridge, Massachusetts London,
England: Belknap Press of Harvard University Press.
Rabin, M. (1993). Incorporating fairness into game theory and
economics. American Economic Review, 83, 1281-1302.
Ramsey, F. P. (1964). Truth and probability, 1926. In J. H. Kyberg, & H.
E. Smokler (Eds.), Studies in Subjective Probability (pp. 62-92).
New York: Wiley.
Rapoport, A. (1989). Prisoner's dilemma. In J. Eatwell, M. Milgate, & P.
Newman (Eds.), The New Palgrave: Game Theory (pp. 199-
204). New York.
Richmond, B. (2001). An introduction to systems thinking. Hanover,
NH, USA: iThink Software from High Performance Systems,
Inc.
Ryder, L. (2009). Introduction to General Relativity. Cambridge:
Cambridge University Press.
Sabelli, H. (2005). Bios: A Study of Creation. Singapore: World
Scientific.
Sally, D. (1995). Conversation and cooperation in social dilemmas: a
meta-analysis of experiments from 1958-1992. Rationality and
Society, 7(1), 58-92.
Senge, P. M. (1990). The Fifth Discipline: The Art and Practice of the
Learning Organization. New York: Currency Doubleday.
Shubik, M. (1991). Game Theory in the Social Sciences: Concepts and
Solutions. Cambridge: MIT Press.
Slotline, J.-J. E., & Li, W. (1991). Applied Non-Linear Control.
Englewood Cliffs, NJ: Prentice Hall.
Smith, A. (1776). An Inquiry into the Nature of and Causes of The
Wealth of Nations. (1 ed.) London: W. Strahan.
Sofroniou, M., & Knapp, R. (2008). Advanced Numerical Differential
Solving in Mathematica. Retrieved from www.wolfram.com.
Stanovich, K. (2004). How to Think Straight About Psychology (Seventh
ed.). Boston, MA: Allyn & Bacon Pearson Education, Inc.
804 Geometry, Language and Strategy—Vol. 2

Steenrod, N. (1951). The Topology of Fibre Bundles. Princeton:


Princeton University Press.
Stiglitz, J. E. (2012). The Price of Inequality. New York: W. W. Norton
and Company.
Stock Market. (2015). Wolfram|Alpha. Retrieved December 15, 2015,
from Wolfram Alpha: http://www.wolframalpha.com/
Tavani, H. T. (2011). Ethics and Technology: Ethical Issues in an Age of
Information and Communications (3rd ed.). John Wiley & Sons.
Thom, R. (1975). Structural Stability and Morphogenesis. (D. H. Fowler,
Trans.) Reading: The Benjamin/Cummings Publishing
Company, Inc.
Thomas, G. H. (1980). Introductory lectures on fibre bundles and
topology for physicists. Rivista del Nuovo Cimento, 3(4), 1-119.
Thomas, G. H. (2006). Geometry, Language and Strategy. New Jersey:
World Scientific.
Thomas, G. H. (2013, June). Decision Making, Chaos and Determinism.
Retrieved July 10, 2015, from www.decisionprocesstheory.com:
http://decisionprocesstheory.com/white-papers/the-dynamics-of-
decision-processes/decision-making-chaos-and-determinism/
Thomas, G. H., & Kane, K. (2008). Physical Decision Theory, the
Prisoner's Dilemma and a New Foundation for Nash
Equilibrium. unpublished.
Thomas, G. H., & Kane, K. (2010). A dynamic theory of strategic
decision making applied to the prisoner's dilemma. In A. A.
Minai, D. Braha, & B.-Y. Yaneer (Ed.), Unifying themes in
complex systems: Vol VI, Proceedings of the Sixth International
Conference on Complex Systems (pp. 275-282). Springer Verlag.
Thomas, G. H., Sabelli, H., Kauffman, L. H., & Kovacevic, L. (2010).
Biotic patterns in the Schrödinger Equation and the evolution of
the cosmos. In A. A. Minai, D. Braha, & B.-Y. Yaneer (Ed.),
Unifying themes in complex systems: Vol. VI, Proceedings of the
Sixth International Conference on Complex Systems. Springer
Verlag.
Tolman, R. C. (1987). Relativity, Thermodynmaics and Cosmology. New
York: Dover Publications, Inc.
Tzu, S. (1988). The Art of War. (T. Cleary, Trans.) Boston: Shambala.
Bibliography 805

