A Brief History of Analysis: Detlef D. Spalt

Download as pdf or txt
Download as pdf or txt
You are on page 1of 265

Detlef D.

Spalt

A Brief History
of Analysis
With Emphasis on Philosophy,
Concepts, and Numbers, Including
Weierstraß‘ Real Numbers
Detlef D. Spalt

A Brief History of Analysis


With Emphasis on Philosophy, Concepts, and
Numbers, Including Weierstraß’ Real
Numbers
Detlef D. Spalt
Darmstadt, Hessen, Germany

ISBN 978-3-031-00649-4 ISBN 978-3-031-00650-0 (eBook)


https://doi.org/10.1007/978-3-031-00650-0

Mathematics Subject Classification: 01Axx, 97-xx

© Springer Nature Switzerland AG 2022


This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether
the whole or part of the material is concerned, specifically the rights of reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or
information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar
methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.

This book is published under the imprint Birkhäuser, www.birkhauser-science.com by the registered
company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
to Juliane
in memory of 19 May 1990
and 27 January 2017
Preface

For Whom Is This Book Written?

This book will be of interest for all those who ask themselves: what is mathematics?
how is it accomplished? from where did its concepts originate?
Some proofs are shown—not the celebrated ones but the fundamental, the
important ones. Among them is one which does not add up but ends in controversy
(Chap. 5). One of the renowned and favourite proofs of devotional mathematical lit-
erature is called into question and proven to be only partially conclusive (Chap. 14).
Not uplift but knowledge is our aim, not well-being but thought.
Which thought led up to today’s mathematics (especially: modern Analysis)?

Who Can Understand This Book?

The understanding of this book does not require an academic degree in mathematics.
However, you need some acquaintance with the central concepts of Analysis
(function and series, continuity and convergence, differential and integral). The
details are always given on the page.

What Is at Stake?

This book shows in a historical way:


• How and by what means was Analysis accomplished? (the formation of
mathematics, the philosophical aspect)
• How do the objects (the concepts) of Analysis develop and change? (the
modelling of mathematics, the conceptual aspect)

vii
viii Preface

Two new mathematical discoveries are added:


• Today’s common Analysis may easily be changed and given another shape—
which is to say: some other theorems become valid. Our theory makes up the
archetype of modern Analysis, which was created by Cauchy.
• Besides the two constructions of the real numbers known today, there exists
a third version, which is completely different. It was invented by Weierstraß
and taught in his foundational analysis course but was not understood and,
consequently, not passed on.
This are two historical issues as well as two items of mathematical news: the history
of mathematics is able to present novelties even to mathematicians!

Who Has Contributed?

My thanks go primarily to a dozen master’s students of mathematics who stimulated


me by their interest, their vigilance and their criticisms (Mr. Marco Pavić was
especially engaged) in the winter semester 2016/17, as this book is based on the
lecture notes written after each session, and subsequently also to Frankfurt am Main
university for granting a fifth—unpaid—teaching assignment in the Department of
Didactics of Mathematics. These students demonstrated that there still exists an
ongoing student interest in the history of mathematics, even though there is no
proper room for this specialization in the mathematics curriculum after the Bologna
Process.
Inevitably, my special thanks go to Prof Dr Jürgen Wolfart for his organizational
contribution as well as for his thoughtful hint to look at the recently discovered
student’s lecture notes from Weierstraß’ lessons in winter 1880/81. Without this
fact, these notes would never have come to my attention.
History and didactics of mathematics have in common that they both focus and
reflect on present mathematics. This may cause a mutually fruitful relationship
as has happened in this case. In particular, I am indebted to Prof Dr Reinhard
Oldenburg, now Augsburg, and Prof Dr Harald Riede, formerly Koblenz, for
their reference to the formula concerning the concept of “continuous derivability”
(p. 144).
As always, I am indebted to my very persistent friends Hassan Givsan and Bernd
Arnold for their everlasting willingness to criticize my tentative formulations in
respect to essence and linguistic quality, thereby helping to clarify and to improve
the book. Finally, Juliane engaged in a final criticism with élan and success: thank
you!
However, without the curiosity of Mrs Iris Ruhmann from Springer, her open-
mindedness and her courage to include this book in her programme, these lecture
notes, as usual, would have been lost as a single document in the university library,
Preface ix

thus remaining unknown to the public. Mrs Ruhmann is entitled to the merit for this
publication, and I thank her very much for that.

Darmstadt, Germany Detlef D. Spalt


15 June 2018

Preface to the Translation

This translation includes one additional mathematical and one further historical
novelty:
• The mathematical novelty is Weierstraß’ concept of real number (in Chap. 12).
Although I had embarked on the study of the newly found lecture notes
taken in Weierstraß’ lessons of winter 1880/81 in preparing my lecture in
winter 2016/17, I failed to understand it correctly. It was only in spring 2017
when I succeeded in unravelling the puzzle and understanding Weierstraß’
fundamentally new idea of a number concept. It is of such a revolutionary
essence that this book seems to be the first documentation of this mathematical
idea in print in English—as no referee of an international first rank journal was
able to understand and accept it for publication within 5 years.1
• The historical novelty is the unearthing of Traugott Müller’s schoolbook from
1838, wherein Müller clearly defines the (positive) irrational numbers in the
very same manner we usually call “Dedekind-cuts” (in Chap. 13).
As to the burden of translation, I am indebted to the longstanding pair of friends,
who taught British A-level Mathematics, Juliane Horn and John D. Smith (London),
who transformed my often colloquial German into decent English. Clearly, no one
could have done better than they did!

Darmstadt, Germany Detlef D. Spalt


7 December 2021

1 See Detlef D. Spalt 2021, The concept of real number, following Karl Weierstraß. URL https://

www.detlef-spalt.de.
Detlef D. Spalt 2022, Die Grundlegung der Analysis durch Karl Weierstraß – eine bislang
unbekannte Konstruktion der natürlichen und der reellen Zahlen. (Springer).
Contents

1 The Invention of the Mathematical Formula . . . . . . . . . . . . . . . . . . . . . . . . . . . 1


Who Invented the Mathematical Formula? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
How Did Descartes Invent the Mathematical Formula? . . . . . . . . . . . . . . . . . . . 2
What Is x for Descartes? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
Literature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2 Numbers, Line Segments, Points—But No Curved Lines . . . . . . . . . . . . . . 11
Mathematics Is in Need of Systematization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
True and False Roots . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
The Geometrical Advantage of Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
Descartes’ Geometrical Successes and His Failure . . . . . . . . . . . . . . . . . . . . . . . . 19
Literature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3 Lines and Variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
From Two to Infinity: Leibniz’ Conception of the World . . . . . . . . . . . . . . . . . . 21
Leibniz’ Mathematical Writings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
Leibniz’ Theorem: Fresh from the Creator! . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
The Precise Calculation of Areas Bounded by Curves: The Integral . . . . . . 28
Leibniz’ Neat Construction of the Concept of a Differential . . . . . . . . . . . . . . 32
What Is x (and What Is dx) for Leibniz? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
Literature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4 Indivisible: An Old Notion (Or, What Is the Continuum
Made of?) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
A Modern Theory?. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
The Continuum and Why It Does Not Consist of Points. . . . . . . . . . . . . . . . . . . 40
The Indivisible . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
Newton’s Method of Fluxions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
Literature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

xi
xii Contents

5 Do Infinite Numbers Exist?—An Unresolved Dispute


Between Leibniz and Johann Bernoulli . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
A Correspondence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
The Subject of the Controversy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
Harmony. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
Johann Bernoulli’s Exciting Position . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
Another Shared (Mathematical) Point of View . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
. . . with Different, Even Contrary Consequences . . . . . . . . . . . . . . . . . . . . . . . . . . 53
Johann Bernoulli’s Position in Dispute . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
Johann Bernoulli Provides the Evidence for His View . . . . . . . . . . . . . . . . . . . . . 55
The End of This Debate: The Disagreement Continues to Exist . . . . . . . . . . 56
Looking Ahead . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
Considering the Real Significance of This Problem: An
Inconsistency in the Actual Mathematical Thinking . . . . . . . . . . . . . . . . . 57
Literature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
6 Johann Bernoulli’s Rules for Differentials—What Does
“Equal” Mean? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
Johann Bernoulli’s Rules for Differentials—Part 1: Preparation . . . . . . . . . . 59
What Does “Equal” Mean? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
Johann Bernoulli’s Rules for Differentials—Part 2: Execution . . . . . . . . . . . . 63
The First Book Containing the Rules for Differentials Stems
from de l’Hospital . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
Literature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
7 Euler and the Absolute Reign of Formal Calculation . . . . . . . . . . . . . . . . . . 69
The Invention of the Principal Notion of Analysis: “Function” . . . . . . . . . . . 69
The Components of Functions: Quantities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
How Did Euler Denote Functions?. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
A Standard Form for Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
A Daring Calculation with Infinite Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
Euler’s Concepts of Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
Analysis Without Continuity and Convergence . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
Continuity According to Euler . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
Euler’s Second Notion of Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
Convergence According to Euler . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
To Sum up Euler’s Algebraic Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
Literature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
8 Emphases in Algebraic Analysis After Euler . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
d’Alembert: Philosophical Legitimation of Algebraic Analysis
as Well as His Critique of Euler’s Concept of Function . . . . . . . . . . . . . 91
d’Alembert’s Impulse: Condorcet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
Lagrange: Making Algebra the Sole Foundation of Analysis . . . . . . . . . . . . . 94
Contents xiii

