You are on page 1of 53

Complex cobordism and stable

homotopy groups of spheres Ravenel


D.C
Visit to download the full and correct content document:
https://textbookfull.com/product/complex-cobordism-and-stable-homotopy-groups-of-s
pheres-ravenel-d-c/
More products digital (pdf, epub, mobi) instant
download maybe you interests ...

Foundations of stable homotopy theory First Published


Edition Cambridge University Press.

https://textbookfull.com/product/foundations-of-stable-homotopy-
theory-first-published-edition-cambridge-university-press/

Complex Semisimple Quantum Groups and Representation


Theory 1st Edition Christian Voigt

https://textbookfull.com/product/complex-semisimple-quantum-
groups-and-representation-theory-1st-edition-christian-voigt/

Stable isotope forensics : methods and forensic


applications of stable isotope analysis Second Edition
Meier-Augenstein

https://textbookfull.com/product/stable-isotope-forensics-
methods-and-forensic-applications-of-stable-isotope-analysis-
second-edition-meier-augenstein/

Theory of Groups and Symmetries Finite groups Lie


groups and Lie Algebras 1st Edition Alexey P. Isaev

https://textbookfull.com/product/theory-of-groups-and-symmetries-
finite-groups-lie-groups-and-lie-algebras-1st-edition-alexey-p-
isaev/
Spheres of Transnational Ecoviolence: Environmental
Crime, Human Security, and Justice Peter Stoett

https://textbookfull.com/product/spheres-of-transnational-
ecoviolence-environmental-crime-human-security-and-justice-peter-
stoett/

Handbook of Homotopy Theory 1st Edition Haynes Miller


(Editor)

https://textbookfull.com/product/handbook-of-homotopy-theory-1st-
edition-haynes-miller-editor/

Handbook of Homotopy Theory 1st Edition Haynes Miller


(Editor)

https://textbookfull.com/product/handbook-of-homotopy-theory-1st-
edition-haynes-miller-editor-2/

Stein Manifolds and Holomorphic Mappings The Homotopy


Principle in Complex Analysis Ergebnisse der Mathematik
und ihrer Grenzgebiete 3 Folge A Series of Modern
Surveys in Mathematics Franc Forstneri■
https://textbookfull.com/product/stein-manifolds-and-holomorphic-
mappings-the-homotopy-principle-in-complex-analysis-ergebnisse-
der-mathematik-und-ihrer-grenzgebiete-3-folge-a-series-of-modern-
surveys-in-mathematics-franc-forstneric/

Theory of Groups and Symmetries: Representations of


Groups and Lie Algebras, Applications 1st Edition
Alexey P. Isaev

https://textbookfull.com/product/theory-of-groups-and-symmetries-
representations-of-groups-and-lie-algebras-applications-1st-
edition-alexey-p-isaev/
Complex Cobordism and
Stable Homotopy Groups of Spheres

Douglas C. Ravenel

Department of Mathematics, University of Rochester, Rochester,


New York
To my wife, Michelle
Contents

List of Figures xi

List of Tables xiii

Preface to the first edition xv

Preface to the second edition xvii

Commonly Used Notations xix

Chapter 1. An Introduction to the Homotopy Groups of Spheres 1

1. Classical Theorems Old and New 2


Homotopy groups. The Hurewicz and Freudenthal theorems. Stable stems.
The Hopf map. Serre’s finiteness theorem. Nishida’s nilpotence theorem. Cohen,
Moore and Neisendorfer’s exponent theorem. Bott periodicity. The J-homomorphism.
2. Methods of Computing π∗ (S n ) 5
Eilenberg–Mac Lane spaces and Serre’s method. The Adams spectral sequence.
Hopf invariant one theorems. The Adams–Novikov spectral sequence. Tables in low
dimensions for p = 3.
3. The Adams–Novikov E2 -term, Formal Group Laws, and the Greek
Letter Construction 12
Formal group laws and Qillen’s theorem. The Adams–Novikov E2 -term as
group cohomology. Alphas, beta and gamma. The Morava–Landweber theorem
and higher Greek letters. Generalized Greek letter elements.
4. More Formal Group Law Theory, Morava’s Point of View, and the
Chromatic Spectral Sequence 19
The Brown–Peterson spectrum. Classification of formal group laws. Morava’s
group action, its orbits and stabilizers. The chromatic resolution and the chromatic
spectral sequence. Bockstein SSs. Use of cyclic subgroups to detect Arf invariant
elements. Morava’s vanishing theorem. Greek letter elements in the chromatic
spectral sequence.
5. Unstable Homotopy Groups and the EHP Spectral Sequence 24
The EHP sequences. The EHP SS. The stable zone. The inductive method.
The stable EHP SS. The Adams vector field theorem. James periodicity. The
J-spectrum. The SS for J∗ (RP ∞ ) and J∗ (BΣp ). Relation to the Segal conjecture.
The Mahowald root invariant.

Chapter 2. Setting up the Adams Spectral Sequence 41

v
vi CONTENTS

1. The Classical Adams Spectral Sequence 41


Mod (p) Eilenberg–Mac Lane spectra. Mod (p) Adams resolutions. Differen-
tials. Homotopy inverse limits. Convergence. The extension problem. Examples:
integral and mod (pi ) Eilenberg–Mac Lane spectra.
2. The Adams Spectral Sequence Based on a Generalized Homology
Theory 49
E∗ -Adams resolutions. E-completions. The E∗ -Adams spectral sequence. As-
sumptions on the spectrum E. E∗ (E) is a Hopf algebroid. The canonical Adams
resolution. Convergence. The Adams filtration.
3. The Smash Product Pairing and the Generalized Connecting
Homomorphism 53
The smash product induces a pairing in the Adams spectral sequence. A map
that is trivial in homology raises Adams filtration. The connecting homomorphism
in Ext and the geometric boundary map.

Chapter 3. The Classical Adams Spectral Sequence 59

1. The Steenrod Algebra and Some Easy Calculations 59


Milnor’s structure theorem for A∗ . The cobar complex. Multiplication by p in
the E∞ -term. The Adams spectral sequence for π∗ (M U ). Computations for M O,
bu and bo.
2. The May Spectral Sequence 67
May’s filtration of A∗ . Nonassociativity of May’s E1 -term and a way to avoid
it. Computations at p = 2 in low dimensions. Computations with the subalgebra
A(2) at p = 2.
3. The Lambda Algebra 77
Λ as an Adams E1 -term. The algebaraic EHP SS. Serial numbers. The Curtis
algorithm. Computations below dimension 14. James periodicity. The Adams
vanishing line. d1 is multiplication by λ−1 . Illustration for S 3 .
4. Some General Properties of Ext 86
Exts for s ≤ 3. Behavior of elements in Ext2 . Adams’ vanishing line of slope
1/2 for p = 2. Periodicity above a line of slope 1/5 for p = 2. Elements not
annihilated by any periodicity operators and their relation to im J. An elementary
proof that most of these elements are nontrivial.
5. Survey and Further Reading 94
Exotic cobordism theories. Decreasing filtrations of A∗ and the resulting SSs.
Application to M Sp. Mahowald’s generalizations of Λ. vn -periodicity in the Adams
spectral sequence. Selected references to related work.

Chapter 4. BP -Theory and the Adams–Novikov Spectral Sequence 103

1. Quillen’s Theorem and the Structure of BP∗ (BP ) 103


Complex cobordism. Complex orientation of a ring spectrum. The formal
group law for a complex oriented homology theory. Quillen’s theorem equating the
Lazard and complex cobordism rings. Landweber and Novikov’s theorem on the
structure of M U∗ (M U ). The Brown–Peterson spectrum BP . Quillen’s idempotent
operation and p-typical formal group laws. The structure of BP∗ (BP ).
2. A Survey of BP -Theory 111
Bordism groups of spaces. The Sullivan–Baas construction. The Johnson–
Wilson spectrum BP hni. The Morava K-theories K(n). The Landweber filtration
CONTENTS vii

and exact functor theorems. The Conner–Floyd isomorphism. K-theory as a func-


tor of complex cobordism. Johnson and Yosimura’s work on invariant regular ideals.
Infinite loop spaces associated with M U and BP ; the Ravenel–Wilson Hopf ring.
The unstable Adams–Novikov spectral sequence of Bendersky, Curtis and Miller.
3. Some Calculations in BP∗ (BP ) 117
The Morava-Landweber invariant prime ideal theorem. Some invariant regular
ideals. A generalization of Witt’s lemma. A formula for the universal p-typical
formal group law. Formulas for the coproduct and conjugation in BP∗ (BP ). A
filtration of BP∗ (BP ))/In .
4. Beginning Calculations with the Adams–Novikov Spectral Sequence 130
and Miller. Low dimensional Ext of the algebra of Steenrod reduced powers.
Bockstein SSs leading to the Adams–Novikov E2 -term. Calculations at odd primes.
Toda’s theorem on the first nontrivial odd primary Novikov differential. Chart for
p = 5. Calculations and charts for p = 2. Comparison with the Adams spectral
sequence.

Chapter 5. The Chromatic Spectral Sequence 147

1. The Algebraic Construction 148


Greek letter elements and generalizations. The chromatic resolution, SS, and
cobar complex. The Morava stabilizer algebra Σ(n). The change-of-rings theorem.
The Morava vanishing theorem. Signs of Greek letter elements. Computations with
βt . Decompsibility of γ1 . Chromatic differentials at p = 2. Divisibility of α1 βp .
2. Ext1 (BP∗ /In ) and Hopf Invariant One 158
Ext0 (BP∗ ). Ext0 (M10 ). Ext1 (BP∗ ). Hopf invariant one elements. The Miller-
Wilson calculation of Ext1 (BP∗ /In ).
3. Ext(M 1 ) and the J-Homomorphism 165
Ext(M 1 ). Relation to im J. Patterns of differentials at p = 2. Computations
with the mod (2) Moore spectrum.
4. Ext2 and the Thom Reduction 172
Results of Miller, Ravenel and Wilson (p > 2) and Shimomura (p = 2) on
Ext2 (BP∗ ). Behavior of the Thom reduction map. Arf invariant differentials at
p > 2. Mahowald’s counterexample to the doomsday conjecture.
5. Periodic Families in Ext2 178
Smith’s construction of βt . Obstructions at p = 3. Results of Davis, Mahowald,
Oka, Smith and Zahler on permanent cycles in Ext2 . Decomposables in Ext2 .
6. Elements in Ext3 and Beyond 183
Products of alphas and betas in Ext3 . Products of betas in Ext4 . A possible
obstruction to the existence of V (4).

Chapter 6. Morava Stabilizer Algebras 187

1. The Change-of-Rings Isomorphism 187


Theorems of Ravenel and Miller. Theorems of Morava. General nonsense about
Hopf algebroids. Formal group laws of Artin local rings. Morava’s proof. Miller
and Ravenel’s proof.
viii CONTENTS

2. The Structure of Σ(n) 192


Relation to the group ring for Sn . Recovering the grading via an eigenspace
decomposition. A matrix representation of Sn . A splitting of Sn when p - n.
Poincaré duality and and periodic cohomology of Sn .
3. The Cohomology of Σ(n) 198
A May filtration of Σ(n) and the May SS. The open subgroup theorem. Coho-
mology of some associated Lie algebras. H 1 and H 2 . H ∗ (S(n)) for n = 1, 2, 3.
4. The Odd Primary Kervaire Invariant Elements 213
The nonexistence of certain elements and spectra. Detecting elements with the
cohomology of Z/(p). Differentials in the Adams spectral sequence.
5. The Spectra T (m) 220
A splitting theorem for certain Thom spectra. Application of the open subgroup
theorem. Ext0 and Ext1 .