Ury, W. (1993). Getting Past No: Negotiating Your Way from


Confrontation to Cooperation (revised ed.). Bantam Books.
Von Neumann, J. (1928). Zur Theorie der Gesellschaftspiele.
Mathematische Annalen, 100, 295-320.
Von Neumann, J., & Morgenstern, O. (1944). Theory of Games and
Economic Behavior. New York: John Wiley & Sons, Inc.
Wald, R. M. (1984). General Relativity. Chicago: The University of
Chicago Press.
Walton, M. (1986). The Deming Management Method. New York:
Putnum Publishing Group.
Warner, F. W. (1971). Foundations of Differentiable Manifolds and Lie
Groups. Glenview, IL: Scott Foresman and Company.
Wikipedia. (2012). Boom and Bust. Retrieved from
http://en.wikipedia.org/wiki/Boom_and_bust
Wikipedia. (2015). Distributed Element Model. Retrieved November 24,
2015, from https://en.wikipedia.org/wiki/Distributed_element_
model
Wikipedia. (2015). Lumped Element Model. Retrieved November 24,
2015, from Lumped Element Model: https://en.wikipedia.org/
wiki/Lumped_element_model
Wikipedia. (2015). Mobility Analogy. Retrieved November 24, 2015,
from Wikipedia: https://en.wikipedia.org/wiki/Mobility_analogy
#Inertance
Wikipedia. (2015, October 15). Seven Generation Sustainability.
Retrieved December 17, 2015, from Wikipedia: https://
en.wikipedia.org/wiki/Seven_generation_sustainability
Wikipedia; Henry Wriston. (2012). Retrieved May 28, 2012, from
Wikipedia: http://en.wikipedia.org/wiki/Henry_Merritt_Wriston
Williams, J. D. (1966). The compleat strategyst. New York: McGraw-
Hill Book Company.
Wolfram, S. (1992). Mathematica: A System for Doing the Mathematics
by Computer (Second ed.). Reading, MA: Addison-Wesley.
Zermelo, E. (1913). Uber eine Anwendung der Mengenlelhre auf die
Theorie des Schachspiels. Proceedings Fifth International
Congress of Mathematics, 2, pp. 501-504.
May 2, 2013 14:6 BC: 8831 - Probability and Statistical Theory PST˙ws

This page intentionally left blank


Index

0-form, 40 altruism, 422


1-form, 41 altruists, 565
2-form, 41 Ampere’s law, 48
applied, 13
absolute acceleration, 247, 550 applied forces, 77
absolute force, 779
absolute stress, 82 beta, 274
acceleration, 50, 89, 706, 739, 742, blame game, 419
746 bond forces, 487, 760, 781, 784
acceleration potential field, 681 bond matrix, 510
acceleration vector, 145 bond shear, 222
accommodating, 266, 320, 730 bond shear tensor, 513
accountable, 422, 426 bounce, 723
act, 39 boundary difference, 389
act of will, 730 breaker waves, 708
action, 499 buyers, 456
active, 14, 65, 425, 763 buy-in, 514, 775
active acceleration, 479, 637, 716
active agents, 421, 426 capacitance, 745
active geometry, 338, 650, 661 capacitive subsystems, 729
active geometry acceleration, 247, 550 capital accumulation, 724
active metric components, 65 capital loss, 724
active null geodesic, 362 Cauchy-Kowalewsky existence
active payoffs, 272 theorem, 207, 224, 510
active strategy, 170, 192, 376 causal connectivity, 473
active-co-moving frame, 252 causal mechanisms, 714
active-internal geometry, 650 causal relationship, 16
agents, 114, 119 causality, 706, 709, 744
aggressive, 266, 320 causality assumption, 653
aggressively, 730 causality principle, 345, 346, 363
aggressor, 283 causally connected, 709