The Idea of Lagrange . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95


The Fundamental Gap in Lagrange’s Proof. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
Literature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
9 Bolzano: The Republican Revolutionary of Analysis . . . . . . . . . . . . . . . . . . . 101
The Situation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
A New Meaning of Convergence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
Continuity with a New Meaning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
Bolzano’s Revolutionary Concept of Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
Mathematical Consequences of Bolzano’s Notion of Function. . . . . . . . . . . . 112
Literature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
10 Cauchy: The Bourgeois Revolutionary as Activist
of the Restoration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
The Heart of Cauchy’s Revolution of Analysis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
Mathematical View of Cauchy’s Revolution of Analysis . . . . . . . . . . . . . . . . . . 116
Cauchy’s Concept of Variable Is Determined by “values” . . . . . . . . . . . . . . . . . 117
Cauchy Derives “number” from “quantity” . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
“Function” and “value of a function” in Cauchy . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
Some Very Surprising Consequences from Cauchy’s Concept of
“value of a function” . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
Excursus: Preview of a Failed Revolution of Analysis in the
Years of 1958 and 1961 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
Cauchy’s Concept of Convergence: A Big Misunderstanding . . . . . . . . . . . . . 135
Cauchy’s Concept of Derivative—Again a Misunderstanding . . . . . . . . . . . . 142
Cauchy’s Concept of the Integral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
Literature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
11 The Interregnum: Analysis on Swampy Ground . . . . . . . . . . . . . . . . . . . . . . . 149
On the Utility of History of Mathematics. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
The Teaching of Analysis Without a Curriculum . . . . . . . . . . . . . . . . . . . . . . . . . . 151
Analysis as Freestyle Wrestling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
The Foundational Uncertainty of Limes Calculation:
A Prominent Misunderstanding by Prominent Scholars . . . . . . . . . . . . . 160
Literature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
12 Weierstraß: The Last Effort Towards a Substancial Analysis . . . . . . . . 167
A Famous Man . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
The Core of His Fame . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
What is Weierstraß’ Understanding of a “Function”? . . . . . . . . . . . . . . . . . . . . . . 171
Weierstraß’ Concept of Number . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
Infinite Series. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
Weierstraß in Retrospect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
Literature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
xiv Contents

13 Analysis’ Detachment from Reality—and the Introduction


of the Actual Infinite into the Foundations of Mathematics . . . . . . . . . . . 187
A Pessimistic Mood . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
Georg Cantor’s Construction of the Real Numbers . . . . . . . . . . . . . . . . . . . . . . . . 189
The Form of the Real Numbers, Invented by Traugott Müller,
Joseph Bertrand, and Richard Dedekind. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199
The Axiomatic Characterization of the Real Numbers by David
Hilbert in the Years 1899 and 1900 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
The Price of Success: The Inclusion of the Actual Infinite in
Mathematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
Literature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
14 Analysis with or Without Paradoxes? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
Built on Very Thin Ice: Cantor’s Diagonal Argument . . . . . . . . . . . . . . . . . . . . . 215
The significance of the “Diagonal Argument” . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
Paradox I: Conditionally Convergent Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
Paradox II: Methods of Summation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
Paradox III: The Convergence of Function Series. . . . . . . . . . . . . . . . . . . . . . . . . . 224
Paradox IV: The Term-by-Term Integration of Series. . . . . . . . . . . . . . . . . . . . . . 224
Is an Analysis Without Those Paradoxes Possible? . . . . . . . . . . . . . . . . . . . . . . . . 226
Schmieden Dissolves the Paradoxes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
The First Formal Version of a Nonstandard-Analysis in the Year 1958 . . . 233
Finale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
A Satisfying Finish . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240
Literature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240

Author Index. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243


Subject Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
Introduction: The Four Big Topics
of This Book

All those who are not yet familiar with the history of mathematics are advised to
start directly with Chap. 1. However, those who have already had a first encounter
with the history or with the philosophy of mathematics, will get a first orientation
by this detailed overview of the four topics we discussed on p. vii. From p. xviii to
p. xx the shortest possible history of analysis is given.

The Configuration of Mathematics—or: Designing


Mathematical Theories

From the works of the classical Greek philosophers we know: mathematics is


conceptual thought. Whilst Babylonian accountants put up with calculations for
millennia (thereby producing the abstract object “number”), the Greek philosophers
invented the critical discourse, i.e. concepts. With Euclid (c-300), mathematics took
the form it still has today: definition–theorem–proof.
Without definition and without proof, there is no theorem. This book tries to
shed light on those two areas of mathematics, which are usually overlooked: the
definitions (or: “concepts”) and the proofs (which rely on those definitions). It turns
out that these are the decisive issues.

To Define Is Hard Work!

Not the Babylonians, but only Euclid introduced a definition of the object: concept
of “number”.
The definition is the starting point for mathematics as we know it.

xv
xvi Introduction: The Four Big Topics of This Book

The definition tells us what is the subject under discussion, e.g. “number” or
“continuity”. And this is the very spot at which the labour of mathematical thinking
starts: to define the concepts. Mathematicians are not empirical scientists, who can
rely on sensual perception.
Mathematical thinking is not empirical, it does not consider the objects of our
experience.
Which should those be?

Is a Mathematical Proof Beyond Reproach?

Another kind of mathematical labour is to prove a mathematical statement. This


reverses the perspective. The mathematical proofs explore the capabilities of
conceptual expressions.
A proof must be strong, at best irrefutable.
The most forceful proofs in mathematics are the proofs by contradiction. Math-
ematics must be free of contradictions. (Otherwise, everything could be proved: a
theorem together with its negation.) However, not all mathematical proofs work by
contradiction. Some proofs show: a definition is used correctly. Others demonstrate:
so it must be, this is sufficient!
Mostly, the favourite proofs are informal and work without higher concepts. They
are not technical. In case they are completely informal, they suffer from a severe
shortcoming: they are not necessarily conclusive. In Chaps. 5 and 14 I shall present
two such examples. Only a formal, technical proof ends the discussion and shows:
there is no room for developing further arguments.
However, even the greatest did not always succeed and exhaust all the possibili-
ties. In Chaps. 3 and 9, we shall see how two great thinkers reached the limits of their
thinking, essentially, about the same issue. (We may wonder if they were aware of
it.) As a further example of the change of mathematical knowledge we shall present
a simple diagram, which could not have been drawn by Leibniz (or well as by his
contemporaries and his predecessors), on p. 27.

From Confusion to Clarity

Mathematical thinking is formed within definitions and proofs. Through a plenitude


of attempts, concepts are created, tested and rejected. This is how through creative
thinking mathematics is produced. In the end, the so-called mathematical knowledge
can be seen to be a conceptual construction—and as such it is always open to
discussion, in its foundational aspects as well as in its concrete formulations.
Introduction: The Four Big Topics of This Book xvii

The transformation of concepts is at the heart of the historical development of


mathematics.
It is a fundamental object of this book to present this confusion of concepts as
well as of proofs. And that not only in general, but in detail. Great mathematical
thinkers (like Gottfried Wilhelm Leibniz, Johann Bernoulli, Leonhard Euler, Karl
Weierstraß, Georg Cantor) were thoroughly convinced by their own mathematical
arguments, which today’s mathematics values to be not reliable. Some of them will
be cited in their original form.

Growing Insight in the Formative Power of Definitions in


Mathematics

Already, the early modern mathematical thinkers have been aware of the possi-
bilities given by definitions or by mere creativity to shape mathematics (Chaps. 1
till 5). Only in the nineteenth century did a wider professional public take hold
of these possibilities. Only then they agreed: geometry does not only exist in the
plane but also, for instance, on the sphere—however, without validity of Pythagoras’
Theorem, which only holds in the plane. Usually this is expressed as follows:
there also exists a “non-Euclidean” geometry (which is in the true sense a “non-
Pythagorean” one).
It took another century for the insight that such different configurations of a
theory are not at all confined to geometry but are also possible for the most important
and most influential mathematical theory, the Calculus, which we also call Analysis.
A key-word therefore is Nonstandard-analysis. In this book it is shown: there also
exists a Cauchy an Analysis—which is different from each of the theories known
today.

The Change, Seen from a Philosophical Viewpoint

Having acquired this understanding of the formative power of human thinking, it


is not difficult to take the trouble to assess mathematical thinking anew. This will
require a precise decomposition of the details of thought in mathematics, which we
examine.
We shall attempt a reassessment of the history of the most important branch
of mathematics so far: of analysis.
And that in a way which focuses on the foundational ideas and concepts of the
field. The decisive changes (or “revolutions”) in the development of the theory will
be shown. We shall not only stress the importance of naming the new but also
point out which price is paid for any change. There is no innovation, which merely
xviii Introduction: The Four Big Topics of This Book

yields new results, considered to be “good”, without the loss of something old. This
is true also for the way of reasoning—and, consequently, for the development of
mathematics as a whole. Usually, the relinquished parts are kept secret—not here.
(An exception is the first chapter: the birth of modern mathematics. Even now, I am
unable to express this huge transformation, this emergence of modern mathematics,
in just a few words.)
This book concentrates on the (very) broad developments and the fundamental
aspects of the relevant theoretical thought, the philosophical ones. My book Die
Analysis im Wandel und im Widerstreit, which was published in 2015, is much more
detailed (not least in mathematical aspects) and, consequently, more voluminous.
To sum up: I shall pay particular attention to the philosophical aspects of math-
ematical thought. However, the main focus will always be mathematics (analysis)
and its actual development, not philosophical peculiarities and quibbles. It is the
actual mathematics in the minds of the greatest mathematicians which is at stake
here. This is presented by scrutinizing their original texts.