Chapter 7. Computing Stable Homotopy Groups with the Adams–Novikov


Spectral Sequence 225

1. The method of infinite descent 227


Γ(m+1), A(m), and G(m+1, k −1). Weak injective comodules. Quillen opera-
tions. i-free comodules. The small descent spectral sequence and topological small
descent spectral sequence. Input/output procedure. The 4-term exact sequence.
Hat notation. A generalization of the Morava-Landweber theorem. The λ bi and
0
Dm+1 . Poincaré series. Ext1Γ(m+1) . Properties of weak injectives.
2
2. The comodule Em+1 238
1 2 2
D1 . The Poincaré series for Em+1 . Ext(Em+1 ) below dimension p2 |b v1 |. The
comodules B and U .
3. The homotopy of T (0)(2) and T (0)(1) 249
The j-freeness of B. The elements ui,j . A short exact sequence for U . π∗ (T (0)(2) ).
Cartan–Eilenberg differentials for T (0)(1) .
4. The proof of Theorem 7.3.15 261
P (1)∗ . Ẽ12 , D̃12 , and Ẽ13 . P -free comodules. A filtration of D̃12 . A 4-term exact
sequence of P (1)∗ -comodules. Ci and the skeletal filtration SS.
5. Computing π∗ (S 0 ) for p = 3 277
The input list I. Computation of differentials in the topological small descent
spectral sequence.
6. Computations for p = 5 282
The differential on γ3 and the nonexistence of V (3) for p = 5. The input list I.
Differentials.

Appendix A1. Hopf Algebras and Hopf Algebroids 299

1. Basic Definitions 301


Hopf algebroids as cogroup objects in the category of commutative algebras.
Comodules. Cotensor products. Maps of Hopf algebroids. The associated Hopf
algebra. Normal maps. Unicursal Hopf algebroids. The kernel of a normal map.
Hopf algebroid extensions. The comodule algebra structure theorem. Invariant
ideals. Split Hopf algebroids.
CONTENTS ix

2. Homological Algebra 309


Injective comodules. The derived functors Cotor and Ext. Relative injectives
and resolutions. The cobar resolution and complex. Cup products. Ext isomor-
phisms for invariant ideals and split Hopf algebroids.
3. Some Spectral Sequences 315
The resolution spectral sequence. Filtered Hopf algebroids. Filtrations by
powers of the unit coideal. The SSassicated with a Hopf algebroid map. Change-
of-rings isomorphism. The Cartan–Eilenberg spectral sequence. A formulation due
to Adams. The E2 -term for a cocentral extesion of Hopf algebras.
4. Massey Products 323
Definitions of n-fold Massey products and indeterminacy. Defining systems.
Juggling theorems: associativity and commutativity formulas. Convergence of
Massey products in SSs. A Leibnitz formula for differentials. Differentials and
extensions in SSs.
5. Algebraic Steenrod Operations 332
Construction, Cartan formula and Adem relations. Commutativity with sus-
pension. Kudo transgression theorem.
Appendix A2. Formal Group Laws 339

1. Universal Formal Group Laws and Strict Isomorphisms 339


Definition and examples of formal group laws. Homomorphisms, isomorphisms
and logarithms. The universal formal group law and the Lazard ring. Lazard’s
comparison lemma. The Hopf algebroid V T . Proof of the comparison lemma.
2. Classification and Endomorphism Rings 351
Hazewinkel’s and Araki’s generators. The right unit formula. The height of a
formal group law. Classification in characteristic p. Finite fields, Witt rings and
division algebras. The endomorphism ring of a height n formal group law.
Appendix A3. Tables of Homotopy Groups of Spheres 361
Bibliography 377
List of Figures

1.2.15 The Adams spectral sequence for p = 3, t − s ≤ 45. 11


1.2.19 The Adams–Novikov spectral sequence for p = 3, t − s ≤ 45 13
1.5.9 The EHP spectral sequence for p = 2 and k ≤ 7. 27
1.5.24 A portion of the E2 -term of the spectral sequence converging to
J∗ (RP ∞ ). 36

3.2.9 The May E2 -term for p = 2 and t − s ≤ 13 72


3.2.17 The May spectral sequence for ExtA(2)∗ (Z/(2), A(0)∗ ). 76
3.3.10 The EHP spectral sequence (3.3.7) for t − s ≤ 14 81
3.3.18 The unstable Adams E2 -term for S 3 . 85
3.4.20 Exts−1,t H∗ (W ). 94

4.4.16 Exts,t
BP∗ (BP ) (BP∗ , BP∗ /I1 ) for p = 5 and t − s ≤ 240. 136
4.4.21 The Adams–Novikov spectral sequence for p = 5, t − s ≤ 240, and
s ≥ 2. 138
4.4.23 Ext(BP∗ /In ) for p = 2 and t − s < 29 for n = 2, 3, 4. 140
4.4.32 Ext(BP∗ /I1 ) for p = 2 and t − s ≤ 26 142
4.4.45 Ext(BP∗ ) for p = 2, t − s ≤ 25. 144
4.4.46 ExtA∗ (Z/2, Z/2) for t − s ≤ 25. 145
(1)
7.3.17 ExtΓ(1) (T0 ) 262

A3.1a The Adams spectral sequence for p = 2, t − s ≤ 29. 362


A3.1b The Adams spectral sequence for p = 2, 28 ≤ t − s ≤ 45 363
A3.1c The Adams spectral sequence for p = 2, 44 ≤ t − s ≤ 61. 364
A3.2 The Adams–Novikov spectral sequence for p = 2, t − s ≤ 39. 365

xi
List of Tables

4.4.48 Correspondence between Adams–Novikov spectral sequence and Adams


spectral sequence permanent cycles for p = 2, 14 ≤ t − s ≤ 24 146

A3.3 π∗S at p = 2a 366


A3.4 3-Primary Stable Homotopy Excluding im J a 369
A3.5 5-Primary Stable Homotopy Excluding im J 370
A3.6 Toda’s calculation of πn+k (S n ) for k ≤ 19. 376

xiii
Preface to the first edition

My initial inclination was to call this book The Music of the Spheres, but I was
dissuaded from doing so by my diligent publisher, who is ever mindful of the sensi-
bilities of librarians. The purpose of this book is threefold: (i) to make BP -theory
and the Adams–Novikov spectral sequence more accessible to nonexperts, (ii) to
provide a convenient reference for workers in the field, and (iii) to demonstrate the
computational potential of the indicated machinery for determining stable homo-
topy groups of spheres. The reader is presumed to have a working knowledge of
algebraic topology and to be familiar with the basic concepts of homotopy theory.
With this assumption the book is almost entirely self-contained, the major excep-
tions (e.g., Sections 5.4, 5.4, A1.4, and A1.5) being cases in which the proofs are
long, technical, and adequately presented elsewhere.
The subject matter is a difficult one and this book will not change that fact.
We hope that it will make it possible to learn the subject other than by the only
practical method heretofore available, i.e., by numerous exhausting conversations
with one of a handful of experts. Much of the material here has been previously
published in journal articles too numerous to keep track of. However, a lot of
the foundations of the subject, e.g., Chapter 2 and Appendix 1, have not been
previously worked out in sufficient generality and the author found it surprisingly
difficult to do so.
The reader (especially if she is a graduate student) should be warned that many
portions of this volume contain more than he is likely to want or need to know. In
view of (ii), results are given (e.g., in Sections 4.3, 6.3, and A1.4) in greater strengh
than needed at present. We hope the newcomer to the field will not be discouraged
by abundance of material.
The homotopy groups of spheres is a highly computational topic. The serious
reader is strongly encouraged to reproduce and extend as many of the computations
presented here as possible. There is no substitute for the insight gained by carrying
out such calculations oneself.
Despite the large amount of information and techniques currently available,
stable homotopy is still very mysterious. Each new computational breakthrough
heightens our appreciation of the difficulty of the problem. The subject has a highly
experimental character. One computes as many homotopy groups as possible with
existing machinery, and the resulting data form the basis for new conjectures and
new theorems, which may lead to better methods of computation. In contrast with
physics, in this case the experimentalists who gather data and the theoreticians
who interpret them are the same individuals.
The core of this volume is Chapters 2–6 while Chapter 1 is a casual nontechnical
introduction to this material. Chapter 7 is a more technical description of actual
computations of the Adams–Novikov spectral sequence for the stable homotopy

xv
xvi PREFACE TO THE FIRST EDITION

groups of spheres through a large range of dimensions. Although it is likely to be


read closely by only a few specialists, it is in some sense the justification for the
rest of the book, the computational payoff. The results obtained there, along with
some similar calculations of Tangora, are tabulated in Appendix 3.
Appendices 1 and 2 are utilitarian in nature and describe technical tools used
throughout the book. Appendix 1 develops the theory of Hopf algebroids (of which
Hopf algebras are a special case) and useful homological tools such as relative
injective resolutions, spectral sequences, Massey products, and algebraic Steenrod
operations. It is not entertaining reading; we urge the reader to refer to it only
when necessary.
Appendix 2 is a more enjoyable self-contained account of all that is needed
from the theory of formal group laws. This material supports a bridge between
stable homotopy theory and algebraic number theory. Certain results (e.g., the
cohomology of some groups arising in number theory) are carried across this bridge
in Chapter 6. The house they inhabit in homotopy theory, the chromatic spectral
sequence, is built in Chapter 5.
The logical interdependence of the seven chapters and three appendixes is dis-
played in the accompanying diagram.
It is a pleasure to acknowledge help received from many sources in preparing
this book. The author received invaluable editorial advice from Frank Adams, Peter
May, David Pengelley, and Haynes Miller. Steven Mitchell, Austin Pearlman, and
Bruce McQuistan made helpful comments on various stages of the manuscript,
which owes its very existence to the patient work of innumerable typists at the
University of Washington.
Finally, we acknowledge financial help from six sources: the National Science
Foundation, the Alfred P. Sloan Foundation, the University of Washington, the
Science Research Council of the United Kingdom, the Sonderforschungsbereich of
Bonn, West Germany, and the Troisième Cycle of Bern, Switzerland.

1
2

4 A2
A1
5

6
7
A3
Preface to the second edition

The subject of BP -theory has grown dramatically since the appearance of the
first edition 17 years ago. One major development was the proof by Devinatz, Hop-
kins and Smith (see Devinatz, Hopkins and Smith [2] and Hopkins and Smith [3])
of nearly all the conjectures made in Ravenel [8]. An account of this work can be
found in our book Ravenel [13]. The only conjecture of Ravenel [8] that remains
is Telescope Conjecture. An account of our unsuccessful attempt to disprove it is
given in Mahowald, Ravenel, and Shick [1].
Another big development is the emergence of elliptic cohomology and the theory
of topological modular forms. There is still no comprehensive introduction to this
topic. Some good papers to start with are Ando, Hopkins and Strickland [1],
Hopkins and Mahowald [2], Landweber, Ravenel and Stong [8], and Rezk [1], which
is an account of the still unpublished Hopkins-Miller theorem.
The seventh and final chapter of the book has been completely rewritten and is
nearly twice as long as the original. We did this with an eye to carrying out future
research in this area.
I am grateful to the many would be readers who urged me to republish this
book and to the AMS for its assistance in getting the original manuscript retypeset.
Peter Landweber was kind enough to provide me with a copious list of misprints
he found in the first edition. Nori Minami and Igor Kriz helped in correcting some
errors in § 4.3. Mike Hill and his fellow MIT students provided me with a timely
list of typos in the online version of this edition. Hirofumi Nakai was very helpful
in motivating me to make the revisions of Chapter 7.

xvii
Commonly Used Notations

Z Integers
Zp p-agic integers
Z(p) Integers localized at p
Z/(p) Integers mod p
Q Rationals
Qp p-adic numbers
P (x) Polynomial algebra on generators x
E(x) Exterior algebra on generators x
2 Cotensor product (Section A1.1)
Given suitable objects A, B, and C and a map f : A → B, the evident map
A ⊗ C → B ⊗ C is denoted by f ⊗ C.

xix
CHAPTER 1

An Introduction to the Homotopy Groups


of Spheres

This chapter is intended to be an expository introduction to the rest of the book.