807
808 Geometry, Language and Strategy—Vol. 2

cause, 52 common good solution, 348


central, 246 common ground, 456, 463, 517
central active acceleration, 300 communication compatible, 695
central co-moving holonomic frame, communication ellipsoid, 656
211, 520 communication hyperboloid, 656
central frame model, 567 commutator, 137
central holonomic frame, 144, 231, co-moving basis, 138, 144, 164
244, 249, 495 co-moving coordinate basis, 101, 120,
central time, 211, 231, 245, 520 378, 694
centrally co-moving frame, 339 co-moving frame, 380
centrally co-moving frame hypothesis, co-moving orthonormal coordinate,
161 164
centrally co-moving hypothesis, 120, co-moving orthonormal coordinate
146, 149, 209, 211, 217, 237 basis, 138, 144, 371
centripetal acceleration, 74, 76, 124, co-moving payoffs, 141
128, 133 co-moving player payoff, 554, 563
certainty, 438 co-moving player strategy bias, 563,
challenges, 777 564, 691
chaos, 332 competition, 262, 724
characteristic behavior, 12 competitive acceleration, 247, 551
characteristic payoff, 74 competitive games, 446
characteristic players, 780 competitive valuation force, 779
characteristic vector potential, 215 composite payoff, 508, 511, 513, 559,
charge, 193, 265 652, 693
charge of player j, 192 composite scalar field, 688
choice region, 351, 383, 403 conductivity, 202, 212, 213, 217
circulation, 41, 49 conductivity model, 118, 161, 200, 220,
closed loop thinking, 470 242, 516, 536
closed loops, 453, 471 connected, 709
Club of Rome, 467 connection, 60, 309
coalition, 458 connectivity, 479
code of conduct, 7, 38, 120, 262, 234, consensus, 458
263, 277, 282, 324, 371, 413, 420, conserved, 106
421, 431, 485, 502, 529, 706, 759, conserved charge, 71
760, 783, 793 conserved flow, 612
code of conduct behavior, 265 constraints, 13
code of conduct group, 793 consumers, 456
code of conduct strategy, 566 context, 100, 460
coercion, 458 contextual frame, 131, 461, 478
cognition, 458 contextual frame potential, 461
common cause, 135 continuity viewpoint, 121
common good, 306 continuous change, 50
Index 809

continuous matter, 80 decision process value, 400


contravariant vector, 62 decision relative preference, 501
control, 421 decision value field, 380
convertible, 477, 478 dependent, 372, 423
cooperation, 262, 724 dependent players, 218
cooperation potential, 452, 456 determinism, 430, 445, 709
cooperative acceleration, 247, 551 diffeomorphism, 89
cooperative behaviors, 265 distance, 88
cooperative opportunity, 479 distributed element systems, 728
cooperative payoffs, 134, 567 divergence equations, 194
cooperative shear, 567 dominates, 447
cooperative support force, 779 dual tensor, 45
coordinate basis, 60, 121 dummy, 423
coordinate displacement, 58 dynamic competition, 431
Coriolis effect, 74, 76, 124, 128, 133 dynamic cooperation, 431
Coulomb’s Law, 46 dynamic law of competition, 430, 445
covariance principle, 98 dynamic law of cooperation, 431
covariance, 54 dynamic law of organization, 431
covariant, 33, 58, 61, 92 dynamic scenarios, 265
covariant derivative, 57, 63 dynamic strategy bonding fields, 338
covariant gauge, 571
covariant vector, 62 economic equivalence principle, 445,
create the future, 798 453, 463
critical exponents, 714 effect, 52
cross product, 41 effort, 724
curl equations, 194 egoists, 565
current density, 45 elastic perfect fluid, 102
current of player α, 201 elasticity, 479, 723
curvature tensor, 64 electric field, 28, 194, 195
electro-gravitational waves, 210
Decision Analysis, 470 electromagnetic field, 31
decision engineering, 6, 262 electromagnetic radiation, 30
decision isobars, 782, 790 energy, 201, 727, 729, 745
decision mass, 273, 596, 628, 636, 743 energy density, 199, 201
decision preference scale, 500 energy flow, 170, 192
decision pressure process, 692 energy flow velocity, 617
decision process pressure, 679 energy storage, 728
decision process theory, 7, 9, 13, 14, energy-momentum stress tensor, 84
20, 24, 38, 39, 50, 51, 54, 62, 65, engagement, 38, 193, 514
69, 73, 75, 95, 108, 114, 115, 119, engagement payoff, 766
135, 149, 163, 189, 217, 261, 265, entitled, 671, 730, 747
335, 414, 470, 650 entitlement, 38, 759, 765, 783
810 Geometry, Language and Strategy—Vol. 2