The Formation of Mathematics—or: The Transformations of


Analysis

Analysis is the most important mathematical theory of modern times and thus—
since the nineteenth century—also of the industrial era. It is a dominant part of this
cultural and economic development and one of its necessary driving forces.
This book tries to trace how the central concepts of analysis in the last just
350 years were established: convergence and continuity, differential and integral,
number and function. And it also offers some of the related proofs.

The Foundational Years

It starts with the premise of it all: the creation of the formula by Descartes (not
by Viète!—see Chaps. 1 and 2). Because: without a formula there is no modern
mathematics. (This has rarely ever been told.)
Already the very first mathematical formula contains an “unknown”. The next
stage of development liquified it: the (discrete) “unknown” became the (continuous)
“variable”. Newton as well as Leibniz had this idea (Chaps. 3 and 4). Only fairly
recently it was established that Leibniz had created very precise mathematical con-
cepts and arguments; for instance, he formulated a clear-cut proof of convergence
and he invented the precise concept, which we call today “Riemann integral”
(Chap. 3)—naturally in his own conceptual language.
Thereby, a philosophical abyss opened up: do infinite numbers exist? Two
otherwise always fruitfully cooperating friends from the Champion’s League of
Mathematicians, Leibniz and Johann Bernoulli, were unable to find an agreement
upon this question: see Chap. 5.
Introduction: The Four Big Topics of This Book xix

An Era of Pomposity: Algebraic Analysis

Following Johann Bernoulli, who transferred the focus of mathematical activ-


ity from geometry to calculations (or algebra, Chap. 6), his pupil Euler shaped
the doctrine into “Algebraic Analysis”, which is ruled by the formula, now
called “function”. In the Zeitgeist of absolutism of the eighteenth century, Euler
institutionalized a formalistic etiquette, which disregarded some common-sense
calculations (Chap. 7). D’Alembert and Lagrange worked on a further development
along these lines. However, they did not end up with a genuine redesign (Chap. 8).
Lagrange’s attempt at a foundation failed because of a basic deficiency within his
concepts.

The Implosion of Algebraic Analysis—and a First Attempt to


Replace It

For the newly emerging practical perspective of the engineers, which became
dominant in the nineteenth century, the formulae were of no use if they did not
result in numbers. Bolzano as well as Cauchy did justice to this demand through
being the first to move the “value” into the centre of the theory (Chaps. 9 and
10); Cauchy even succeeded in a complete theory. This gave analysis a new twist.
Because nobody has yet described this transformation, I had to coin a new name for
this doctrine: “Value Analysis”.
However, Value Analysis suffered from a difficult defect of omission: until then,
nobody had invented a definition for a concept of “number”, which was suitable for
analysis—neither Leibniz, nor, especially, Euler. Ultimately, not even Bolzano did
manage it, whereas Cauchy did not even try to do so. This interregnum lasted for
about two generations (Chap. 11).

Implementation of a Capricious Value Analysis

Finally, Weierstraß (the last mathematician who dealt with a traditional form of
analysis, Chap. 12), Cantor and Dedekind (Chap. 13) succeeded in cutting the
Gordian knot. They swallowed the bitter pill and conferred the civic rights of
mathematics to the actual infinite with the help of their concepts of the real
numbers. Thereafter, an agreement on the true concept of function (from Riemann,
not from Dirichlet!) was easily reached.
The ideas which are well known today are those of Cantor and Dedekind, for they
were printed in 1872. Weierstraß’ concept of the real numbers is still unknown
and is presented here in print for the very first time in English. His idea was disclosed
in the lecture notes of one of his students, which was discovered only in 2016; it is
presented here in a short summary (Chap. 12).
xx Introduction: The Four Big Topics of This Book

Outlook: Axiomatics, Analysis Within Set-Theory and a New


Kind of Formal Calculation

One generation later, Hilbert disavowed the traditional kind of mathematical


thinking and paved the way for the structural mathematics à la Bourbaki of the
twentieth century by his axiomatization of the concept of number (Chap. 13). At
first, this attempt was not without flaws—even a Hilbert can make a mistake (p. 206).
The transformation of Value Analysis into Set Analysis in the twentieth century
is not a subject of this book, but the basic philosophical problem of this approach
is clearly indicated here (p. 199). This would quickly run into technical details, and
besides, the result can be found in any modern textbook. But for the sake of clarity,
we emphasize
Set Analysis only arose in the twentieth century
As a note to the specialists: Georg Cantor defines each converging sequence (“Cantor
sequence”) of rational numbers to be a real number (p. 191). (He defined two “numbers” of this
kind to be “equal”, if they differ by an “infinitely small quantity”, which is a sequence converging
on the limit zero.) Cantor did not create the concept of “equivalence relation” (see pp. 62, 120),
which also originates in the twentieth century. He created set-theory only two decades after his
definition of a real number. It is very astonishing (and only known since 2017): the first and only
mathematician who fruitfully used set-theoretical constructions to create a concept for the real
numbers in the nineteenth century was Karl Weierstraß (Chap. 12)—besides, almost without being
noticed, Traugott Müller and François Bertrand, and, of course, somewhat overbearingly: Richard
Dedekind (Chap. 13).

Finally, in Chap. 14, I shall reveal a little-known branch of analysis, which today
is usually hidden behind logic or axiomatics. Here, my starting point is another
unpublished manuscript, for once a notebook from the splendid calculator Curt
Schmieden dating back to the early 1950s. Schmieden did two things. First, he
changed the concept of the real numbers. Secondly, he transformed the way of
calculation with the infinite: instead of describing the infinite in the formulae only
(a) at the start and (b) thereafter only by “ . . . ”, he stuck to the demand (c) to be
precise also at the end of the formulae by saying explicitly what one means by the
infinite there. Schmieden demanded a precise end for the infinite. (Consequently,
his formulae are longer than the usual ones, and, moreover, they require twice the
usual effort—because of the necessity to think about their ending, not only about
their start. However, these efforts always stay elementary.) With the help of Detlef
Laugwitz, Schmieden’s ideas became the first impulse for the invention of a new
theory, which nowadays is called “Nonstandard-analysis”, in 1958. However, in
my opinion, the potential of Schmieden’s ideas is far from being exhausted.
Introduction: The Four Big Topics of This Book xxi

The First Mathematical News in This Book: The Archetype of


Today’s Analysis (from Cauchy)

The central concepts of modern analysis, “convergence” and “continuity”, were


formulated crystal-clearly by Bolzano and Cauchy. However, it was just Cauchy
who developed the theory completely. He based it upon the old basic concept of
“function” (to be a formula) and the new concept of “value of the function” and
added the precise notions of “derivative” and “integral”.
Historically, Cauchy’s notion of “value of a function” was absolutely new. This
mathematical object had never been defined before. Surprisingly to us, Cauchy
thought about this concept rather differently from the way we do (p. 124).
This different understanding of the concept of “value of the function” necessarily
implies consequences for those concepts, which derive from it. These are especially
the notions of “convergence”, “continuity” and “derivative”. They got meanings
which differ from the meanings of their modern counterparts. (Cauchy’s definitions
of “convergence”, “continuity” and “derivative” read just like ours, but their
meaning is different!) “Different” implies: they make some other theorems true,
not those we have today.
In Chap. 10, this is shown in detail. To sum up: the archetype of analysis was
different (and it was also more easy) from its modern development with its changing
concepts of “function” and “value of the function”. (Seen from this perspective,
Cauchy’s concepts may not appear as such a surprise?)

The Second Mathematical News in This Book: A Third


Construction of the Real Numbers (by Weierstraß)

Besides the well-known sources of the two traditional concepts of the “real
numbers” (it is true: I add a hitherto unknown early predecessor of Dedekind, from
1838!), I present a completely new one: from a manuscript which was found by
chance in the mathematical library of the university Frankfurt am Main as recently
as 2016. This precise construction of the concept of real numbers is the most
elementary one known today and has never been re-invented by others.1
The fact that Weierstraß lectured on it during and after Cantor’s and Dedekind’s
ideas were already published might show that Weierstraß thought his own idea to be
superior to the other two.
It is my opinion that Weierstraß’ construction of the real numbers prospectively
should really replace the others, because it is the most elementary one. Cantor’s
construction uses the concept of convergence—which is a petitio principii within

1 Detlef D. Spalt 2022, Die Grundlegung der Analysis durch Karl Weierstraß – eine bislang

unbekannte Konstruktion der natürlichen und der reellen Zahlen. (Springer).


xxii Introduction: The Four Big Topics of This Book

the scope of a deductive composition of analysis. And compared to Dedekind’s


(actually: Traugott Müller’s!) idea, Weierstraß’ is more elementary, because it does
not rely on the puzzling concept of “least upper bound”.
No gain without pain. The said passages (in Chap. 12) are very elementary mathematics, but
as they are yet unknown, they are completely new territory for any reader—and understanding
something new always requires much more effort than the recognition of something already known.
Thus, even mathematicians will have to read these passages with some more care than others.
Mathematics is not trivial. That is why these subsections go into some details.