We will informally describe the spectral sequences of Adams and Novikov, which
are the subject of the remaining chapters. Our aim here is to give a conceptual
picture, suppressing as many technical details as possible.
In Section 1 we list some theorems which are classical in the sense that they
do not require any of the machinery described in this book. These include the
Hurewicz theorem 1.1.2, the Freudenthal suspension theorem 1.1.4, the Serre finite-
ness theorem 1.1.8, the Nishida nilpotence theorem 1.1.9, and the Cohen–Moore–
Neisendorfer exponent theorem 1.1.10. They all pertain directly to the homotopy
groups of spheres and are not treated elsewhere here. The homotopy groups of
the stable orthogonal group SO are given by the Bott periodicity theorem 1.1.11.
In 1.1.12 we define the J-homomorphism from πi (SO(n)) to πn+i (S n ). Its image
is given in 1.1.13, and in 1.1.14 we give its cokernel in low dimensions. Most of the
former is proved in Section 5.3.
In Section 2 we describe Serre’s method of computing homotopy groups using
cohomological techniques. In particular, we show how to find the first element of
order p in π∗ (S 3 ) 1.2.4. Then we explain how these methods were streamlined by
Adams to give his celebrated spectral sequence 1.2.10. The next four theorems
describe the Hopf invariant one problem. A table showing the Adams spectral
sequence at the prime 2 through dimension 45 is given in 1.2.15. In Chapter 2
we give a more detailed account of how the spectral sequence is set up, including
a convergence theorem. In Chapter 3 we make many calculations with it at the
prime 2.
In 1.2.16 we summarize Adams’ method for purposes of comparing it with
that of Novikov. The basic idea is to use complex cobordism (1.2.17) in place
of ordinary mod (p) cohomology. Fig. 1.2.19 is a table of the Adams–Novikov
spectral sequence for comparison with Fig. 1.2.15.
In the next two sections we describe the algebra surrounding the E2 -term of
the Adams–Novikov spectral sequence. To this end formal group laws are defined
in 1.3.1 and a complete account of the relevant theory is given in Appendix 2. Their
connection with complex cobordism is the subject of Quillen’s theorem (1.3.4) and
is described more fully in Section 4.1. The Adams–Novikov E2 -term is described in
terms of formal group law theory (1.3.5) and as an Ext group over a certain Hopf
algebra (1.3.6).
The rest of Section 3 is concerned with the Greek letter construction, a method
of producing infinite periodic families of elements in the E2 -term and (in favorable
cases) in the stable homotopy groups of spheres. The basic definitions are given in

1
2 1. INTRODUCTION TO THE HOMOTOPY GROUPS OF SPHERES

1.3.17 and 1.3.19 and the main algebraic fact required is the Morava–Landweber
theorem (1.3.16). Applications to homotopy are given in 1.3.11, 1.3.15, and 1.3.18.
The section ends with a discussion of the proofs and possible extensions of these
results. This material is discussed more fully in Chapter 5.
In Section 4 we describe the deeper algebraic properties of the E2 -term. We
start by introducing BP and defining a Hopf algebroid. The former is a minimal
wedge summand of M U localized at a prime. A Hopf algebroid is a generalized
Hopf algebra needed to describe the Adams–Novikov E2 -term more conveniently in
terms of BP (1.4.2). The algebraic and homological properties of such objects are
the subject of Appendix 1.
Next we give the Lazard classification theorem for formal group laws (1.4.3)
over an algebraically closed field of characteristic p, which is proved in Section A2.2.
Then we come to Morava’s point of view. Theorem 1.3.5 describes the Adams–
Novikov E2 -term as the cohomology of a certain group G with coefficients in a
certain polynomial ring L. Spec(L) (in the sense of abstract algebraic geometry)
is an infinite dimensional affine space on which G acts. The points in Spec(L) can
be thought of as formal group laws and the G-orbits as isomorphism classes, as
described in 1.4.3. This orbit structure is described in 1.4.4. For each orbit there
is a stabilizer or isotropy subgroup of G called Sn . Its cohomology is related to
that of G (1.4.5), and its structure is known. The theory of Morava stabilizer al-
gebras is the algebraic machinery needed to exploit this fact and is the subject of
Chapter 6. Our next topic, the chromatic spectral sequence (1.4.8, the subject of
Chapter 5), connects the theory above to the Adams–Novikov E2 -term. The Greek
letter construction fits into this apparatus very neatly.
Section 5 is about unstable homotopy groups of spheres and is not needed for
the rest of the book. Its introduction is self-explanatory.

1. Classical Theorems Old and New


We begin by recalling some definitions. The nth homotopy group of a connected
space X, πn (X), is the set of homotopy classes of maps from the n-sphere S n to X.
This set has a natural group structure which is abelian for n ≥ 2.
We now state three classical theorems about homotopy groups of spheres.
Proofs can be found, for example, in Spanier [1].
1.1.1. Theorem. π1 (S 1 ) = Z and πm (S 1 ) = 0 for m > 1. 
n n
1.1.2. Hurewicz’s Theorem. πn (S ) = Z and πm (S ) = 0 for m < n.
A generator of πn (S n ) is the class of the identity map. 
For the next theorem we need to define the suspension homomorphism
σ : πm (S n ) → πm+1 (S n+1 ).
1.1.3. Definition. The kth suspension Σk X of a space X is the quotient of
I × X obtained by collapsing ∂I k × X onto ∂I k , ∂I k being the boundary of I k ,
k

the k-dimensional cube. Note that Σi Σj X = Σi+j X and Σk f : Σk X → Σk Y is the


quotient of 1 × f : I k × X → I k × Y . In particular, given f : S m → S n we have
Σf : S m+1 → S n+1 , which induces a homomorphism πm (S n ) → πm+1 (S n+1 ). 
1.1.4. Freudenthal Suspension Theorem. The suspension homomorphism
σ : πn+k (S n ) → πn+k+1 (S n+q ) defined above is an isomorphism for k < n − 1 and
a surjection for k = n − 1. 
1. CLASSICAL THEOREMS OLD AND NEW 3

1.1.5. Corollary. The group πn+k (S n ) depends only on k if n > k + 1. 


1.1.6. Definition. The stable k-stem or kth stable homotopy group of spheres
πkS is πn+k (S n ) for n > k + 1. The groups πn+k (S n ) are called stable if n > k + 1
and unstable if n ≤ k + 1. When discussing stable groups we will not make any
notational distinction between a map and its suspensions. 
The subsequent chapters of this book will be concerned with machinery for
computing the stable homotopy groups of spheres. Most of the time we will not
be concerned with unstable groups. The groups πkS are known at least for k ≤ 45.
See the tables in Appendix 3, along with Theorem 1.1.13. Here is a table of πkS for
k ≤ 15:

k 0 1 2 3 4 5 6 7 8
πkS Z Z/(2) Z/(2) Z/(24) 0 0 Z/(2) Z/(240) (Z/(2))2

k 9 10 11 12 13 14 15
πkS (Z/2)3 Z/6 Z/(504) 0 Z/(3) (Z/(2))2 Z/(480) ⊕ Z/(2)

This should convince the reader that the groups do not fall into any obvious pattern.
Later in the book, however, we will present evidence of some deep patterns not
apparent in such a small amount of data. The nature of these patterns will be
discussed later in this chapter.
When homotopy groups were first defined by Hurewicz in 1935 it was hoped
that πn+k (S n ) = 0 for k > 0, since this was already known to be the case for n = 1
(1.1.1). The first counterexample is worth examining in some detail.
1.1.7. Example. π3 (S 2 ) = Z generated by the class of the Hopf map η : S 3 → S 2
defined as follows. Regard S 2 (as Riemann did) as the complex numbers C with a
point at infinity. S 3 is by definition the set of unit vectors in R4 = C2 . Hence a
point in S 3 is specified by two complex coordinates (z1 , z2 ). Define η by
(
z1 /z2 if z2 6= 0
η(z1 , z2 ) =
∞ if z2 = 0.
It is easy to verify that η is continuous. The inverse image under η of any point
in S 2 is a circle, specifically the set of unit vectors in a complex line through the
origin in C2 , the set of all such lines being parameterized by S 2 . Closer examination
will show that any two of these circles in S 3 are linked. One can use quaternions
and Cayley numbers in similar ways to obtain maps ν : S 7 → S 4 and σ : S 15 → S 8 ,
respectively. Both of these represent generators of infinite cyclic summands. These
three maps (η, ν, and σ) were all discovered by Hopf [1] and are therefore known
as the Hopf maps.
We will now state some other general theorems of more recent vintage.
1.1.8. Finiteness Theorem (Serre [3]). πn+k (S n ) is finite for k > 0 except
when n = 2m, k = 2m − 1, and π4m−1 (S 2m ) = Z ⊕ Fm , where Fm is finite. 
S S
L
The next theorem concerns the ring structure of π∗ = k≥0 πk which is in-
duced by composition as follows. Let α ∈ πiS and β ∈ πjS be represented by
f : S n+i → S n and g : S n+i+j → S n+i , respectively, where n is large. Then
4 1. INTRODUCTION TO THE HOMOTOPY GROUPS OF SPHERES

S
αβ ∈ πi+j is defined to be the class represented by f · g : S n+i+j → S n . It can be
shown that βα = (−1)ij αβ, so π∗S is an anticommutative graded ring.
1.1.9. Nilpotence Theorem (Nishida [1]). Each element α ∈ πkS for k > 0
is nilpotent, i.e., αt = 0 for some finite t. 
For the next result recall that 1.1.8 says π2i+1+j (S 2i+1 ) is a finite abelian group
for all j > 0.
1.1.10. Exponent Theorem (Cohen, Moore, and Neisendorfer [1]). For p ≥ 5
the p-component of π2i+1+j (S 2i+1 ) has exponent pi , i.e., each element in it has
order ≤ pi . 
This result is also true for p = 3 (Neisendorfer [1]) as well, but is known to be
false for p = 2. For example, the 2-component of 3-stem is cyclic of order 4 (see Fig.
3.3.18) on S 3 and of order 8 on S 8 (see Fig. 3.3.10). It is also known (Gray [1]) to
be the best possible, i.e., π2i+1+j (S 2i+1 ) is known to contain elements of order pi
for certain j.
We now describe an interesting subgroup of π∗S , the image of the Hopf–White-
head J-homomorphism, to be defined below. Let SO(n) be the space of n×n special
orthogonal matrices over RSwith the standard topology. SO(n) is a subspace of
SO(n + 1) and we denote n>0 SO(n) by SO, known as the stable orthogonal
group. It can be shown that πi (SO) = πi (SO(n)) if n > i + 1. The following result
of Bott is one of the most remarkable in all of topology.
1.1.11. Bott Periodicity Theorem (Bott [1]; see also Milnor [1]).