entitlement payoff, 558, 563, 765 free fall behavior, 249


equilibrium, 24 freeway game, 420
equilibrium strategy, 25, 269 frequency domain, 650
equivalent decision process, 234
ethics, 422 game beta, 348
exact, 66, 124, 127, 143, 147 game equivalency, 547, 569, 578, 706
exclusive ownership, 529 game equivalency principle, 546
expansion tensor, 145 game equivalent payoffs, 579
experience, 722 game strategy, 547
extensive form, 35, 429 game theory, 14, 165, 334, 408
external geometry, 650 game value, 24, 25, 274, 554, 564, 652,
692
factions, 573 gauge independent holonomic, 796
far zone, 391 gauge structure, 360
Faraday’s Law, 44 gauge transformation, 62
field intensity, 43 gauge transformation associated with
field storage, 706 player j, 66
finite element analysis, 230 general coordinate basis, 56
fixed coordinate basis, 59 geodesic curves, 67
flows, 471 geometric forms, 703
flow-β, 696 geometric structure, 61
flux, 47 geometry, 649
focus boundary, 548 geometry of change, 37
focus direction, 532, 557 givers, 266, 456, 463
focus strategy, 547 glaring, 424
focus subspace, 548 global connections, 470, 477
focused strategy, 529 global connectivity, 473
foraging, 376 gradient, 153, 166
forced behavior, 12, 739 gradient behaviors, 155
forced harmonic, 592, 595, 621, 628, gradient effects, 723
744 gravitas, 745
form factor, 312 gravitas subsystems, 729
fractional investment, 443 gravitational radiation waves, 391
frame derivative, 58 gravitomagnetism, 133
frame effects, 478 greedy, 266, 320
frame of reference, 92
frame rotation, 146 harmonic, 706
frame-wave equation, 226, 231 harmonic behaviors, 331
frame-wave phasors, 380 harmonic coordinates, 129, 130, 143
frame-waves, 231, 520 harmonic gauge, 129, 142, 246, 521
free fall, 496, 545, 550, 551, 582, 641, harmonic gauge condition, 130
706, 734, 745 harmonic polynomials, 228, 235, 255
Index 811

harmonic power series approximation, inertial flow, 262


583, 637 inertial forces, 76, 77, 81, 95, 100
harmonic steady state, 337 inertial frame of reference, 51, 76
harmonic steady state waves, 12, 716, inertial mass, 481
739, 781 inertial media, 77, 78, 85, 87, 88
harmonics, 231 inertial potential, 304
hedge strategy, 570 inertial pressure, 324
hidden, 65, 126 inertial sources, 101
hidden-in-plain-sight, 760, 781, 784 initial boundary, 549
high frequency harmonics, 706, 744 institutional player, 528, 562
holonomic, 120, 122, 147 integral curve, 73, 214
holonomic co-moving coordinate intensive form, 36, 376, 430
basis, 120 interactive decision theory, 409
holonomic constraints, 21 inter-dependence, 422
holonomic coordinate basis, 53 interest flow, 70, 456, 463, 464, 483,
honoring contracts, 706, 725, 745 489
humanity, 416 interest flow density, 453
hyper-cylinder, 74 interest opportunity, 479
internal geometry, 650
impending event, 202 internal group, 137
inactive, 14, 65, 425, 763 internal maps, 425
inactive dimensions, 65 internal payoff, 275, 444, 573
inactive geometry, 650 internal preference, 444
inactive metric components, 65 internal symmetry, 120
inactive player, 404 intertance, 745
inactive strategy, 170, 192, 376 intrinsic view, 334
incomplete games, 265 invisible hand, 297, 302, 412, 413
independent players, 137 irreducible decision process, 372
indifference surface, 220 isometry, 65, 119, 136, 315, 520, 725
individual, 706, 725, 745 isotropic, 100
individual player, 528
individuality, 706 just decision process, 281
inductance, 745 just solution, 282
inductive subsystems, 729
inequality flow, 612 killing field, 421
inertance, 729 Killing vector, 170
inertia, 265, 464, 481, 489, 743 kinetic energy, 19
inertial, 76, 146 known behaviors, 7, 284, 373, 380,
inertial acceleration, 301, 303, 324 485, 487, 489, 503
inertial behaviors, 77 known conditions, 469, 485
inertial cohesion, 78, 81, 84
inertial currents, 132 Lagrangian, 22
812 Geometry, Language and Strategy—Vol. 2