The Historiographical Hallmarks of This Book


In Substance

It is the maxim of this book to take the former mathematics in its own rights
and to present the mathematical thinking of previous mathematicians as it was—
instead of transforming it into that caricature, which results from its translation into
modern language and modern thought. Only in this way we are able to do justice
to history, because the earlier facts existed without the later ones and without our
actual thinking—just as nowadays we do not have the faintest idea about the shape
of mathematics of the next or even the following century. Only in case of Weierstraß’
concept of number, I make an exception as a concession to the reader: to ease their
understanding.
This attitude is quite unusual, and it has to be implemented especially in
Chaps. 10 and 11, i.e. when we look at the nineteenth century. For this is a period
in which mathematicians, who are not historians, sometimes are led astray. This is
because the “Value Analysis” of that time appears to have much similarity with
today’s set-theoretical analysis. The historical layman as well as the historical
laywoman feels compelled to get along with those texts: look, these are the same
concepts as ours!
Those who are not aware of the fact that mathematical thinking underwent
transformations over time and also that mathematical concepts are invented (or who
is not at least willing to think that this could be the case) will assume that the former
concepts are our modern ones and subsequently will often reach wrong conclusions.
They will take Cauchy’s analysis to be identical with today’s—which definitely is a
mathematical error. It was just its archetypal form.
Unfortunately, I myself got into this trap when I worked for the first time on original texts
between 1977 and 1986. However, not only mathematics changes, but also—at least sometimes—
one’s own understanding of the world. Some succeed in learning new things.

During the last half century, all this led to some grotesque misunderstandings
of the analysis of the nineteenth century. The most dramatic example thereof
(mathematically as well as historically) will be shown in Chap. 10 in some detail,
another one in Chap. 11. At the beginning of Chap. 11, a comprehensive presentation
Introduction: The Four Big Topics of This Book xxiii

of my historiographical method is given, which is an example of the complex facts


stated in Chap. 10: the historiographical judgement of thoughts and concepts within
analysis which differs from all our actual theories—but which is, nevertheless,
consistent in itself (just as the Cauchy ian Analysis). History of mathematics is
not only able to detect (seen from today) single new ideas (like Weierstraß’ concept
of numbers) but even an entire new theory! (Well, of course one has to reckon with
that . . . )
For that reason, Chap. 10 is of a somewhat more complicated structure than the others.
There is not only the subject presented (i.e.: how did Cauchy think analysis?) but also the
two (!) mathematical misconceptions that Cauchy’s thinking will be unravelled—together with
the identification of the errors involved.

In Method

This book is solely based on original sources. (All of them are given in English,
sometimes the translations are new.2 ) Normally, these texts are well known or, at
least, easily accessible (in our digital world far more easily then ever). In Germany,
the digitalization is sadly behind many other countries: for instance the French one,
with its portal Gallica, or the Swiss one, with the portal www.e-rara.ch which
presents the noblest sources. Of general help was also the Californian Internet
Archive; however, things seem to have changed.

All Told

This book aims to ease access to the history of analysis. It gives an overview of the
important lines of its development: particularly by scrutinizing the definitions of the
basic notions, followed by some detailed discussion.
Mathematics changes—like all the other kinds of human thought. “Changes”
means: it becomes different. This becoming different is the subject of history of
mathematics—whereas the actual mathematical thinking usually strives not for a
difference, but, instead, for more.

2 My German book Die Analysis im Wandel und im Widerstreit has, besides the German wording,

also the original.


Chapter 1
The Invention of the Mathematical
Formula

Today, mathematics is unthinkable without formulae. Without:

(a + b)2 = (a + b)(a + b) = a 2 + 2ab + b2

nothing works at all.


Did formulae always exist? Or did somebody invent them? Of course, somebody
must have invented them, they cannot just appear from nowhere!
Then, who did invent the mathematical formula? How can anybody have such an
idea? And why?
These are the three questions, which will be addressed in this chapter.

Who Invented the Mathematical Formula?

Strangely, in school we learn many mathematical formulae, but not where and when
they were created.
It is even more odd that this is nowhere written down! There is no encyclopedia,
including Wikipedia (at least not until today, 2021), that mentions who first thought
of the mathematical formula.
However, reading books can educate and here is the answer to that question:
The mathematical formula was invented by René Descartes.
René Descartes—who is he and does one need to know him?
Yes, one should know René Descartes, at least a little bit! He lived from 1596
until 1650. Being French, as is obvious from his name, he nevertheless lived many
years in the Netherlands. In the twenty years between 1629 and 1649 he moved
about 24 times in order to avoid being disturbed by the authorities.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 1


D. D. Spalt, A Brief History of Analysis,
https://doi.org/10.1007/978-3-031-00650-0_1
2 1 The Invention of the Mathematical Formula

René Descartes thought in ways unacceptable to the officials. The Churches of


central Europe were part of the ruling elite which possessed secular power. It was,
for instance, the Catholic Church that sentenced Galileo Galilei in 1613 for his claim
that the Earth is moving and that the Sun is at the centre of the cosmos. When Galilei
repeated his belief in 1613, he was forced, under threat of torture, to revoke his
announcement and incarcerated by the clerical officials. During that time the Thirty
Years’ War (1618–48) was waged. What begun as a war about religion had turned
into a war solely about power and influence.
Descartes did not want to share Galilei’s destiny. Thus, he preferred not to publish
certain texts. In August 1634, Descartes managed to view Galilei’s forbidden text,
in secrecy, for a few hours. There he read what he could not publish.
Come to that, the text contained no mathematical formulae. Galilei did not know
any. For Galilei it was circles, triangles, and other geometrical objects that according
to him constituted the “letters” of mathematics. These kind of “letters” had already
existed for two thousand years—Descartes radically changed all this.

How Did Descartes Invent the Mathematical Formula?

Descartes, too, is a child of his time. As with every thinker, his thought is shaped
by his experience. This experience was the increasing importance of individual
vocation in order to gain social status, as well as the mechanization of the world.
Independently of their status at birth, an ever increasing number of artisans, artists,
and engineers shaped their lives through their professional activities. Wearable
watches, which were, however, not very precise, had existed already for about a
hundred years (“Nuremberg Egg”); pumps became important in order to provide
big cities, such as London and Paris, with enough water. Fortresses were built, and
so on.
It was Descartes’ special strength to express his Zeitgeist by means of concepts.
He thus anchored philosophical thought in the thinking of the individual: “It is
thinking, thinking alone, which cannot be separated from me: I am, I exist, this is
certain”. This “I” is of course not the man Descartes but the abstract encapsulation of
the concrete individual: I forge iron, therefore I can live (from it)! It is not the social
status at birth that determines a life, but one’s own doing. Descartes subsequently
becomes the founder of modern philosophy, a philosophy that fitted the new, the
bourgeois and mechanized world of the people. And this without the backing of the
authorities. This is a revolution.
And how does Descartes set the foundation for modern mathematics? There are
two aspects which act in unison.
The first aspect is the method. Somewhat pedantically, Descartes demands: Focus
your penetrating mind on the simplest things! Disregard everything superfluous
and unnecessary! Represent everything by clearly perceptible figures. Keep your
notations as brief as possible!
How Did Descartes Invent the Mathematical Formula? 3

Decisive for all this was Descartes’ basic conviction: everything which can be
grasped “clearly and distinctly” is also true. And only that is true.
The second aspect is this: only Geometry is real mathematics. Thus: numbers
and calculations are only then mathematics if they are transferred into Geometry.
The reasons behind Descartes’ thinking will be discussed later. First of all, the
focus should be on how Descartes came up with the mathematical formula.

Transfer Arithmetic into Geometry!

So if real mathematics is restricted to Geometry, it will be necessary to shift the


numbers into Geometry. And, of course, the calculations as well.
To transfer the numbers into Geometry is simple: The number 1 becomes the
length 1, the number 2 the . . . So, instead of mere numbers (as it is with arithmetic)
we have straight lines of the relevant lengths: geometrical objects.
We still need to deal with the calculations, which must happen with straight lines.
Adding and subtracting is simple: the straight lines are put together or they are
taken away from each other. (Of course, negative numbers do not exist: what should
that be, a “negative” number?)
There are still more complex forms of calculations: multiplying and dividing.
This proved problematic for Descartes. First, he did the obvious and declared:
the product of two lines is an area: “If we multiply a by b we put them
a
together at right angles in the following way b in order to get the rectangle
b
a
.
However, if one wanted to multiply the result by c, one would have to imagine
ab

ab as a line a b ab in order to gain abc”.