Z
 if i ≡ −1 mod 4
πi (SO) = Z/(2) if i = 0 or 1 mod 8 

0 otherwise.

We will now define a homomorphism J : πi (SO(n)) → πn+i (S n ). Let α ∈


πi (SO(n)) be the class of f : S i → SO(n). Let Dn be the n-dimensional disc, i.e.,
the unit ball in Rn . A matrix in SO(n) defines a linear homeomorphism of Dn to
itself. We define fˆ: S i ×Dn → Dn by fˆ(x, y) = f (x)(y), where x ∈ S i , y ∈ Dn , and
f (x) ∈ SO(n). Next observe that S n is the quotient of Dn obtained by collapsing
its boundary S n−1 to a single point, so there is a map p : Dn → S n , which sends
the boundary to the base point. Also observe that S n+i , being homeomorphic to
the boundary of Di+1 × Dn , is the union of S i × Dn and Di+1 × S n−1 along their
common boundary S i × S n−1 . We define f˜: S n+i → S n to be the extension of
pfˆ: S i × Dn → S n to S n+i which sends the rest of S n+i to the base point in S n .
1.1.12. Definition. The Hopf–Whitehead J-homomorphism J : πi (SO(n)) →
πn+i (S n ) sends the class of f : S i → SO(n) to the class of f˜: S n+i → S n as
described above. 
We leave it to the skeptical reader to verify that the above construction actually
gives us a homomorphism.
Note that both πi (SO(n)) and πn+i (S n ) are stable, i.e., independent of n, if
n > i + 1. Hence we have J : πk (SO) → πkS . We will now describe its image.
1.1.13. Theorem (Adams [1] and Quillen [1]). J : πk (SO) → πkS is a monomor-
phism for k ≡ 0 or 1 mod 8 and J(π4k−1 (SO)) is a cyclic group whose 2-component
2. METHODS OF COMPUTING π∗ (S n ) 5

is Z(2) /(8k) and whose p-component for p ≥ 3 is Z(p) /(pk) if (p − 1) | 2k and 0 if


(p − 1) - 2k, where Z(p) denotes the integers localized at p. In dimensions 1, 3, and
7, im J is generated by the Hopf maps (1.1.7) η, ν, and σ, respectively. If we denote
by xk the generator in dimension 4k − 1, then ηx2k and η 2 x2k are the generators
of im J in dimensions 8k and 8k + 1, respectively. 
The image of J is also known to a direct summand; a proof can be found for
example at the end of Chapter 19 of Switzer [1]. The order of J(π4k−1 (SO)) was
determined by Adams up to a factor of two, and he showed that the remaining
ambiguity could be resolved by proving the celebrated Adams conjecture, which
Quillen and others did. Denote this number by ak . Its first few values are tabulated
here.
k 1 2 3 4 5 6 7 8 9 10
ak 24 240 504 480 264 65,520 24 16,320 28,728 13,200
The number ak has interesting number theoretic properties. It is the denominator
of Bk /4k, where Bk , is the kth Bernoulli number, and it is the greatest common
divisor of numbers nt(n) (n2k −1) for n ∈ Z and t(n) sufficiently large. See Adams [1]
and Milnor and Stasheff [5] for details.
Having determined im J, one would like to know something systematic about
coker J, i.e., something more than its structure through a finite range of dimensions.
For the reader’s amusement we record some of that structure now.
1.1.14. Theorem. In dimensions ≤ 15, the 2-component of coker J has the
following generators, each with order 2:
η 2 ∈ π2S , ν 2 ∈ π6S , ν̄ ∈ π8S , ην̄ = ν 3 ∈ π9S , µ ∈ π9S ,
S
ηµ ∈ π10 , σ 2 ∈ π14
S
, S
κ ∈ π14 and S
ηκ ∈ π15 .
(There are relations η 3 = 4ν and η 2 µ = 4x3 ). For p ≥ 3 the p-component of coker J
has the following generators in dimensions ≤ 3pq − 6 (where q = 2p − 2), each with
order p:
S S
β1 ∈ πpq−2 , α1 β1 ∈ π(p+1)q−3
S
where α1 = x(p−1)/2 ∈ πq−1 is the first generator of the p-component of im J,
β12 ∈ π2pq−4
S
, α1 β12 ∈ π(2p+1)q−5
S
, S
β2 ∈ π(2p+1)q−2 ,
S
α1 β2 ∈ π(2p+2)q−3 , and β13 ∈ π3pq−6
S
. 
The proof and the definitions of new elements listed above will be given later
in the book, e.g., in Section 4.4.

2. Methods of Computing π∗ (S n )
In this section we will informally discuss three methods of computing homotopy
groups of spheres, the spectral sequences of Serre, Adams, and Novikov. A fourth
method, the EHP sequence, will be discussed in Section 5. We will not give any
proofs and in some cases we will sacrifice precision for conceptual clarity, e.g., in
our identification of the E2 -term of the Adams–Novikov spectral sequence.
The Serre spectral sequence (circa 1951) (Serre [2]) is included here mainly
for historical interest. It was the first systematic method of computing homotopy
groups and was a major computational breakthrough. It has been used as late as
the 1970s by various authors (Toda [1], Oka [1, 2, 3]), but computations made
6 1. INTRODUCTION TO THE HOMOTOPY GROUPS OF SPHERES

with it were greatly clarified by the introduction of the Adams spectral sequence
in 1958 in Adams [3]. In the Adams spectral sequence the basic mechanism of the
Serre spectral sequence information is organized by homological algebra.
For the 2-component of π∗ (S n ) the Adams spectral sequence is indispensable to
this day, but the odd primary calculations were streamlined by the introduction of
the Adams–Novikov spectral sequence (Adams–Novikov spectral sequence) in 1967
by Novikov [1]. It is the main subject in this book. Its E2 -term contains more
information than that of the Adams spectral sequence; i.e., it is a more accurate
approximation of stable homotopy and there are fewer differentials in the spectral
sequence. Moreover, it has a very rich algebraic structure, as we shall see, largely
due to the theorem of Quillen [2], which establishes a deep (and still not satisfac-
torily explained) connection between complex cobordism (the cohomology theory
used to define the Adams–Novikov spectral sequence; see below) and the theory of
formal group laws. Every major advance in the subject since 1969, especially the
work of Jack Morava, has exploited this connection.
We will now describe these three methods in more detail. The starting point
for Serre’s method is the following classical result.
1.2.1. Theorem. Let X be a simply connected space with Hi (X) = 0 for i < n
for some positive integer n ≥ 2. Then
(a) (Hurewicz [1]). πn (X) = Hn (X).
(b) (Eilenberg and Mac Lane [2]). There is a space K(π, n), characterized up
to homotopy equivalence by
(
π if i = n
πi (K(π, n)) =
0 if i 6= n.
If X is above and π = πn (X) then there is a map f : X → K(π, n) such that Hn (f )
and πn (f ) are isomorphisms. 
1.2.2. Corollary. Let F be the fiber of the map f above. Then
(
πi (X) for i ≥ n + 1
πi (F ) = 
0 for i ≤ n.
In other words, F has the same homotopy groups as X in dimensions above
n, so computing π∗ (F ) is as good as computing π∗ (X). Moreover, H∗ (K(π, n)) is
known, so H∗ (F ) can be computed with the Serre spectral sequence applied to the
fibration F → X → K(π, n).
Once this has been done the entire process can be repeated: let n0 > n be the
dimension of the first nontrivial homology group of F and let Hn0 (F ) = π 0 . Then
πn0 (F ) = πn0 (X) = π 0 is the next nontrivial homotopy group of X. Theorem 1.2.1
applied to F gives a map f 0 : F → K(π 0 , n0 ) with fiber F 0 , and 1.2.2 says
(
πi (X) for i > n0
πi (F 0 ) =
0 for i ≤ n0 .
Then one computes H∗ (F 0 ) using the Serre spectral sequence and repeats the pro-
cess.
As long as one can compute the homology of the fiber at each stage, one can
compute the next homotopy group of X. In Serre [3] a theory was developed
which allows one to ignore torsion of order prime to a fixed prime p throughout the
2. METHODS OF COMPUTING π∗ (S n ) 7

calculation if one is only interested in the p-component of π∗ (X). For example, if


X = S 3 , one uses 1.2.1 to get a map to K(Z, 3). Then H∗ (F ) is described by:
1.2.3. Lemma. If F is the fibre of the map f : S 3 → K(Z, 3) given by 1.2.1,
then (
Z/(m) if i = 2m and m > 1
Hi (F ) = 
0 otherwise.

1.2.4. Corollary. The first p-torsion in π∗ (S 3 ) is Z/(p) in π2p (S 3 ) for any


prime p. 
Proof of 1.2.3. (It is so easy we cannot resist giving it.) We have a fibration
ΩK(Z, 3) = K(Z, 2) → F → S 3
and H ∗ (K(Z, 2)) = H ∗ (CP ∞ ) = Z[x], where x ∈ H 2 (CP ∞ ) and CP ∞ is an
infinite-dimensional complex projective space. We will look at the Serre spectral
sequence for H ∗ (F ) and use the universal coefficient theorem to translate this to the
desired description of H∗ (F ). Let u be the generator of H 3 (S 3 ). Then in the Serre
spectral sequence we must have d3 (x) = ±u; otherwise F would not be 3-connected,
contradicting 1.1.2. Since d3 is a derivation we have d3 (xn ) = ±nuxn−1 . It is easily
seen that there can be no more differentials and we get
(
Z/(m) if i = 2m + 1, m > 1
H i (F ) =
0 otherwise
which leads to the desired result. 
If we start with X = S n the Serre spectral sequence calculations will be much
easier for πk+n (S n ) for k < n − 1. Then all of the computations are in the stable
range, i.e., in dimensions less than twice the connectivity of the spaces involved.
i f
This means that for a fibration F − →X− → K, the Serre spectral sequence gives a
long exact sequence
i ∗ f∗ d
(1.2.5) · · · → Hj (F ) −→ Hj (X) −−→ Hj (K) −
→ Hj−1 (F ) → · · · ,
where d corresponds to Serre spectral sequence differentials. Even if we know
H∗ (X), H∗ (K), and f∗ , we still have to deal with the short exact sequence
(1.2.6) 0 → coker f∗ → H∗ (F ) → ker f∗ → 0.
It may lead to some ambiguity in H∗ (F ), which must be resolved by some other
means. For example, when computing π∗ (S n ) for large n one encounters this prob-
lem in the 3-component of πn+10 (S n ) and the 2-component of πn+14 (S n ). This
difficulty is also present in the Adams spectral sequence, where one has the pos-
sibility of a nontrivial differential in these dimensions. These differentials were
first calculated by Adams [12], Liulevicius [2], and Shimada and Yamanoshita [3]
by methods involving secondary cohomology operations and later by Adams and
Atiyah [13] by methods involving K-theory
The Adams spectral sequence of Adams [3] begins with a variation of Serre’s
method. One works only in the stable range and only on Q the p-component. Instead
of mapping X to K(π, n) as in 1.2.1, one maps to K = j>0 K(H j (X; Z/(p)), j) by
a certain map g which induces a surjection in mod (p) cohomology. Let X1 be the
fiber of g. Define spaces Xi and Ki inductively by Ki = j>0 K(H j (Xi ; Z/(p)), j)
Q
8 1. INTRODUCTION TO THE HOMOTOPY GROUPS OF SPHERES