law of competition, 452, 463 Nash equilibrium, 25, 462, 481


law of cooperation, 456, 463, 482 Nash solution, 306, 348
law of opportunity, 360 natural harmonics, 592
law of organization, 464, 478 natural resonance, 552, 569, 581
Lie product, 125, 137 natural taxonomy, 371
limits to choice, 383, 389, 390, 403, natural units, 109, 371
468, 567, 706, 745 near zone, 391
limits to growth, 468, 567 negotiation fields, 47, 70
line element, 65 negotiation flux, 47
line integral, 56 negotiation plane, 47
linear programming, 449 negotiation plane normal, 48
linear transformation, 443 negotiation strategies, 47
locality, 14, 50, 707 net-wealth-strategy, 613
locally flat, 50, 52, 58 network-strategy, 613
locked behavior, 525, 567, 595, 596, network connections, 706, 714, 744,
733 753
longitudinal, 676 network connectivity, 715, 722
Lorentz metric, 91 network player bonding, 723
lumped element approximation, 745 network relationship, 16, 52
lumped element systems, 728 neutral, 556
neutral decision process, 556
magnetic field, 194, 195 neutrality principle, 732
magnetic flux, 44 n-forms, 42
mass, 17 no strategic bias, 555, 575
mathematical stability, 264 non-dependent player, 218
matter, 79 non-inertial, 76
measure, 51 non-inertial forces, 76
measure independence, 51, 59 non-inertial frame, 76
me-conversation, 420 non-zero sum game, 270
mesh, 583 normal form, 14, 24, 268, 371, 376,
metric, 51, 92, 315 430
Minkowski metric, 53, 560 normal frame, 495
mixed strategy, 15, 432, 463 normal time, 695
morphogenesis, 706, 708, 744 normal-form coordinate basis, 66,
morphogenic, 708 120, 124, 127, 129, 694
morphogenic changes, 722 normal-form holonomic basis, 503
moves, 15, 429 normality, 14
mutual player support, 760, 779, 783 normalized flows, 612
mutual player support potentials, 778, null geodesic, 345, 654
779 null hypothesis, 434
numerical method of lines, 230, 361,
narrative, 479, 720 475, 528
Index 813

observer players, 485 phase transitions, 722


observer-agents, 426 phasor, 520, 742
observers, 423, 485 phasor preferences, 615
opportunity, 464, 479, 483, 489 phasor representation, 495
opportunity cost, 482, 483, 489 phasor solutions, 231
opportunity energy momentum tensor, phasors, 12, 90, 136, 163, 210, 231,
478 497
organizational cooperative effects, 461 physical decision theory, 14
organizational current, 453 play, 15, 430
organizational dynamics, 100, 102, player aggression, 634, 750
200 player bonding, 196, 197, 220, 301,
organizational equilibrium, 462 324, 506, 507, 556, 652, 692
organizational field, 133 player bonding coefficient, 309
organizational opportunity, 479 player bonding potential, 223, 513,
organizational payoffs, 461 780
organizational system, 131, 198, 293, player bonding scalar, 681
302, 460 player bonding shear, 222
organizing source, 460, 461 player bonding shear components, 309
organizing sources, 100, 461, 724 player current, 193, 241
orientation flux field, 64, 73, 95, 96, player effort, 500
198, 496 player effort strategy, 547
orientation potential, 57 player engagement, 150, 171, 192,
orthogonal time, 695 193, 301, 320, 324, 505, 506, 552,
orthogonal transformation, 296 559, 652, 692, 730, 759, 770, 783
orthonormal co-moving coordinate player engagement payoff, 759, 771
basis, 120 player entitlement, 554, 555
orthonormal coordinate basis, 60 player entitlement payoff, 552, 555
ownership, 425 player fixed frame model, 120, 147,
ownership model, 220, 242, 516 149, 164, 167, 169, 191, 485, 496,
owns, 425 796
player impact, 516, 519, 776
parallel translation, 57, 62 player independence hypothesis, 252
passion, 220, 515 player interest, 196, 197, 301, 324,
passively, 730 355, 513, 564, 652, 692, 759, 775,
payoff direction, 451 783
payoff field, 68 player interest field, 377
payoff potentials, 65 player interest flow, 222, 266, 375,
payoffs, 23 453, 456, 458, 482
perfect fluid, 200 player interests, 317
persistency, 14, 65, 124, 166, 191 player ownership, 220
persistent, 70, 124, 137 player ownership model, 796
persistent behaviors, 7, 119 player ownership potential, 219
814 Geometry, Language and Strategy—Vol. 2