These were Descartes’ initial thoughts, but it becomes immediately evident: this
is not the way to do it! An area is not a straight line but, according to Descartes’
idea, the product ab is first an area but then, after multiplication by c, the same
product ab suddenly is a straight line.
It is, therefore, justified that Descartes interrupts his uncompleted text at that
point—which he really did.
(He himself did not gave a reason for it. One can only guess. Yet, this is simple
considering what we just have said.)
About ten years later, Descartes presented the solution to his problem. In the
only book on mathematics ever written by him (and which was of course called The
4 1 The Invention of the Mathematical Formula

Geometry), he explains multiplication completely differently:

“Let, for instance, AB be the unity. In order to multiply BD by BC, I only have
to join the points A and C, then to draw DE as a parallel to CA and BE will be the
product of this multiplication”.
Descartes has two new ideas: First, fix a “unity” in a completely arbitrary way.
Secondly, use the principle of similar triangles. Then all becomes simple. If AB =
1, BD = a and BC = b, then all one needs to do is to connect A and C and to draw
the parallel to AC through D—finished! The intersection gives us point E and the
line segment BE. According to the principle of similarity we have:

BD BE a BE
= resp. = , therefore ab = 1 · BE = BE .
1 BC 1 b
Knowing the principle of similarity, it is simple. It was known to mathematicians at
least since it was mentioned by Euclid, about two thousand years ago.
This is all: thus Arithmetic is transferred into Geometry, just as Descartes had
wished. (Division is easy, the principle of similarity is used slightly differently.)
From this point on, instead of calculating, one has to draw!
Now we still miss the formulae! They appear, at least by Descartes, all by
themselves.

Solve Problems!

Let us read a longer piece from Descartes’ Geometry, it is worth doing! Remember
that Descartes is now able to calculate with straight lines.
If, then, we wish to solve any problem, we first suppose the solution already effected,
and give names to all the lines that seem needful for its construction—to those that are
unknown as well as to those that are known. Then, making no distinction between known
and unknown lines, we must unravel the difficulty in any way that shows most naturally the
relations between these lines, until we find it possible to express a single quantity in two
ways. This will constitute an equation, since the terms of one of these two equations are
together equal to the terms of the other. We must find as many such equations as there are
supposed to be unknown lines.
How Did Descartes Invent the Mathematical Formula? 5

Incidentally, since antiquity, this procedure has been called “Analysis” (in
opposition to Synthesis), leading to today’s term “analysis”. Descartes’ text carries
on for a few more sentences and then he writes some formulae:
I may express this as follows:

z o b,
e or

z 2
o −az + b2 ,
e or

z3 o +az2 + b2 z − c3 ,
e or

z 4
o az − c z + d ,
e 3 3 4
resp.

That is, z, which I take for the unknown quantity, is equal to b; or, the square of z is equal
to the square of b diminished by a multiplied by z; or, the cube of z is equal to a multiplied
by the square of z plus the square of b multiplied by z, diminished by the cube of c; and
similarly for the others.

These are the first original equations given as formulae in mathematics.


(To be precise: not entirely original. Descartes used a different equality sign than
that which is used today. But this is not important. It is decisive that a sign is a
symbol that is used instead of a word, as it was the case.) But the formulae do
not depend on language. Formulae are a new language: the language of the new
mathematics.
What is so exciting about it?
Yet, Descartes did not attend any school where he could have learnt this. His
contemporaries wrote such things completely differently.
For instance, François Viète (1540–1603). Viète is officially named as the
“founder of algebra”, by which one really says that he invented formulaic expres-
sions. But he did not. Because in 1593, Viète at most comes up with such things as
this:

B times A
B times A −B times H
+ are said to be equal to B .
D F
Interestingly, Viète puts vowels in place of unknowns, while Descartes is using
the last letters of the alphabet. (The vowels represent for Viète the unknowns—
in similarity to the inventors of the letter-based writing system, the Phoenicians,
who only used consonants and who considered vowels to be “unknown letters”;
in Arabic too, vowels were never written, till today. Only the Greeks introduced
letters to represent vowels. But this observation by Otto Hamborg is only added in
parentheses.)
6 1 The Invention of the Mathematical Formula

However, Viète often writes equations completely differently, for instance, like
this:
Given be B times A to the square plus D planar times A equals Z corporal.

This looks fairly awful. A careful translation, after some thought, could be
written as follows:

B times A2 + D planar · A is equal Z corporal

Descartes would simply have written:

az2 + bz = c

Descartes would completely have omitted the denotations “planar” and “corporal”.
Yet, Viète uses those notations. He thus shows what kind of object he is considering.
In other words: for Viète it is not just the letter that denotes the object in question,
but the meaning of the letter is determined even further. That means, for Viète, letters
are not general symbols but are denoting, together with their labels, specific kinds
of objects.
The last point is of ideological importance: Viète put particular importance on
the “Law of Homogeneity”. That means all the objects which are to be added (or
subtracted) must be of the same kind, i.e. lines, areas or solids. For, according to
Viète, it is impossible to add something “one-dimensional” to something “two-
dimensional”. And we remember: Descartes thought initially the same! But only
in the beginning.
The mature Descartes can think very differently. Because the mature Descartes—
we remember his application of the principle of similarity!—has chosen a “unity”.
Descartes can insert this unity (let us call it “1”) whenever he likes to do so as a
factor or several factors. Therefore, he can read the last equation thus:

az2 + bz · 1 = c · 1 · 1

or also like:

az2
+ bz = c · 1
1
and now it is “homogeneous”! (Of course, he could write something completely
differently, such as:

bz
az2 · 1 + = c · 1.
1·1
It makes no real sense, but it is not wrong!)
How Did Descartes Invent the Mathematical Formula? 7

The later Descartes has shown: the Law of Homogeneity is superfluous. It is an


unnecessary boundary for mathematical thought. More precisely: for calculations.
Descartes managed to do this by defining calculation in a new way. He started by
fixing a “unity”, which makes calculations with line segments easy.
Viète was not able to do this. Because of this he had to fall back on additional
attributes such as “planar” and “corporal” besides his letters.
We learn: that which is for one person “the first and generally valid Law
of Equations”, is, if one thinks about matters completely differently, entirely
superfluous. In this case, the matter is calculating.
The fixing of a unity makes the Law of Homogeneity superfluous.
Thus, what is allegedly a law of thinking can be proven to be obsolete by means
of a different form of thought. If one thinks differently, one’s judgement might
change too. This seems self-evident and applies to mathematics too. Mathematical
judgements too (propositions and theorems) are based on initial rules. They can be
changed by thinking differently or something different. One needs, of course, an idea
how this could be done. (As we saw with Descartes, and as we will see yet again,
such ideas are only found with difficulty.)!
Descartes did have such an idea: the arbitrary fixing of a unity and the application
of the principle of similarity. Strangely, after all: Descartes’ first equations are
homogeneous! Is he following old habits?
And as we are already penetrating fairly deeply into this matter: By utilizing the
principle of similarity, Descartes assumes the validity of Euclidian mathematics.
He does this without spelling it out. Maybe he was not even aware of it. Because
the first sentence in his textbook for mathematics says: “Any problem in Geometry
can easily be reduced to such terms that a knowledge of the lengths of certain
straight lines is sufficient for its construction”. There is no real justification for the
self- congratulatory tone of the statement: without giving it a mentioning, Descartes
assumes Euclidian theory.
Considering all this, it appears to be completely nonsensical to designate Viète as
the inventor of the mathematical formula. It is Descartes, who clearly deserves this
honour. His idea—it took him years of thinking in order to come up with it—finally
looks simple: in order to solve a problem, only write down the relevant straight lines
(numbers), by means of SINGLE letters, as well as the arithmetical operations which
connect them. For the known quantities use the first letters of the alphabet, for the
unknowns the last. That this is not really easy, as can be seen in school: many have
problems in grasping this idea.
Of course, this does not mean that Viète’s partly formalized way of notating was
entirely useless. Quite the opposite: Viète was the first who was able to write down
how to get the solution of an existing equation in an orderly and succinct way. Viète
also gives examples of how to calculate the coefficients (which he did not call so)
of an equation from its solutions—a procedure which we now refer to as “Viète’s
formulae”.
8 1 The Invention of the Mathematical Formula

Why Does Descartes Have Those Ideas?

Why does Descartes only consider Geometry to be proper mathematics, but not
Calculating (Arithmetic)? The answer is deeply anchored in Descartes’ world view.
We know already that Descartes kept away from the authorities as much as
possible. He knew: if his real thoughts became known, he would be persecuted
like Galilei. So, what were Descartes’ revolutionary thoughts about the world?
These: a human being is not what is usually claimed, a combination of body
and mind. Instead, body and mind are two fundamentally different and, therefore,
separate entities. More precisely: there are only two “substances” which exist in the
world. One is matter and one is mind; or in a different formulation, extension and
thought. (That is why one usually speaks about Cartesian dualism: for Descartes
there exist only two fundamental substances.)
All things in the world are mere appearances of those two substances. Solids
and space are only “modi” (the plural of the Latin word “modus”) of matter (or
extension); soul and time (namely: memory) modi of the mind (or of thought).
Strong words, which are of course heretical: God merely a modus of thought?
But this was how Descartes thought. He also wrote it down. Yet, in order to
avoid trouble, he could not express his thoughts so clearly. For that reason, his
philosophical texts contain inconsistent statements: safety measures. In order to
understand Descartes, one needs to read his texts very carefully while considering
his political situation.
Thus, according to Descartes, there are exactly two substances: matter (or
extension) and mind (or thought) with their exemplifications of body and soul.
Bodies or solids, in being modi of matter are appearances caused by motion.
Souls are modi of mind. Subsequently, body and soul are really distinct; a unity of
both does not exist. Not really distinct are, however, matter and bodies; thought and
mind; your soul and my soul; soul and God.
But what are “numbers” for Descartes? Obviously they are not bodies. Although
we count bodies, the numbers we deduce in the process, are not independently
existing entities. “Numbers” derive from counted objects and thus are connected to
them. If one disregards these connections one only arrives at “nonsense” (Descartes
might also have thought of numerological practices). Instead, one has to “abstract”
from the “numbers”.
If a right-angled triangle has shorter sides of lengths 3 and 4, the arithmetician
says that the third side has length 5. The mathematician,
√ that means the geometri-
cian, however, will say that it has the length a 2 + b2 (if a and b are the shorter
sides): “The two parts a 2 and b2 remain distinct, while they are mixed up in the
number 5”.
The number is distinguished from the counted objects in the same way as the
quantity from the extension. Therefore, there cannot be a science of numbers (no
arithmetic)—but definitely one of extension (Geometry).
Descartes, therefore, had to transfer Arithmetic into Geometry where numbers
became (as a matter of simplification) straight line segments. But, as a consequence,
What Is x for Descartes? 9