and Xi+1 is the fiber of g : Xi → Ki (this map is defined in Section 2.1, where the
Adams spectral sequence is discussed in more detail). Since H ∗ (gi ) is onto, the
analog of 1.2.5 is an short exact sequence in the stable range
(1.2.7) 0 ← H ∗ (Xi ) ← H ∗ (Ki ) ← H ∗ (ΣXi+1 ) ← 0,
where all cohomology groups are understood to have coefficients Z/(p). Moreover,
H ∗ (Ki ) is a free module over the mod (p) Steenrod algebra A, so if we splice
together the short exact sequences of 1.2.7 we get a free A-resolution of H ∗ (X)
(1.2.8) 0 ← H ∗ (X) ← H ∗ (K) ← H ∗ (Σ1 K1 ) ← H ∗ (Σ2 K2 ) ← · · ·
Each of the fibration Xi+1 → Xi → Ki gives a long exact sequence of homotopy
groups. Together these long exact sequences form an exact couple and the asso-
ciated spectral sequence is the Adams spectral sequence for the p-component of
π∗ (X). If X has finite type, the diagram
(1.2.9) K → Σ−1 K1 → Σ−2 K2 → · · ·
(which gives 1.2.8 in cohomology) gives a cochain complex of homotopy groups
whose cohomology is ExtA (H ∗ (X); Z/(p)). Hence one gets
1.2.10. Theorem (Adams [3]). There is a spectral sequence converging to the
p-component of πn+k (S n ) for k < n − 1 with
E2s,t = Exts,t s,t
A (Z/(p), Z/(p)) =: H (A)
s,t
and dr : Ers,t → Ers+r,t+r−1 . Here the groups E∞ for t − s = k form the associated
graded group to a filtration of the p-component of πn+k (S n ). 
Computing this E2 -term is hard work, but it is much easier than making similar
computations with Serre spectral sequence. The most widely used method today is
the spectral sequence of May [1, 2] (see Section 3.2). This is a trigraded spectral
sequence converging to H ∗∗ (A), whose E2 -term is the cohomology of a filtered form
of the Steenrod algebra. This method was used by Tangora [1] to compute E2s,t
for p = 2 and t − s ≤ 70. Most of his table is reproduced here in Fig. A3.1a–c.
Computations for odd primes can be found in Nakamura [2].
As noted above, the Adams E2 -term is the cohomology of the Steenrod algebra.
Hence E21,∗ = H 1 (A) is the indecomposables in A. For p = 2 one knows that A
i
is generated by Sq 2 for i ≥ 0; the corresponding elements in E21,∗ are denoted by
i i
hi ∈ E21,2 . For p > 2 the generators are the Bockstein β and P p for i ≥ 0 and the
i
corresponding elements are a0 ∈ E21,1 and hi ∈ E21,qp , where q = 2p − 2.
For p = 2 these elements figure in the famous Hopf invariant one problem.
1.2.11. Theorem (Adams [12]). The following statements are equivalent.
i
(a) S 2 −1 is parallelizable, i.e., it has 2i −1 globally linearly independent tangent
vector fields.
(b) There is a division algebra (not necessarily associative) over R of dimen-
sion 2i .
i i
(c) There is a map S 2·2 −1 → S 2 of Hopf invariant one (see 1.5.2).
i i+1
(d) There is a 2-cell complex X = S 2 ∪ e2 [the cofiber of the map in (c)] in
i+1 i
which the generator of H 2 (X) is the square of the generator of H 2 (X).
i
(e) The element hi ∈ E21,2 is a permanent cycle in the Adams spectral sequence.

2. METHODS OF COMPUTING π∗ (S n ) 9

Condition (b) is clearly true for i = 0, 1, 2 and 3, the division algebras being
the reals R, the complexes C, the quaternions H and the Cayley numbers, which
are nonassotiative. The problem for i ≥ 4 is solved by
1.2.12. Theorem (Adams [12]). The conditions of 1.2.11 are false for i ≥ 4
and in the Adams spectral sequence one has d2 (hi ) = h0 h2i−1 6= 0 for i ≥ 4. 
For i = 4 the above gives the first nontrivial differential in the Adams spectral
sequence. Its target has dimension 14 and is related to the difficulty in Serre’s
method referred to above.
The analogous results for p > 2 are
1.2.13. Theorem (Liulevicius [2] and Shimada and Yamanoshita [3]). The
following are equivalent.
i+1 i
(a) There is a map S 2p −1 → Sb2p with Hopf invariant one (see 1.5.3 for the
definition of the Hopf invariant and the space Sb2m ).
i i i i+1
(b) There is a p-cell complex X = S 2p ∪ e4p ∪ e6p ∪ · · · ∪ e2p [the cofiber
of the map in (a)] whose mod (p) cohomology is a truncated polynomial algebra on
one generator.
i
(c) The element hi ∈ E21,qp is a permanent cycle in the Adams spectral se-
quence. 
The element h0 is the first element in the Adams spectral sequence above
dimension zero so it is a permanent cycle. The corresponding map in (a) suspends
to the element of π2p (S 3 ) given by 1.2.4. For i ≥ 1 we have
1.2.14. Theorem (Liulevicius [2] and Shimada and Yamanoshita [3]). The
conditions of 1.2.13 are false for i ≥ 1 and d2 (hi ) = a0 bi−1 , where bi−1 is a
i
generator of E22,qp (see Section 5.2). 
For i = 1 the above gives the first nontrivial differential in the Adams spectral
sequence for p > 2. For p = 3 its target is in dimension 10 and was referred to
above in our discussion of Serre’s method.
Fig. 1.2.15 shows the Adams spectral sequence for p = 3 through dimension
45. We present it here mainly for comparison with a similar figure (1.2.19) for the
Adams–Novikov spectral sequence. E2s,t is a Z/(p) vector space in which each basis
element is indicated by a small circle. Fortunately in this range there are just two
bigradings [(5,28) and (8,43)] in which there is more than one basis element. The
vertical coordinate is s, the cohomological degree, and the horizontal coordinate
is t − s, the topological dimension. These extra elements appear in the chart to
the right of where they should be, and the lines meeting them should be vertical.
A dr is indicated by a line which goes up by r and to the left by 1. The vertical
lines represent multiplication by a0 ∈ E21,1 and the vertical arrow in dimension
zero indicates that all powers of a0 are nonzero. This multiplication corresponds to
multiplication by p in the corresponding homotopy group. Thus from the figure one
can read off π0 = Z, π11 = π45 = Z/(9), π23 = Z/(9) ⊕ Z/(3), and π35 = Z/(27).
Lines that go up 1 and to the right by 3 indicate multiplication by h0 ∈ E21,4 ,
while those that go to the right by 7 indicate the Massey product hh0 , h0 , −i (see
A1.4.1). The elements a0 and hi for i = 0, 1, 2 were defined above and the elements
b0 ∈ E22,12 , k0 ∈ E22,28 , and b1 ∈ E22,36 are up to the sign the Massey products
hh0 , h0 , h0 i, hh0 , h1 , h1 i, and hh1 , h1 , h1 i, respectively. The unlabeled elements in
Another random document with
no related content on Scribd:
nucleus with planetary negative electrons, and of the picture that
Niels Bohr has imagined by working these in with the “quanta” of
Planck.[479] The atoms of Leucippus and Democritus were different in
form and magnitude, that is to say, they were purely plastic units,
“indivisible,” as their name asserts, but only plastically indivisible.
The atoms of Western physics, for which “indivisibility” has quite
another meaning, resemble the figures and themes of music; their
being or essence consisting in vibration and radiation, and their
relation to the processes of Nature being that of the “motive” to the
“movement.”[480] Classical physics examines the aspect, Western the
working, of these ultimate elements in the picture of the Become; in
the one, the basic notions are notions of stuff and form, in the other
they are notions of capacity and intensity.
There is a Stoicism and there is a Socialism of the atom, the
words describing the static-plastic and the dynamic-contrapuntal
ideas of it respectively. The relations of these ideas to the images of
the corresponding ethics is such that every law and every definition
takes these into account. On the one hand—Democritus’s multitude
of confused atoms, put there, patient, knocked about by the blind
chance that he as well as Sophocles called ἀνάγκη, hunted like
Œdipus. On the other hand—systems of abstract force-points
working in unison, aggressive, energetically dominating space (as
“field”), overcoming resistances like Macbeth. The opposition of
basic feelings makes that of the mechanical Nature-pictures.
According to Leucippus the atoms fly about in the void “of
themselves”; Democritus merely regards shock and countershock as
a form of change of place. Aristotle explains individual movements
as accidental, Empedocles speaks of love and hate, Anaxagoras of
meetings and partings. All these are elements also of Classical
tragedy; the figures on the Attic stage are related to one another just
so. Further, and logically, they are the elements of Classical politics.
There we have minute cities, political atoms ranged along coasts
and on islands, each jealously standing for itself, yet ever needing
support, shut-in and shy to the point of absurdity, buffeted hither and
thither by the planless orderless happenings of Classical history,
rising to-day and ruined to-morrow. And in contrast—the dynastic
states of our 17th and 18th Centuries, political fields of force, with
cabinets and great diplomats as effective centres of purposeful
direction and comprehensive vision. The spirit of Classical history
and the spirit of Western history can only be really understood by
considering the two souls as an opposition. And we can say the
same of the atom-idea, regarded as the basis of the respective
physics. Galileo who created the concept of force and the Milesians
who created that of ἀρχή, Democritus and Leibniz, Archimedes and
Helmholtz, are “contemporaries,” members of the same intellectual
phases of quite different Cultures.
But the inner relationship between atom-theory and ethic goes
further. It has been shown how the Faustian soul—whose being
consists in the overcoming of presence, whose feeling is loneliness
and whose yearning is infinity—puts its need of solitude, distance
and abstraction into all its actualities, into its public life, its spiritual
and its artistic form-worlds alike. This pathos of distance (to use
Nietzsche’s expression) is peculiarly alien to the Classical, in which
everything human demanded nearness, support and community. It is
this that distinguishes the spirit of the Baroque from that of the Ionic,
the culture of the Ancien Régime from that of Periclean Athens. And
this pathos, which distinguishes the heroic doer from the heroic
sufferer, appears also in the picture of Western physics as tension. It
is tension that is missing in the science of Democritus; for in the
principle of shock and countershock it is denied by implication that
there is a force commanding space and identical with space. And,
correspondingly, the element of Will is absent from the Classical
soul-image. Between Classical men, or states, or views of the world,
there was—for all the quarrelling and envy and hatred—no inner
tension, no deep and urging need of distance, solitude, ascendancy;
and consequently there was none between the atoms of the Cosmos
either. The principle of tension (developed in the potential theory),
which is wholly untranslatable into Classical tongues and
incommunicable to Classical minds, has become for Western
physics fundamental. Its content follows from the notion of energy,
the Will-to-Power in Nature, and therefore it is for us just as
necessary as for the Classical thought it is impossible.