player ownership rule, 218 proper-active, 147


player ownership structures, 252, 796 proper-inactive, 147
player passion, 218, 506, 515, 564, proper-player, 184
651, 668, 692, 760, 777, 783 proper-time, 214, 284, 337, 503
player payoff, 331, 515, 765 proper-time coordinate, 147, 151
player payoff vector, 511 public-interest, 302, 413, 422
player potential, 170 pure strategies, 15, 430
player potential vector fields, 193 purely resource driven processes, 502
player preference scale, 279, 500 purely skill driven processes, 502
player relative preference, 279, 500
player scale, 282 quantum of action, 108
player self-support, 780
player shear, 652, 692 random events, 11
player shear matrix, 681 rational behavior, 438
player shear potential, 780 reduced bond shear potential, 326
player skill preferences, 547 reduced player strategy bias, 677
player stakes, 273, 348, 362 reduced pressure tensor, 221
player strategy bias, 554, 652, 692 reduced transfer function, 619
players, 16, 71, 114, 119, 191, 193, reduced transfer modulus, 619, 650
485, 797 reduced transfer phase, 650
polarization, 732 reduced transfer phase shift, 619
political commentary game, 420 reflection and transmission scenarios,
population value rule, 611 261
positive energy density assumption, relative player effort, 281, 282, 573
654 relative stress, 80
potential energy, 19 resilience, 200, 294, 382, 480
prefer, 443 resonant behaviors, 331
preference, 438, 500 resonant harmonic, 592
pressure, 102, 199, 201 resonant harmonic frequency, 594
pressure gradients, 723 resonant scenarios, 261, 271
principle of least action, 13, 77, 78, rigid bodies, 84
116 risk, 438
principle of opportunity, 360 rotating frames, 56
principle of possible change, 346, 363 rules of the game, 763
Prisoner Dilemma Model (PDM), 304
producers, 456 scalar field, 55, 56, 61
productivity factor, 40, 46 scale invariance, 548, 572, 574
propaganda, 606 scaled payoff, 45
proper active strategy, 170, 192 seasonal effects, 569
proper charge, 150 seasonal gauge, 256, 571, 640
proper inactive strategy, 192 seasonal harmonics, 592, 744
proper relative player effort, 286
Index 815

seasonal payoff, 241, 246, 331, 569, steady state surface geometry, 262
688, 737, 739 still point, 503, 545, 550, 582, 737
seasonal player, 570 stocks, 471
seasonal potential, 216, 246, 518, 530 Stokes’ Law, 42
seasonal timeframe, 592, 594, 612 strain, 196, 303, 309
seasonal vector potential, 241, 517, strategic boundary, 40, 49
559 strategic clustering, 567
self, 375, 422 strategic decomposition, 371
self-interest, 302, 413, 421 strategic form, 430
self-love, 416 strategic occupation, 75
sellers, 456 strategic opinion, 501
shear potential scalar, 681 strategic surface, 40, 45
signal, 52 strategic viscosity, 221
single active strategy, 234 strategic volume, 40
sinusoid, 742 strategy contours, 790
skill boundary, 556 streamline proper-time scalar field,
skill preference subspace, 556 214
social distinctions, 705 streamline solutions, 214, 234
solipsistic, 568 streamlines, 345
space dimension, 130, 147 stress, 309
spacelike, 141 stress tensor, 198, 201
spaceline, 364 stress tensor components, 101
speed, 89 strictly determined, 432, 447
spin, 196 strong ownership, 252
staff-seconds, 39 structural coupling, 200
staff-years, 39 structural distinctions, 705
standard of behavior, 277, 421, 436 structural stability, 264, 303, 706, 707,
static games, 335 708, 744
stationary, 161 structure constants, 59
stationary flow, 12 subsidy-seeking, 418
stationary holonomic frame, 796 successful decision structures, 724,
stationary line gauge, 330 745
steady state, 209, 210 sustainability, 706, 745
steady state behaviors, 403, 448 sustainable, 764
steady state flow, 12 symmetric decision process, 281, 372
steady state geometry, 299, 301, 324 symmetric normal-form coordinate
steady state harmonics, 228 basis, 284, 503
steady state hypothesis, 211, 217, 221, symmetric prisoner’s dilemma, 340
525 symmetry, 105
steady state orthonormal frame, 211 system response harmonics, 497, 591
steady state scenarios, 261, 264, 272 systems dynamics, 262
steady state solution, 447
816 Geometry, Language and Strategy—Vol. 2