one had to calculate with lines. Therefore, it was important to pay attention to
the correct method: “Direct your attention toward the simplest objects! Disregard
everything superfluous and unnecessary! Express everything by figures of the
intuition! Keep your notations as short as possible!”
The result reached by Descartes was the purely symbolic mathematical formula.
But the formula is the decisive means of construction for modern mathematics.
Nobody before Descartes had this idea.
Six years before the publication of Descartes’ Geometry, with its purely symbolic
formulae, another book containing such formulae was published. It was found
among the writings of Thomas Harriot, who had died ten years earlier in 1621.
(Harriot even wrote down the equality sign which is used until today and which was
introduced two generations earlier in England.) Admittedly, the author did not use
the symbolism of powers, which often led to confusing and prolonged calculations.
Besides, all the equations follow the Law of Homogeneity—clearly, because Harriot
did not have the same revolutionary idea as Descartes.
Descartes always denied having had knowledge of this book before 1637 and
such knowledge could not be proven until today.
It became quickly apparent that mathematics had gained a completely new
perspective for further development. A case in point is that a purely symbolic
equation becomes a new mathematical object—for which one can create a theory.
One can, for instance, carry out calculations with them.

What Is x for Descartes?

Clearly, Descartes’ equations deal with “numbers”. The known numbers are denoted
by the letters at the beginning of the alphabet; others are unknowns: those marked
by “x”. Initially, Descartes rather writes “z”, where we, today, usually write “x”,
yet, in the last chapter he changes to “x”.
To calculate an unknown x may be simple:

x−1=0

has obviously the value 1.


Sometimes, there are several values. In the last chapter of his book, Descartes
treats equations as the objects of calculations:
If one multiplies the two equations,

x−2=0 and x−3=0

one gets

x 2 − 5x + 6 = 0
10 1 The Invention of the Mathematical Formula

or

x 2 = 5x − 6 ,

and this is an equation, where the quantity x can take the value 2 as well as the value 3.

Thus, x is not necessarily uniquely determined. Of course, there could be even


more values which are represented by x; one merely has to multiply more than two
of those equations.
Of course, Descartes has no “negative” numbers—because his “numbers” are
line segments. If an equation leads to such a result, as in

x + 5 = 0,

then it has a “false solution” which is 5.


The result:
Descartes’ “x” denotes one or more UNKNOWN numbers.
Descartes did not need to say what “numbers” are: he had transferred them into
lines. The meaning of line segments (“lengths”) could already be found in Euclid. He
also holds that there are “rational” and “irrational” (for Euclid “incommensurable”)
numbers.
Albeit, until the concept of “irrational” number was finally established another
212 years would pass. We will come to this quite toward the end, in Chaps. 12 and
13.

Literature

Bos, H. J. M. (2001). Redefining geometrical exactness. Descartes’ transformation of the early


modern concept of construction. Berlin: Springer.
Cajori, F. (1929). Controversies on mathematics between Wallis, Hobbes and Barrow. The
Mathematics Teacher, 22, 146–151.
Descartes, R. (1637a). Discours de la methode. Leyde: Ian Maire. http://gallica.bnf.fr/ark:/12148/
btv1b86069594.r=descartes+discours.langDE.
Descartes, R. (1637b). Geometrie. Darmstadt:Wissenschaftliche Buchgesellschaft, 1981. reprint of
the 2nd ed. from 1923, Leipzig. German by Ludwig Schlesinger 1894.
Hobbes, T. (1662). Seven philosophical problems and two propositions of geometry. In W.
Molesworth (Ed.) The english works of Thomas Hobbes (Vol. VII). London: John Bolton, 1845.
Second reprint: Aalen: Scientia Verlag, 1966.
Irrlitz, G. (1980). René Descartes – Ausgewählte Schriften. (Leipzig: Philipp Reclam jun.)
Schmitz, M. (2010). Analysis – eine Heuristik wissenschaftlicher Erkenntnis. Freiburg im Breis-
gau: Verlag Karl Alber.
Smith, D. E., & Latham, M. L. (Eds.). (1954). The geometry of René Descartes. New York: Dover.
Stevin, S. (1586). De Beghinselen der Weeghconst. Leyden: Christoffel Plantijn. http://www.
google.com/books?id=_wo8AAAAcAAJ&hl=de.
Chapter 2
Numbers, Line Segments, Points—But
No Curved Lines

Mathematics Is in Need of Systematization

In 1637 Descartes invented the pure symbolic formula for mathematics. This
became possible only after having unmasked the Law of Homogeneity as null and
void, as an unnecessary curtailment of thinking.
Only in the course of time is the quality of an idea revealed. Often it is the novelty
of an idea that hinders its appreciation and its dissemination. Not everybody loves
news; clinging to habits is more conducive to comfort.
Thomas Hobbes (1588–1679) was a contemporary of Descartes, being born eight
years earlier and dying 29 years after him. The famous Encyclopedia Britannica
describes him as a philosopher and a political theorist but Hobbes considered
himself also to be a competent mathematician. Nonetheless, until today he is not
known for this quality to historians of mathematics. This is generally agreed and
will not be questioned here.
In this way, Hobbes may be a good example of a thinker who is highly esteemed
in some matters (like philosophy and politics) but cuts a poor figure in other subjects
(mathematics). √ √ √ √
Considering a calculation like 9 · 2 = 92 · 2 = 81 · 2 = 162, he asks,
in the year 1662, how this could be possible—for the first product is a rectangle but
the second a line segment.
I see the calculation in numbers is right, though false in lines. The reason whereof can
be no other than some difference between multiplying numbers into lines or planes, and
multiplying lines into the same lines or planes.

Hobbes was unable to reason without the Law of Homogeneity in the year 1662,
although Descartes had uncovered it as a superfluous barrier for reasoning in 1637.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 11


D. D. Spalt, A Brief History of Analysis,
https://doi.org/10.1007/978-3-031-00650-0_2
12 2 Numbers, Line Segments, Points—But No Curved Lines

True and False Roots

Descartes supplied mathematics with the new entity: the pure formal equation. And
what do mathematicians prefer? They prefer to calculate. If there exist new entities,
they calculate or attempt to calculate with them.
And so did Descartes. At first he multiplied his new objects, the equations.
The equation

x−2=0

is very simple. Its solution is natural: x = 2. Descartes adheres to the traditional


name for the solutions of an equation and says: 2 is the “root” of this equation.
The equation

x−3=0

has root 3. And now it happens: the equations are manipulated; they are multiplied.
We have already seen this in the preceding chapter. The product of these two
equations is:

(x − 2) · (x − 3) = x 2 − 5x + 6 = 0 .

The left side directly shows that the two roots are 2 and 3, for a product a · b is zero,
if at least one of the two factors is 0.

What Are False Roots? And What Is Their Use?

Now, instead of x − 3 = 0 let us choose the equation

x + 3 = 0.

What is the root of this equation? Today we say: the root is −3. But Descartes did
not! Instead he says: this equation has the “false” root 3.
Why?
Well, he has transferred arithmetic to geometry and taken numbers to be line
segments. To consider 3 as a line segment seems easy if you have a unity, as
previously proposed by Descartes. However, what about −3? What kind of line
segment could this be?
Today, mathematicians, physicists, or engineers will answer immediately: “the
opposed line segment!”
That is what we have been taught: a straight line has a direction. However, this is
anything but clear. A straight line does not have a direction by itself. (Maybe it has
True and False Roots 13

two of them, but surely not one. Indeed, the idea of the directions of a straight line
had been proposed by the Dutch engineer Simon Stevin (1548/9–1620) in the year
1586, but in 1637 the philosopher Descartes was not interested in this proposal.)
Who uses “opposed to” has to assign a direction to all straight lines. He or she
only knows directed lines as opposed to straight lines per se.
Descartes did not see the need for this distinction. Why should he have done?
What is the use of straight lines having a direction in geometry? None at all! Euclid
did not know straight lines with a direction. And as we will see, Leibniz, two
generations later, did not subscribe to the idea that straight lines have a direction
in geometry. But without “directed” lines there can be no “opposed” line segments!
It is only calculation with line segments that needs a direction, not the calculation of
line segments. This was shown by Descartes.
When we think, in various ways we presume things without being aware of them,
for example, that straight lines always have a “direction”. Obviously, this view is
possible. But as Descartes shows, it is far from necessary to think in this manner.
“Negative” line segments do not exist in classical geometry and neither in Descartes’
view.