IV
Every atomic theory, therefore, is a myth and not an experience. In
it the Culture, through the contemplative-creative power of its great
physicists, reveals its inmost essence and very self. It is only a
preconceived idea of criticism that extension exists in itself and
independently of the form-feeling and world-feeling of the knower.
The thinker, in imagining that he can cut out the factor of Life, forgets
that knowing is related to the known as direction is to extension and
that it is only through the living quality of direction that what is felt
extends into distance and depth and becomes space. The cognized
structure of the extended is a projection of the cognizing being.
We have already[481] shown the decisive importance of the depth-
experience, which is identical with the awakening of a soul and
therefore with the creation of the outer world belonging to that soul.
The mere sense-impression contains only length and breadth, and it
is the living and necessary act of interpretation—which, like
everything else living, possesses direction, motion and irreversibility
(the qualities that our consciousness synthesizes in the word Time)
—that adds depth and thereby fashions actuality and world. Life itself
enters into the experiences as third dimension. The double meaning
of the word “far,” which refers both to future and to horizon, betrays
the deeper meaning of this dimension, through which extension as
such is evoked. The Becoming stiffens and passes and is at once
the Become; Life stiffens and passes and is at once the three-
dimensional Space of the known. It is common ground for Descartes
and Parmenides that thinking and being, i.e., imagined and
extended, are identical. “Cogito, ergo sum” is simply the formulation
of the depth-experience—I cognize, and therefore I am in space. But
in the style of this cognizing, and therefore of the cognition-product,
the prime-symbol of the particular Culture comes into play. The
perfected extension of the Classical consciousness is one of
sensuous and bodily presence. The Western consciousness
achieves extension, after its own fashion, as transcendental space,
and as it thinks its space more and more transcendentally it
develops by degrees the abstract polarity of Capacity and Intensity
that so completely contrasts with the Classical visual polarity of
Matter and Form.
But it follows from this that in the known there can be no
reappearance of living time. For this has already passed into the
known, into constant “existence,” as Depth, and hence duration (i.e.,
timelessness) and extension are identical. Only the knowing
possesses the mark of direction. The application of the word “time”
to the imaginary and measurable time-dimension of physics is a
mistake. The only question is whether it is possible or not to avoid
the mistake. If one substitutes the word “Destiny” for “time” in any
physical enunciation, one feels at once that pure Nature does not
contain Time. The form-world of physics extends just as far as the
cognate form-world of number and notion extend, and we have seen
that (notwithstanding Kant) there is not and cannot be the slightest
relation of any sort between mathematical number and Time. And yet
this is controverted by the fact of motion in the picture of the world-
around. It is the unsolved and unsolvable problem of the Eleatics—
being (or thinking) and motion are incompatible; motion “is” not (is
only “apparent”).
And here, for the second time, Natural science becomes dogmatic
and mythological. The words Time and Destiny, for anyone who uses
them instinctively, touch Life itself in its deepest depths—life as a
whole, which is not to be separated from lived-experience. Physics,
on the other hand—i.e., the observing Reason—must separate
them. The livingly-experienced “in-itself,” mentally emancipated from
the act of the observer and become object, dead, inorganic, rigid, is
now “Nature,” something open to exhaustive mathematical
treatment. In this sense the knowledge of Nature is an activity of
measurement. All the same, we live even when we are observing
and therefore the thing we are observing lives with us. The element
in the Nature-picture in virtue of which it not merely from moment to
moment is, but in a continuous flow with and around us becomes, is
the copula of the waking-consciousness and its world. This element
is called movement, and it contradicts Nature as a picture, but it
represents the history of this picture. And therefore, as precisely as
Understanding is abstracted (by means of words) from feeling and
mathematical space from light-resistances (“things”[482]), so also
physical “time” is abstracted from the impression of motion.
Physics investigates Nature, and consequently it knows time only
as a length. But the physicist lives in the midst of the history of this
Nature, and therefore he is forced to conceive motion as a
mathematically determinable magnitude, as a concretion of the pure
numbers obtained in the experiment and written down in formulæ.
“Physics,” says Kirchhoff, “is the complete and simple description of
motions.” That indeed has always been its object. But the question is
one not of motions in the picture but of motions of the picture.
Motion, in the Nature of physics, is nothing else but that
metaphysical something which gives rise to the consciousness of a
succession. The known is timeless and alien to motion; its state of
becomeness implies this. It is the organic sequence of knowns that
gives the impression of a motion. The physicist receives the word as
an impression not upon “reason” but upon the whole man, and the
function of that man is not “Nature” only but the whole world. And
that is the world-as-history. “Nature,” then, is an expression of the
Culture in each instance.[483] All physics is treatment of the motion-
problem—in which the life-problem itself is implicit—not as though it
could one day be solved, but in spite of, nay because of, the fact that
it is insoluble. The secret of motion awakens in man the
apprehension of death.[484]
If, then, Nature-knowledge is a subtle kind of self-knowledge—
Nature understood as picture, as mirror of man—the attempt to solve
the motion-problem is an attempt of knowledge to get on the track of
its own secret, its own Destiny.

Only physiognomic tact can, if creative, succeed in this, and in fact


it has done so from time immemorial in the arts, particularly tragic
poetry. It is the thinking man who is perplexed by movement; for the
contemplative it is self-evident. And however completely the former
can reduce his perplexities to system, the result is systematic and
not physiognomic, pure extension logically and numerically ordered,
nothing living but something become and dead.
It is this that led Goethe, who was a poet and not a computer, to
observe that “Nature has no system. It has Life, it is Life and
succession from an unknown centre to an unknowable bourne.” For
one who does not live it but knows it, Nature has a system. But it is
only a system and nothing more, and motion is a contradiction in it.
The contradiction may be covered up by adroit formulation, but it
lives on in the fundamental concepts. The shock and countershock
of Democritus, the entelechy of Aristotle, the notions of force from
the “impetus” of 14th-Century Occamists to the quantum-theory of
radiation, all contain it. Let the reader conceive of the motion within a
physical system as the ageing of that system (as in fact it is, as lived-
experience of the observer), and he will feel at once and distinctly
the fatefulness immanent in, the unconquerably organic content of,
the word “motion” and all its derivative ideas. But Mechanics, having
nothing to do with ageing, should have nothing to do with motion
either, and consequently, since no scientific system is conceivable
without a motion-problem in it, a complete and self-contained
mechanics is an impossibility. Somewhere or other there is always
an organic starting-point in the system where immediate Life enters it
—an umbilical cord that connects the mind-child with the life-mother,
the thought with the thinker.
This puts the fundamentals of Faustian and Apollinian Nature-
science in quite another light. No “Nature” is pure—there is always
something of history in it. If the man is ahistorical, like the Greek, so
that the totality of his impressions of the world is absorbed in a pure
point-formed present, his Nature-image is static, self-contained (that
is, walled against past and future) in every individual moment. Time
as magnitude figures in Greek physics as little as it does in
Aristotle’s entelechy-idea. If, on the other hand, the Man is
historically constituted, the image formed is dynamic. Number, the
definitive evaluation of the become, is in the case of ahistoric man
Measure, and in that of the historical man Function. One measures
only what is present and one follows up only what has a past and a
future, a course. And the effect of this difference is that the inner
inconsistencies of the motion-problem are covered up in Classical
theories and forced into the foreground in Western.
History is eternal becoming and therefore eternal future; Nature is
become and therefore eternally past.[485] And here a strange
inversion seems to have taken place—the Becoming has lost its
priority over the Become. When the intellect looks back from its
sphere, the Become, the aspect of life is reversed, the idea of
Destiny which carries aim and future in it having turned into the
mechanical principle of cause-and-effect of which the centre of
gravity lies in the past. The spatially-experienced is promoted to rank
above the temporal living, and time is replaced by a length in a
spatial world-system. And, since in the creative experience extension
follows from direction, the spatial from life, the human understanding
imports life as a process into the inorganic space of its imagination.
While life looks on space as something functionally belonging to
itself, intellect looks upon life as something in space. Destiny asks:
“Whither?”, Causality asks: “Whence?” To establish scientifically
means, starting from the become and actualized, to search for
“causes” by going back along a mechanically-conceived course, that
is to say, by treating becoming as a length. But it is not possible to
live backwards, only to think backwards. Not Time and Destiny are
reversible, but only that which the physicist calls “time” and admits
into his formulæ as divisible, and preferably as negative or imaginary
quantities.
The perplexity is always there, though it has rarely been seen to
be originally and necessarily inherent. In the Classical science the
Eleatics, declining to admit the necessity of thinking of Nature as in
motion, set up against it the logical view that thinking is a being, with
the corollary that known and extended are identical and knowledge
and becoming therefore irreconcilable. Their criticisms have not
been, and cannot be, refuted. But they did not hinder the evolution of
Classical physics, which was a necessary expression of the
Apollinian soul and as such superior to logical difficulties. In the
“classical” mechanics so-called of the Baroque, founded by Galileo
and Newton, an irreproachable solution of the motion-problem on
dynamic lines has been sought again and again. The history of the
concept of force, which has been stated and restated with all the
tireless passion of a thought that feels its own self endangered by a
difficulty, is nothing but the history of endeavours to find a form that
is unimpeachable, mathematically and conceptually, for motion. The
last serious attempt—which failed like the rest, and of necessity—
was Hertz’s.
Without discovering the true source of all perplexities (no physicist
as yet has done that), Hertz tried to eliminate the notion of force
entirely—rightly feeling that error in all mechanical systems has to be
looked for in one or another of the basic concepts—and to build up
the whole picture of physics on the quantities of time, space and
mass. But he did not observe that it is Time itself (which as direction-
factor is present in the force-concept) that is the organic element
without which a dynamic theory cannot be expressed and with which
a clean solution cannot be got. Moreover, quite apart from this, the
concepts force, mass and motion constitute a dogmatic unit. They so
condition one another that the application of one of them tacitly
involves both the others from the outset. The whole Apollinian
conception of the motion-problem is implicit in the root-word ἀρχή,
the whole Western conception of it in the force-idea. The notion of
mass is only the complement of that of force. Newton, a deeply
religious nature, was only bringing the Faustian world-feeling to
expression when, to elucidate the words “force” and “motion,” he
said that masses are points of attack for force and carriers for
motion. So the 13th-Century Mystics had conceived of God and his
relation to world. Newton no doubt rejected the metaphysical
element in his famous saying “hypotheses non fingo,” but all the
same he was metaphysical through and through in the founding of
his mechanics. Force is the mechanical Nature-picture of western
man; what Will is to his soul-picture and infinite Godhead in his
world-picture. The primary ideas of this physics stood firm long
before the first physicist was born, for they lay in the earliest religious
world-consciousness of our Culture.