takers, 266, 456, 463 uncertainty, 438


tangent vectors, 214 uncoupled, 286
temptation to default, 417, 418 utilitarian solution, 282
tensor field, 55, 56, 62 utility, 40, 437, 724
three W’s, 724 utility current, 48
tidal, 196 utility flow, 45
tidal force, 173
tidal magnetic, 197 valuation field, 47, 70
time bonding, 339, 349, 361 variability, 14, 65
time domain, 650 vector field, 56, 62
time evolution equations, 194 velocity, 89
time flow velocity, 618 volume assumption, 653
time isometry, 706, 725 vortex, 27, 69
time metric factor, 741 vortex-inducing, 14
timelike, 141 vortices, 495
timeline, 364 vorticity, 69, 76, 121, 146, 166, 217,
torsion free, 51, 58 236, 303
total effort strategy, 547 vorticity behaviors, 155, 509
total player effort, 281, 573 vorticity tensor, 145, 169
total player preference, 557
total preference scale, 762 wave equation, 142
tragedy of the commons, 419, 428 wealthy, 609
transfer function, 618, 691, 744 wedge product, 41, 59
transfer modulus, 618 wisdom, 612
transfer phase, 618 work, 19
transient behaviors, 331 workers, 609
transient scenarios, 261 worldview, 266, 375
transverse, 676 WWW model, 612
travelling frame-wave, 386
true players, 485 zero-sum games, 25, 269, 575
SERIES  ON  KNOTS  AND  EVERYTHING

Editor-in-charge: Louis H. Kauffman (Univ. of Illinois, Chicago)

The Series on Knots and Everything: is a book series polarized around the theory of
knots. Volume 1 in the series is Louis H Kauffman’s Knots and Physics.

One purpose of this series is to continue the exploration of many of the themes
indicated in Volume 1. These themes reach out beyond knot theory into physics,
mathematics, logic, linguistics, philosophy, biology and practical experience. All
of these outreaches have relations with knot theory when knot theory is regarded as
a pivot or meeting place for apparently separate ideas. Knots act as such a pivotal
place. We do not fully understand why this is so. The series represents stages in the
exploration of this nexus.

Details of the titles in this series to date give a picture of the enterprise.

Published*:

Vol. 1: Knots and Physics (Third Edition)


by L. H. Kauffman
Vol. 2: How Surfaces Intersect in Space — An Introduction to Topology (Second Edition)
by J. S. Carter
Vol. 3: Quantum Topology
edited by L. H. Kauffman & R. A. Baadhio
Vol. 4: Gauge Fields, Knots and Gravity
by J. Baez & J. P. Muniain
Vol. 5: Gems, Computers and Attractors for 3-Manifolds
by S. Lins
Vol. 6: Knots and Applications
edited by L. H. Kauffman
Vol. 7: Random Knotting and Linking
edited by K. C. Millett & D. W. Sumners
Vol. 8: Symmetric Bends: How to Join Two Lengths of Cord
by R. E. Miles
Vol. 9: Combinatorial Physics
by T. Bastin & C. W. Kilmister
Vol. 10: Nonstandard Logics and Nonstandard Metrics in Physics
by W. M. Honig
Vol. 11: History and Science of Knots
edited by J. C. Turner & P. van de Griend

*The complete list of the published volumes in the series can also be found at
http://www.worldscientific.com/series/skae

EH - Geometry, Language and Strategy.indd 1 18-08-16 2:33:11 PM


Vol. 12: Relativistic Reality: A Modern View
edited by J. D. Edmonds, Jr.
Vol. 13: Entropic Spacetime Theory
by J. Armel
Vol. 14: Diamond — A Paradox Logic
by N. S. Hellerstein
Vol. 15: Lectures at KNOTS ’96
by S. Suzuki
Vol. 16: Delta — A Paradox Logic
by N. S. Hellerstein
Vol. 17: Hypercomplex Iterations — Distance Estimation and Higher Dimensional Fractals
by Y. Dang, L. H. Kauffman & D. Sandin
Vol. 18: The Self-Evolving Cosmos: A Phenomenological Approach to Nature’s
Unity-in-Diversity
by S. M. Rosen
Vol. 19: Ideal Knots
by A. Stasiak, V. Katritch & L. H. Kauffman
Vol. 20: The Mystery of Knots — Computer Programming for Knot Tabulation
by C. N. Aneziris
Vol. 21: LINKNOT: Knot Theory by Computer
by S. Jablan & R. Sazdanovic
Vol. 22: The Mathematics of Harmony — From Euclid to Contemporary Mathematics and
Computer Science
by A. Stakhov (assisted by S. Olsen)
Vol. 23: Diamond: A Paradox Logic (Second Edition)
by N. S. Hellerstein
Vol. 24: Knots in HELLAS ’98 — Proceedings of the International Conference on Knot
Theory and Its Ramifications
edited by C. McA Gordon, V. F. R. Jones, L. Kauffman, S. Lambropoulou &
J. H. Przytycki
Vol. 25: Connections — The Geometric Bridge between Art and Science (Second Edition)
by J. Kappraff
Vol. 26: Functorial Knot Theory — Categories of Tangles, Coherence, Categorical
Deformations, and Topological Invariants
by David N. Yetter
Vol. 27: Bit-String Physics: A Finite and Discrete Approach to Natural Philosophy
by H. Pierre Noyes; edited by J. C. van den Berg
Vol. 28: Beyond Measure: A Guided Tour Through Nature, Myth, and Number
by J. Kappraff
Vol. 29: Quantum Invariants — A Study of Knots, 3-Manifolds, and Their Sets
by T. Ohtsuki
Vol. 30: Symmetry, Ornament and Modularity
by S. V. Jablan