Consequently, the equation x + 3 = 0 does not have the “negative” root −3 but
the “false” root 3. And why should it?
If we now look at the product

(x − 2) · (x + 3) = x 2 + x − 6 = 0 ,

it has the (“true”) root 2 and the “false” root 3.

Turn False into True

Of course, the “false” roots are an eyesore.


Descartes has two ideas on how to get rid of them.

1. Swap the signs of the terms with x, x 3 , x 5 , etc.


That means, instead of

x2 + x − 6 = 0

you consider

x2 − x − 6 = 0 .

Because we have x 2 − x − 6 = (x + 2) · (x − 3), this equation has the (“true”)


root 3 but also the “false” root 2. Easy come, easy go. What turns false roots
into true ones, equally turns true roots into false.
14 2 Numbers, Line Segments, Points—But No Curved Lines

Can’t we do anything about that? Yes, we can:


2. Increase the roots by a sufficiently large quantity.
Consequently, we enlarge the roots of our initial equation

(x − 2) · (x + 3) = x 2 + x − 6 = 0

with the “true” root 2 and the “false” root 3 by, let us say, 4:

y = x + 4.

Then we have to do some calculation:

(x − 2) · (x + 3) = ([y − 4] − 2) · ([y − 4] + 3)
= (y − 6) · (y − 1) = y 2 − 7y + 6 = 0 ,

and indeed we now get two “true” roots for our equation: 6 and 1.

This approach always succeeds: we are always able to find a sufficiently large
number which, if added to all roots of the equation, gives only “true” roots of it. In
this way we do not need to encounter the “false” roots. We may get rid of them.

The Geometrical Advantage of Equations

In his Geometrie, Descartes has the proud boast that he is able to solve a
problem that defeated the ancient mathematicians. This is right, but the problem
is complicated to such a degree that I cannot present it here. We have to confine
ourselves to the fundamentals of the case.

Analysis: From Problem to Equation

Descartes describes how he tackles the problem (all emphases are added):
First, I suppose the thing done, and since so many lines are confusing, I may simplify
matters by considering one of the given lines and one of those to be drawn (as, for example,
AB and BC) as the principal lines, to which I shall try to refer all others.
Call the segment of the line AB between A and B, x, and call BC, y. [. . . ]

We note two important aspects: the first is Descartes’ method. We have already
seen this in Chap. 1 (p. 4): Consider the problem to be solved, name the quantities
and look for relations between them. From antiquity this method has been called
“Analysis” (Fig. 2.1).
The Geometrical Advantage of Equations 15

Fig. 2.1 The Pappos Problem, or: Descartes’ system of coordinates (Discours 1637, pp. 309, 311)

There is a second aspect: Descartes introduces something, which we later, after


its full development, call “coordinate axes”. He has “principal lines” AB and BC
and calls the “points” on AB “x”, those on BC “y”.
So much to consider!
1. At once we are accustomed to the “axis” AB being perpendicular to BC. But
for Descartes things were different: the angle between the axes was arbitrary.
2. For Descartes, “x” and “y” were no longer “unknowns” (unknown numbers)
but “arbitrary” points.
3. The intersection of the axis, B, is not the “origin” of the coordinate system
and does not signify a “zero” for those axes. This is due to the fact (which is
not shown here) that Descartes only works with positive lengths. Naturally, he
adds and subtracts these positive line segments (or at least almost always).
We simplify the matter and avoid Descartes’ tedious path to solve his problem.
We merely examine his solution. It is to be found in the second chapter of The
Geometry:

n p 2
y =m− x+ m2 + ox − x ,
z m

where m, n, o, p, and even z (!) are fixed quantities, which he introduced during his
Analysis.
Please note: this equation is homogeneous. Descartes did not write it in the
elegant way shown above but used the old notation: “mm” instead of “m2 ”, etc.
He gives the following explanation for his equation:
This must give the length of the line BC, leaving AB or x undetermined.
16 2 Numbers, Line Segments, Points—But No Curved Lines

Thus did Descartes reach the goal of his Analysis of the problem: he found
the equation which shows the connection between the “indeterminates” x and y.
In respect of calculating, or to be more sophisticated, algebraically, he solved the
problem.
But according to the classical view, this is not sufficient: it remains to be proved
that the result of the Analysis really does solve the problem. In other words, one
must be able to compose the problem starting from the solution. From ancient times
this process of composing has been called “Synthesis”. Only the Synthesis shows
that the Analysis was right.
This is hardly surprising: after all this is geometry, not arithmetic! Descartes still
has to show the geometrical object which solves the initial (geometrical!) problem.
Our understanding as modern mathematicians who no longer equate mathematics
to geometry but are accustomed to algebra, can easily appreciate that the above
equation solves the problem. Needless to say, for Descartes and his predecessors
things were very different.

Interjection: Continuity

Descartes still has to draw the line representing the arbitrary quantities which are
essential to the equation. There is a categorical distinction between lines and points:
lines are continuous, but points are not.
That lines are continuous objects means that they cannot cross without intersect-
ing.
In the case of two lines of points this is not guaranteed: maybe one of them—or
even both!—does not have a point where they cross, albeit without meeting.
This is the classical contrast between discrete (a line of points) and continuous.
Euclid operates with straight lines and circles. (a) His Parallel Postulate guaran-
tees that two straight lines which are not parallel do intersect. He quietly assumes
that a circle intersects (b) a straight line or (c) another circle, if either object is
sufficiently near. As Euclid only works with those objects, this is all he ever needs
in respect to continuity.
From the continuity of circles follows the continuity of ellipses, parabolas, and
hyperbolas, for these figures are all conic sections.
Any cone emerges if a straight line rotates along a circle in space, one point fixed.
Thus the continuity of the surface of a cone follows from the continuity of both the
straight line and the circle.
A plane is continuous. Consequently, the line of intersection of a plane and
a cone is continuous. But these lines of intersection are ellipses, parabolas, or
hyperbolas, so the continuity of these lines follows from Euclid’s assumptions.
(These arguments are attributed to Menaechmus, a friend of Plato.)
Straight lines and circles can be drawn with the help of a ruler and a compass.
The foregoing considerations show that conic sections can also be so constructed as
continuous lines.
The Geometrical Advantage of Equations 17

Synthesis: From Points to Curved Lines? (I)

As an outcome of his Analysis, Descartes found as the solution of the problem the
equation

n p 2
y =m− x+ m2 + ox − x .
z m

This equation shows how to find for each given indeterminate quantity x a root (or
two roots?) y. These two (or three) numbers x and y denote lengths (see the figure
on p. 15). As measuring starts from point B, x and y can also signify points on the
“principal lines” AB and BC. We know how to work with coordinates: we draw a
parallel to the principal line BC through the point x on AB; and likewise we draw a
parallel to the principal line AB through the point y on BC. The result is one point.
This point represents the pair of roots x, y, calculated beforehand.
But then—how do we get the line from the point?
Well, we do not have one single point, but many: to each number x we find a
number y.
(Maybe not to each, as the expression under the square root must not be negative:
but many; perhaps even infinitely many, as for every two admissible numbers—
which should exist!—we have infinitely many others in between, and they, too,
should be admissible.)
For that reason, the relevant question is: how do we extend the infinitely many
points to become a line?
At this point Descartes becomes somewhat chary. He writes tersely:
If then we should take successively an infinite number of different magnitudes x for the line
[AB], we should also obtain an infinite number [of magnitudes] y for the line [BC], and,
therefore, we have an infinity of different points, such as C, by means of which the required
curved line could be drawn (descrire).

Or:
Having explained the method of determining an infinite number of points lying on any
curved line, I think I have furnished a way to describe them.

The Admissible Curved Lines

“. . . a way to describe them”: what can Descartes “describe”? This he explains at


the very beginning of his second chapter:
Now to treat (tracer) all the curved lines which I mean to introduce here, only one additional
assumption is necessary, namely two or more lines can be moved, one upon the other,
determining by their intersection other curved lines.
18 2 Numbers, Line Segments, Points—But No Curved Lines

With this characterization, and breaking with tradition, Descartes narrows the
range of geometry (for him: mathematics): he only accepts one single movement
for tracing lines.
The spiral, the quadratix [. . . ] must be conceived of as described by two separate movements
whose relation does not admit of exact determination.

(A spiral is created when a radius rotates around a point while the pencil on it
creeps outwards. The speeds of rotating and creeping are independent from each
other, forming a completely arbitrary ratio. As Descartes is unable to perceive
this ratio “clearly and distinctly”, the spiral is not an admissible object in his
mathematics. Analogously with the quadratix.)
The accepted lines are called by Descartes “geometrical” (for him the same as
“mathematical”).
Descartes gives another requirement for the admissibility of lines, namely
that all points of those curved lines which we may call “geometric”, that is, those which
admit of precise and exact measurement must bear a definite relation to all points of a
straight line, and that this relation must be expressed by means of a single equation.