VI

With this it becomes manifest that the physical notion of Necessity,


too, has a religious origin. It must not be forgotten that the
mechanical necessity that rules in what our intellects comprehend as
Nature is founded upon another necessity which is organic and
fateful in Life itself. The latter creates, the former restricts. One
follows from inward certitude, the other from demonstration; that is
the distinction between tragic and technical, historical and physical
logic.
There are, further, differences within the necessity postulated and
assumed by science (that of cause-and-effect) which have so far
eluded the keenest sight. We are confronted here with a question at
once of very great difficulty and of superlative importance. A Nature-
knowledge is (however philosophy may express the relation) a
function of knowing, which is in each case knowing in a particular
style. A scientific necessity therefore has the style of the appropriate
intellect, and this brings morphological differences into the field at
once. It is possible to see a strict necessity in Nature even where it
may be impossible to express it in natural laws. In fact natural laws,
which for us are self-evidently the proper expression-form in science,
are not by any means so for the men of other Cultures. They
presuppose a quite special form, the distinctively Faustian form, of
understanding and therefore of Nature-knowing. There is nothing
inherently absurd in the conception of a mechanical necessity
wherein each individual case is morphologically self-contained and
never exactly reproduced, in which therefore the acquisitions of
knowledge cannot be put into consistently-valid formulæ. In such a
case Nature would appear (to put it metaphorically) as an unending
decimal that was also non-recurring, destitute of periodicity. And so,
undoubtedly, it was conceived by Classical minds—the feeling of it
manifestly underlies their primary physical concepts. For example,
the proper motion of Democritus’s atoms is such as to exclude any
possibility of calculating motions in advance.
Nature-laws are forms of the known in which an aggregate of
individual cases are brought together as a unit of higher degree.
Living Time is ignored—that is, it does not matter whether, when or
how often the case arises, for the question is not of chronological
sequence but of mathematical consequence.[486] But in the
consciousness that no power in the world can shake this calculation
lies our will to command over Nature. That is Faustian. It is only from
this standpoint that miracles appear as breaches of the laws of
Nature. Magian man saw in them merely the exercise of a power that
was not common to all, not in any way a contradiction of the laws of
Nature. And Classical man, according to Protagoras, was only the
measure and not the creator of things—a view that unconsciously
forgoes all conquest of Nature through the discovery and application
of laws.
We see, then, that the causality-principle, in the form in which it is
self-evidently necessary for us—the agreed basis of truth for our
mathematics, physics and philosophy—is a Western and, more
strictly speaking, a Baroque phenomenon. It cannot be proved, for
every proof set forth in a Western language and every experiment
conducted by a Western mind presupposes itself. In every problem,
the enunciation contains the proof in germ. The method of a science
is the science itself. Beyond question, the notion of laws of Nature
and the conception of physics as “scientia experimentalis,”[487] which
has held ever since Roger Bacon, contains a priori this specific kind
of necessity. The Classical mode of regarding Nature—the alter ego
of the Classical mode of being—on the contrary, does not contain it,
and yet it does not appear that the scientific position is weakened in
logic thereby. If we work carefully through the utterances of
Democritus, Anaxagoras, and Aristotle (in whom is contained the
whole sum of Classical Nature-speculation), and, above all, if we
examine the connotations of key-terms like ἀλλοίωσις, ἀνάγκη,
ἐντελέχεια, we look with astonishment into a world-image totally
unlike our own. This world-image is self-sufficing and therefore, for
this definite sort of mankind, unconditionally true. And causality in
our sense plays no part therein.
The alchemist or philosopher of the Arabian Culture, too, assumes
a necessity within his world-cavern that is utterly and completely
different from the necessity of dynamics. There is no causal nexus of
law-form but only one cause, God, immediately underlying every
effect. To believe in Nature-laws would, from this standpoint, be to
doubt the almightiness of God. If a rule seems to emerge, it is
because it pleases God so; but to suppose that this rule was a
necessity would be to yield to a temptation of the Devil. This was the
attitude also of Carneades, Plotinus and the Neo-Pythagoreans.[488]
This necessity underlies the Gospels as it does the Talmud and the
Avesta, and upon it rests the technique of alchemy.
The conception of number as function is related to the dynamic
principle of cause-and-effect. Both are creations of the same
intellect, expression-forms of the same spirituality, formative
principles of the same objectivized and “become” Nature. In fact the
physics of Democritus differs from the physics of Newton in that the
chosen starting-point of the one is the optically-given while that of
the other is abstract relations that have been deduced from it. The
“facts” of Apollinian Nature-knowledge are things, and they lie on the
surface of the known, but the facts of Faustian science are relations,
which in general are invisible to lay eyes, which have to be mastered
intellectually, which require for their communication a code-language
that only the expert researcher can fully understand. The Classical,
static, necessity is immediately evident in the changing phenomena,
while the dynamic causation-principle prevails beyond things and its
tendency is to weaken, or to abolish even, their sensible actuality.
Consider, for example, the world of significance that is connected,
under present-day hypotheses, with the expression “a magnet.”
The principle of the Conservation of Energy, which since its
enunciation by J. R. Mayer has been regarded in all seriousness as
a plain conceptual necessity, is in fact a redescription of the dynamic
principle of causality by means of the physical concept of force. The
appeal to “experience,” and the controversy as to whether judgment
is necessary or empirical—i.e., in the language of Kant (who greatly
deceived himself about the highly-fluid boundaries between the two),
whether it is a priori or a posteriori certain—are characteristically
Western. Nothing seems to us more self-evident and unambiguous
than “experience” as the source of exact science. The Faustian
experiment, based on working hypotheses and employing the
methods of measurement, is nothing but the systematic and
exhaustive exploitation of this “experience.” But no one has noticed
that a whole world-view is implicit in such a concept of “experience”
with its aggressive dynamic connotation, and that there is not and
cannot be “experience” in this pregnant sense for men of other
Cultures. When we decline to recognize the scientific results of
Anaxagoras or Democritus as experiential results, it does not mean
that these Classical thinkers were incapable of interpreting and
merely threw off fancies, but that we miss in their generalizations
that causal element which for us constitutes experience in our sense
of the word. Manifestly, we have never yet given adequate thought to
the singularity of this, the pure Faustian, conception of experience.
The contrast between it and faith is obvious—and entirely superficial.
For indeed exact sensuous-intellectual experience is in point of
structure completely congruent with that heart-experience (as we
may well call it), that illumination which deep religious natures of the
West (Pascal, for instance, whom one and the same necessity made
mathematician and Jansenist) have known in the significant
moments of their being. Experience means to us an activity of the
intellect, which does not resignedly confine itself to receiving,
acknowledging and arranging momentary and purely present
impressions, but seeks them out and calls them up in order to
overcome them in their sensuous presence and to bring them into an
unbounded unity in which their sensuous discreteness is dissolved.
Experience in our sense possesses the tendency from particular to
infinite. And for that very reason it is in contradiction with the feeling
of Classical science. What for us is the way to acquire experience is
for the Greek the way to lose it. And therefore he kept away from the
drastic method of experiment; therefore his physics, instead of being
a mighty system of worked-out laws and formulæ that strong-
handedly override the sense-present (“only knowledge is power”), is
an aggregate of impressions—well ordered, intensified by sensuous
imagery, clean-edged—which leaves Nature intact in its self-
completeness. Our exact Natural science is imperative, the Classical
is θεωρία in the literal sense, the result of passive
contemplativeness.