EH - Geometry, Language and Strategy.indd 2 18-08-16 2:33:11 PM


Vol. 31: Mindsteps to the Cosmos
by G. S. Hawkins
Vol. 32: Algebraic Invariants of Links
by J. A. Hillman
Vol. 33: Energy of Knots and Conformal Geometry
by J. O’Hara
Vol. 34: Woods Hole Mathematics — Perspectives in Mathematics and Physics
edited by N. Tongring & R. C. Penner
Vol. 35: BIOS — A Study of Creation
by H. Sabelli
Vol. 36: Physical and Numerical Models in Knot Theory
edited by J. A. Calvo et al.
Vol. 37: Geometry, Language, and Strategy
by G. H. Thomas
Vol. 38: Current Developments in Mathematical Biology
edited by K. Mahdavi, R. Culshaw & J. Boucher
Vol. 39: Topological Library
Part 1: Cobordisms and Their Applications
edited by S. P. Novikov & I. A. Taimanov
Vol. 40: Intelligence of Low Dimensional Topology 2006
edited by J. Scott Carter et al.
Vol. 41: Zero to Infinity: The Fountations of Physics
by P. Rowlands
Vol. 42: The Origin of Discrete Particles
by T. Bastin & C. Kilmister
Vol. 43: The Holographic Anthropic Multiverse
by R. L. Amoroso & E. A. Ranscher
Vol. 44: Topological Library
Part 2: Characteristic Classes and Smooth Structures on Manifolds
edited by S. P. Novikov & I. A. Taimanov
Vol. 45: Orbiting the Moons of Pluto
Complex Solutions to the Einstein, Maxwell, Schrödinger and Dirac Equations
by E. A. Rauscher & R. L. Amoroso
Vol. 46: Introductory Lectures on Knot Theory
edited by L. H. Kauffman, S. Lambropoulou, S. Jablan & J. H. Przytycki
Vol. 47: Introduction to the Anisotropic Geometrodynamics
by S. Siparov
Vol. 48: An Excursion in Diagrammatic Algebra: Turning a Sphere from Red to Blue
by J. S. Carter
Vol. 49: Hopf Algebras
by D. E. Radford
Vol. 50: Topological Library
Part 3: Spectral Sequences in Topology
edited by S. P. Novikov & I. A. Taimanov

EH - Geometry, Language and Strategy.indd 3 18-08-16 2:33:11 PM


Vol. 51: Virtual Knots: The State of the Art
by V. O. Manturov & D. P. Ilyutko
Vol. 52: Algebraic Invariants of Links (Second Edition)
by J. Hillman
Vol. 53: Knots and Physics (Fourth Edition)
by L. H. Kauffman
Vol. 54: Scientific Essays in Honor of H Pierre Noyes on the Occasion of His 90th Birthday
edited by J. C. Amson & L. H. Kauffman
Vol. 55: Knots, Braids and Möbius Strips
by J. Avrin
Vol. 56: New Ideas in Low Dimensional Topology
edited by L. H. Kauffman & V. O. Manturov
Vol. 57: ADEX Theory: How the ADE Coxeter Graphs Unify Mathematics and Physics
by S.-P. Sirag
Vol. 58: Representing 3-Manifolds by Filling Dehn Surfaces
by R. Vigara & A. Lozano-Rojo
Vol. 59: Geometry, Language and Strategy: The Dynamics of Decision Processes — Vol. 2
by G. H. Thomas

EH - Geometry, Language and Strategy.indd 4 18-08-16 2:33:11 PM

You might also like