VII

We can now say without any hesitation that the form-world of a


Natural science corresponds to those of the appropriate mathematic,
the appropriate religion, the appropriate art. A deep mathematician—
by which is meant not a master-computer but a man, any man, who
feels the spirit of numbers living within him—realizes that through it
he “knows God.” Pythagoras and Plato knew this as well as Pascal
and Leibniz did so. Terentius Varro, in his examination of the old
Roman religion (dedicated to Julius Cæsar), distinguished with
Roman seriousness between the theologia civilis, the sum of
officially-recognized belief, the theologia mythica, the imagination-
world of poets and artists, and the theologia physica of philosophical
speculation. Applying this to the Faustian Culture, that which
Thomas Aquinas and Luther, Calvin and Loyola taught belongs to
the first category, Dante and Goethe belong to the second; and to
the third belongs scientific physics, inasmuch as behind its formulæ
there are images.
Not only primitive man and the child, but also the higher animals
spontaneously evolve from the small everyday experiences an
image of Nature which contains the sum of technical indications
observed as recurrent. The eagle “knows” the moment at which to
swoop down on the prey; the singing-bird sitting on the eggs “knows”
the approach of the marten; the deer “finds” the place where there is
food. In man, this experience of all the senses has narrowed and
deepened itself into experience of the eye. But, as the habit of verbal
speech has now been superadded, understanding comes to be
abstracted from seeing, and thenceforward develops independently
as reasoning; to the instantly-comprehending technique is added the
reflective theory. Technique applies itself to visible near things and
plain needs, theory to the distance and the terrors of the invisible. By
the side of the petty knowledge of everyday life it sets up belief. And
still they evolve, there is a new knowledge and a new and higher
technique, and to the myth there is added the cult. The one teaches
how to know the “numina,” the other how to conquer them. For
theory in the eminent sense is religious through and through. It is
only in quite late states that scientific theory evolves out of religious,
through men having become aware of methods. Apart from this there
is little alteration. The image-world of physics remains mythic, its
procedure remains a cult of conjuring the powers in things, and the
images that it forms and the methods that it uses remain generically
dependent upon those of the appropriate religion.[489]
From the later days of the Renaissance onward, the notion of God
has steadily approximated, in the spirit of every man of high
significance, to the idea of pure endless Space. The God of Ignatius
Loyola’s exercitia spiritualis is the God also of Luther’s “ein’ feste
Burg,” of the Improperia of Palestrina and the Cantatas of Bach. He
is no longer the Father of St. Francis of Assisi and the high-vaulted
cathedrals, the personally-present, caring and mild God felt by
Gothic painters like Giotto and Stephen Lochner, but an impersonal
principle; unimaginable, intangible, working mysteriously in the
Infinite. Every relic of personality dissolves into insensible
abstraction, such a divinity as only instrumental music of the grand
style is capable of representing, a divinity before which painting
breaks down and drops into the background. This God-feeling it was
that formed the scientific world-image of the West, its “Nature,” its
“experience” and therefore its theories and its methods, in direct
contradiction to those of the Classical. The force which moves the
mass—that is what Michelangelo painted in the Sistine Chapel; that
is what we feel growing more and more intense from the archetype
of Il Gesù to the climax in the cathedral façades of Della Porta and
Maderna, and from Heinrich Schütz to the transcendent tone-worlds
of 18th-Century church music; that is what in Shakespearian tragedy
fills with world-becoming scenes widened to infinity. And that is what
Galileo and Newton captured in formulæ and concepts.
The word “God” rings otherwise under the vaulting of Gothic
cathedrals or in the cloisters of Maulbronn and St. Gallen than in the
basilicas of Syria and the temples of Republican Rome. The
character of the Faustian cathedral is that of the forest. The mighty
elevation of the nave above the flanking aisles, in contrast to the flat
roof of the basilica; the transformation of the columns, which with
base and capital had been set as self-contained individuals in space,
into pillars and clustered-pillars that grow up out of the earth and
spread on high into an infinite subdivision and interlacing of lines and
branches; the giant windows by which the wall is dissolved and the
interior filled with mysterious light—these are the architectural
actualizing of a world-feeling that had found the first of all its symbols
in the high forest of the Northern plains, the deciduous forest with its
mysterious tracery, its whispering of ever-mobile foliage over men’s
heads, its branches straining through the trunks to be free of earth.
Think of Romanesque ornamentation and its deep affinity to the
sense of the woods. The endless, lonely, twilight wood became and
remained the secret wistfulness in all Western building-forms, so that
when the form-energy of the style died down—in late Gothic as in
closing Baroque—the controlled abstract line-language resolved
itself immediately into naturalistic branches, shoots, twigs and
leaves.
Cypresses and pines, with their corporeal and Euclidean effect,
could never have become symbols of unending space. But the oaks,
beeches and lindens with the fitful light-flecks playing in their
shadow-filled volume are felt as bodiless, boundless, spiritual. The
stem of the cypress finds conclusive fulfilment of its vertical tendency
in the defined columniation of its cone-masses, but that of an oak
seems, ever restless and unsatisfied, to strain beyond its summit. In
the ash, the victory of the upstriving branches over the unity of the
crown seems actually to be won. Its aspect is of something
dissolving, something expanding into space, and it was for this
probably that the World-Ash Yggdrasil became a symbol in the
Northern mythology. The rustle of the woods, a charm that no
Classical poet ever felt—for it lies beyond the possibilities of
Apollinian Nature-feeling—stands with its secret questions “whence?
whither?” its merging of presence into eternity, in a deep relation with
Destiny, with the feeling of History and Duration, with the quality of
Direction that impels the anxious, caring, Faustian soul towards
infinitely-distant Future. And for that reason the organ, that roars
deep and high through our churches in tones which, compared with
the plain solid notes of aulos and cithara, seem to know neither limit
nor restraint, is the instrument of instruments in Western devotions.
Cathedral and organ form a symbolic unity like temple and statue.
The history of organ-building, one of the most profound and moving
chapters of our musical history, is a history of a longing for the forest,
a longing to speak in the language of that true temple of Western
God-fearing. From the verse of Wolfram von Eschenbach to the
music of “Tristan” this longing has borne fruit unceasingly. Orchestra-
tone strove tirelessly in the 18th Century towards a nearer kinship
with the organ-tone. The word “schwebend”—meaningless as
applied to Classical things—is important alike in the theory of music,
in oil-painting, in architecture and in the dynamic physics of the
Baroque. Stand in a high wood of mighty stems while the storm is
tearing above, and you will comprehend instantly the full meaning of
the concept of a force which moves mass.
Out of such a primary feeling in the existence that has become
thoughtful there arises, then, an idea of the Divine immanent in the
world-around, and this idea becomes steadily more definite. The
thoughtful percipient takes in the impression of motion in outer
Nature. He feels about him an almost indescribable alien life of
unknown powers, and traces the origin of these effects to “numina,”
to The Other, inasmuch as this Other also possesses Life.
Astonishment at alien motion is the source of religion and of physics
both; respectively, they are the elucidations of Nature (world-around)
by the soul and by the reason. The “powers” are the first object both
of fearful or loving reverence and of critical investigation. There is a
religious experience and a scientific experience.
Now it is important to observe how the consciousness of the
Culture intellectually concretes its primary “numina.” It imposes
significant words—names—on them and there conjures (seizes or
bounds) them. By virtue of the Name they are subject to the
intellectual power of the man who possesses the Name, and (as has
been shown already) the whole of philosophy, the whole of science,
and everything that is related in any way to “knowing” is at the very
bottom nothing but an infinitely-refined mode of applying the name-
magic of the primitive to the “alien.” The pronouncement of the right
name (in physics, the right concept) is an incantation. Deities and
basic notions of science alike come into being first as vocable
names, with which is linked an idea that tends to become more and
more sensuously definite. The outcome of a Numen is a Deus, the
outcome of a notion is an idea. In the mere naming of “thing-in-itself,”
“atom,” “energy,” “gravitation,” “cause,” “evolution” and the like is for
most learned men the same sense of deliverance as there was for
the peasant of Latium in the words “Ceres,” “Consus,” “Janus,”
“Vesta.”[490]
For the Classical world-feeling, conformably to the Apollinian
depth-experience and its symbolism, the individual body was
“Being.” Logically therefore the form of this body, as it presented
itself in the light, was felt as its essence, as the true purport of the
word “being.” What has not shape, what is not a shape, is not at all.
On the basis of this feeling (which was of an intensity that we can
hardly imagine) the Classical spirit created as counter-concept[491] to
the form of “The Other” Non-Form viz., stuff, ἀρχή, ὕλη, that which in
itself possesses no being and is merely complement to the actual
“Ent,” representing a secondary and corollary necessity. In these
conditions, it is easy to see how the Classical pantheon inevitably
shaped itself, as a higher mankind side by side with the common
mankind, as a set of perfectly-formed bodies, of high possibilities
incarnate and present, but in the unessential of stuff not
distinguished and therefore subject to the same cosmic and tragic
necessity.
It is otherwise that the Faustian world-feeling experiences depth.
Here the sum of true Being appears as pure efficient Space, which is
being. And therefore what is sensuously felt, what is very
significantly designated the plenum (das Raumerfüllende), is felt as a
fact of the second order, as something questionable or specious, as
a resistance that must be overcome by philosopher or physicist
before the true content of Being can be discovered. Western
scepticism has never been directed against Space, always against
tangible things only. Space is the higher idea—force is only a less
abstract expression for it—and it is only as a counter-concept to
space that mass arises. For mass is what is in space and is logically
and physically dependent upon space. From the assumption of a
wave-motion of light, which underlies the conception of light as a
form of energy, the assumption of a corresponding mass, the
“luminiferous æther” necessarily followed. A definition of mass and
ascription of properties to mass follows from the definition of force
(and not vice versa) with all the necessity of a symbol. All Classical
notions of substantiality, however they differed amongst themselves
as realist or idealist, distinguish a “to-be-formed,” that is, a Nonent,
which only receives closer definition from the basic concept of form,
whatever this form may be in the particular philosophical system. All
Western notions of substantiality distinguish a “to-be-moved,” which
also is a negative, no doubt, but one polar to a different positive.
Form and non-form, force and non-force—these words render as
clearly as may be the polarities that in the two Cultures underlie the
world-impression and contain all its modes. That which comparative
philosophy has hitherto rendered inaccurately and misleadingly by
the one word “matter” signifies in the one case the substratum of
shape, in the other the substratum of force. No two notions could
differ more completely. For here it is the feeling of God, a sense of
values, that is speaking. The Classical deity is superlative shape, the
Faustian superlative force. The “Other” is the Ungodly to which the
spirit will not accord the dignity of Being; to the Apollinian world-
feeling this ungodly “other” is substance without shape, to the
Faustian it is substance without force.

VIII

Scientists are wont to assume that myths and God-ideas are


creations of primitive man, and that as spiritual culture “advances,”
this myth-forming power is shed. In reality it is the exact opposite,
and had not the morphology of history remained to this day an
almost unexplored field, the supposedly universal mythopoetic power
would long ago have been found to be limited to particular periods. It
would have been realized that this ability of a soul to fill its world with
shapes, traits and symbols—like and consistent amongst themselves
—belongs most decidedly not to the world-age of the primitives but
exclusively to the springtimes of great Cultures.[492] Every myth of the
great style stands at the beginning of an awakening spirituality. It is
the first formative act of that spirituality. Nowhere else is it to be
found. There—it must be.
I make the assumption that that which a primitive folk—like the
Egyptians of Thinite times, the Jews and Persians before Cyrus,[493]
the heroes of the Mycenæan burghs and the Germans of the
Migrations—possesses in the way of religious ideas is not yet myth
in the higher sense. It may well be a sum of scattered and irregular
traits, of cults adhering to names, fragmentary saga-pictures, but it is
not yet a divine order, a mythic organism, and I no more regard this
as myth than I regard the ornament of that stage as art. And, be it
said, the greatest caution is necessary in dealing with the symbols
and sagas current to-day, or even those current centuries ago,
amongst ostensibly primitive peoples, for in those thousands of
years every country in the world has been more or less affected by
some high Culture alien to it.
There are, therefore, as many form-worlds of great myth as there
are Cultures and early architectures. The antecedents—that chaos
of undeveloped imagery in which modern folk-lore research, for want
of a guiding principle, loses itself—do not, on this hypothesis,
concern us; but we are concerned, on the other hand, with certain
cultural manifestations that have never yet been thought of as
belonging to this category. It was in the Homeric age (1100-800 B.C.)
and in the corresponding knightly age of Teutonism (900-1200 A.D.),
that is, the epic ages, and neither before nor after them, that the
great world-image of a new religion came into being. The
corresponding ages in India and Egypt are the Vedic and the
Pyramid periods; one day it will be discovered that Egyptian
mythology did in fact ripen into depth during the Third and Fourth
Dynasties.
Only in this way can we understand the immense wealth of
religious-intuitive creations that fills the three centuries of the
Imperial Age in Germany. What came into existence then was the
Faustian mythology. Hitherto, owing to religious and learned
preconceptions, either the Catholic element has been treated to the
exclusion of the Northern-Heathen or vice versa, and consequently
we have been blind to the breadth and the unity of this form-world. In
reality there is no such difference. The deep change of meaning in
the Christian circle of ideas is identical, as a creative act, with the
consolidation of the old heathen cults of the Migrations. It was in this
age that the folk-lore of Western Europe became an entirety; if the
bulk of its material was far older, and if, far later again, it came to be
linked with new outer experiences and enriched by more conscious
treatment, yet it was then and neither earlier nor later that it was
vitalized with its symbolic meaning. To this lore belong the great
God-legends of the Edda and many motives in the gospel-poetry of
learned monks; the German hero-tales of Siegfried and Gudrun,
Dietrich and Wayland; the vast wealth of chivalry-tales, derived from
ancient Celtic fables, that was simultaneously coming to harvest on
French soil, concerning King Arthur and the Round Table, the Holy
Grail, Tristan, Percival and Roland. And with these are to be counted
—beside the spiritual transvaluation, unremarked but all the deeper
for that, of the Passion-Story—the Catholic hagiology of which the
richest floraison was in the 10th and 11th Centuries and which
produced the Lives of the Virgin and the histories of SS. Roch,
Sebald, Severin, Francis, Bernard, Odilia. The Legenda Aurea was
composed about 1250—this was the blossoming-time of courtly epic
and Icelandic skald-poetry alike. The great Valhalla Gods of the
North and the mythic group of the “Fourteen Helpers” in South
Germany are contemporary, and by the side of Ragnarök, the
Twilight of the Gods, in the Völuspa we have a Christian form in the
South German Muspilli. This great myth develops, like heroic poetry,
at the climax of the early Culture. They both belong to the two
primary estates, priesthood and nobility; they are at home in the
cathedral and the castle and not in the village below, where amongst
the people the simple saga-world lives on for centuries, called “fairy-
tale,” “popular beliefs” or “superstition” and yet inseparable from the
world of high centemplation.[494]
Nowhere is the final meaning of these religious creations more
clearly indicated than in the history of Valhalla. It was not an original
German idea, and even the tribes of the Migrations were totally
without it. It took shape just at this time, instantly and as an inward
necessity, in the consciousness of the peoples newly-arisen on the
soil of the West. Thus it is “contemporary” with Olympus, which we
know from the Homeric epos and which is as little Mycenæan as
Valhalla is German in origin. Moreover, it is only for the two higher
estates that Valhalla emerges from the notion of Hel; in the beliefs of
the people Hel remained the realm of the dead.[495]
The deep inward unity of this Faustian world of myth and saga and
the complete congruence of its expression-symbolism has never
hitherto been realized, and yet Siegfried, Baldur, Roland, Christ the
King in the “Heliand,” are different names for one and the same
figure. Valhalla and Avalon, the Round Table and the communion of
the Grail-templars, Mary, Frigga and Frau Holle mean the same. On
the other hand, the external provenance of the material motives and
elements, on which mythological research has wasted an excessive
zeal, is a matter of which the importance does not go deeper than
the surface. As to the meaning of a myth, its provenance proves
nothing. The “numen” itself, the primary form of the world-feeling, is
a pure, necessary and unconscious creation, and it is not

You might also like