In other words: Descartes only accepts those lines as “geometrical” (“mathemat-


ical”) which can be described by only one equation.
Descartes’ equations have the indeterminates only as summands, factors, or in
the denominator of a fraction but never as exponents. Therefore, we can state, in
today’s language, that for Descartes the admissible (“geometrical”) lines are the
algebraic curved lines. Descartes excludes transcendental curved lines from his
geometry (mathematics)—i.e. those whose equations have x in the exponent.

Synthesis: From Points to Curved Lines? (II)

Again: how do we get from the equation to the line? This is unclear and is still
unclear even in Descartes’ writings!
For, amazingly, Descartes was lacking in rigour regarding the transition from
already given points to a continuous line. For example, he constructs the points
of an oval with the help of the sections of two circles having F and G as centres
(Fig. 2.2):
He finished his description of this procedure with the following statement:
[. . . ] we can thus find as many points as may be desired, by drawing lines parallel to 7 8
and describing circles with F and G as centres.

Today this may be sufficient (the reason for this will be understood later), but
it was unacceptable in Descartes’ time and did not meet classical standards. The
intersection of circles around F and G—if they are near enough—is guaranteed by
Euclid, and, therefore, the points of the oval which Descartes constructed do exist.
But how do we know that one of these points is lying on the straight line AR? Maybe
all of these intersections of the circles avoid this line? The figure does not seem to
Descartes’ Geometrical Successes and His Failure 19

Fig. 2.2 Pointwise geometrical construction of an oval (Discours 1637, pp. 352, 358)

look like that—but this is not conclusive! Only if it were proved that the oval could
be constructed would its intersection with the straight line AR be secured, as the
meeting of two continua.
But Descartes is unable to draw this oval.
This shortcoming of Descartes is the more annoying for at the beginning of his
book he accused classical mathematics of exactly this failure. Right at the beginning
of his book he writes:
Then, since there is always an infinite number of different points satisfying these require-
ments, it is also required to discover and trace the curved line containing all these points.
Pappos says that when there are only three or four lines given, this line is one of the three
conic sections, but he does not undertake to determine, describe, or explain the nature of the
line required when the question involves a greater number of lines. He only adds that the
ancients recognized one of them which they had shown to be useful [. . . ] This led me try to
find out whether, by my own method, I could go as far as they had gone.

“As far”—yes, but not further.

Descartes’ Geometrical Successes and His Failure

In Chap. 1 we saw how Descartes embedded arithmetic (the numbers and the
calculations with them) in geometry and that he invented the pure symbolic formula.
In this chapter we described the way in which Descartes analyzed a given
problem of geometry, that is to say, how he translated it into his language of
formulae, how he forms a suitable equation.
However, according to classical standards the Analysis of a problem—even if
it is successful—is not sufficient. On the contrary, an Analysis has to be followed
by a Synthesis: a construction of the problem from the components given by the
Analysis.
20 2 Numbers, Line Segments, Points—But No Curved Lines

But the Synthesis was Descartes’ undoing, at least in general. He did not succeed
in finding a method of tracing the line from the equation. We arrive at the surprising
conclusion:
Descartes did not invent the coordinate system.
Sometimes he succeeded, sometimes not.
In his very detailed study of Descartes’ book the historian of mathematics Henk
Bos summarizes Descartes’ constructions of the examined curved lines as follows:
[. . . ] the reconstructed arguments in the solutions above do not depend on the knowledge of
the equations of the curved lines. They can be achieved entirely without the help of algebra.
Moreover, I have not been able to devise a purely algebraic line of arguments leading in a
natural way from the problems to the turning ruler and moving curve procedures for tracing
the solution curves.

In short: Analysis done—Synthesis gone!


But did an alternative exist? Descartes’ “x” are numbers or line segments; at
best “x” and “y” allow the determination of points, maybe even infinitely many.
But points are not lines; discrete objects are not continuous. Obviously, Descartes
is lacking a continuous x.

Literature

Bos, H. J. M. (2001). Redefining geometrical exactness. Descartes’ transformation of the early


modern concept of construction. New York, Berlin, Heidelberg: Springer.
Cajori, F. (1929). Controversies on mathematics between Wallis, Hobbes and Barrow. The
Mathematics Teacher, 22, 146–151.
Descartes, R. (1637). Discours de la methode. Leyde: Ian Maire.
http://gallica.bnf.fr/ark:/12148/btv1b86069594.r=descartes+discours.lang.DE
Descartes, R. (1637). Geometrie. Darmstadt: Wissenschaftliche Buchgesellschaft, 1981. reprint of
the 2nd edition 1923, Leipzig. German by Ludwig Schlesinger 1894.
Hobbes, T. (1662). Seven philosophical problems and two propositions of geometry. In: W.
Molesworth (Ed.), The English works of Thomas Hobbes (Vol VII). London: John Bolton,
1845. second reprint: Aalen: Scientia Verlag, 1966.
Schmitz, M. (2010). Analysis – eine Heuristik wissenschaftlicher Erkenntnis. Freiburg im Breis-
gau: Karl Alber.
Stevin, S. (1586). De Beghinselen der Weeghconst. Leyden: Christoffel Plantijn. http://www.
google.com/books?id=_wo8AAAAcAAJ&hl=de
Chapter 3
Lines and Variables

From Two to Infinity: Leibniz’ Conception of the World

In Descartes we came to know a radical critic of contemporary rational thinking.


With Gottfried Wilhelm Leibniz (1646–1716) we now turn to an equally radical
diplomat who endeavoured to unite the world’s antagonisms.
According to Leibniz, God created the “best possible world”. His world consists
of prime elements or atoms which Leibniz called “Monads”. He says:
The Monad, of which we shall speak here, is nothing but a simple substance, which enters
into compounds. By ‘simple’ is meant ‘without parts’.

All Monads differ from each other:


Indeed, each Monad must be different from every other. For in nature there are never two
beings which are perfectly alike and in which it is not possible to find an internal difference,
or at least a difference founded upon an intrinsic quality.

Monads are beings, simple beings. Fundamentally,


[. . . ] nothing but this (namely, perceptions and their changes) can be found in a simple
substance. It is also in this alone that all the internal activities of simple substances can
consist.

Leibniz regarded it as a “metaphysical necessity”


that every created being, and consequently the created Monad, is subject to change, and
further that this change is continuous in each.

These central statements are to be found in the small booklet written in 1714,
which was published four years after Leibniz’ death as Monadology.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 21


D. D. Spalt, A Brief History of Analysis,
https://doi.org/10.1007/978-3-031-00650-0_3
22 3 Lines and Variables

This way of thinking is radically antagonistic to Descartes. Whereas Descartes


presumed two substances (extension and thought), Leibniz presumed (actual)
infinitely many (the Monads). In other words, whereas Descartes had principles,
Leibniz had individuals.
Leibniz’ wider conceptions of the world are not connected to his mathematical
thinking. The principal idea is: Monads are nothing else than perception as well as
their change, an important assumption which underlies his entire philosophy. Even
if he does not mention this explicitly, he intends it to be a fundamental fact. We will
soon recognize this.

Leibniz’ Mathematical Writings

Leibniz never worked as a regular mathematician. Nonetheless he is, together with


Isaac Newton, one of the two creators of that mathematical theory which developed
into the most powerful of all, till today: calculus.
Leibniz documented his invention in a lengthy essay; the longest mathematical
text he ever compiled. It was supposed to be printed in Paris after Leibniz’ departure
from the town. But this did not happen and eventually the manuscript was lost.
Leibniz, much later, refused to rewrite the paper: in his eyes, too much time had
passed already.
However, among the vast amount of papers left by Leibniz, some drafts of this
essay were found. An extract from such a draft was first published by Lucie Scholz
in her dissertation in 1934, a complete version in 1993 by Eberhard Knobloch
and meanwhile (2012) also in Leibniz’ Schriften; a French translation by Marc
Parmentier appeared in 2004 and a German translation by Otto Hamborg has been
available since 2007 via the internet and since 2016 as a book.
Reading Leibniz’ papers (which is not very easy as he wrote mostly in Latin)
repudiates all the prejudices of present-day scientists against their predecessors:
they were less intelligent than we are, their ideas were more vague, their reasonings
inconclusive—only today are we qualified to be exact and precise.
The point is this: many earlier scientists did not use “vague” notions and “diffuse”
reasonings—but completely different ones. If one engages with those different
concepts and follows these other ways of reasoning, one realizes that Leibniz’ proofs
are not incorrect at all. Quite the opposite, Leibniz’ ideas are marked by great
ingenuity. His proofs, judged by today’s standards, are as precise as his concepts
allowed for.
Especially the following three concepts: convergence, integral and differential
stand out as great scientific achievements.
Leibniz’ Theorem: Fresh from the Creator! 23

Leibniz’ Theorem: Fresh from the Creator!

The Convergence of Infinite Series

Up until today the basic curriculum of higher mathematics contains “Leibniz’


Theorem”. It is about infinite series, such as

1− 1
2 + 1
4 − 1
8 + −...

or

1− 1
3 + 1
5 − 1
7 + 1
9 − +...

As series are infinite sums, it is not possible to calculate them directly. Nevertheless,
they often have a value, a sum. The two above happen to have a sum, but others do
not. The series

1 − 2 + 3 − 4 + −...

or

1 + 4 + 8 + 16 + . . .

do not have a sum, that is to say: no finite value.


If one ponders for a while perhaps the following idea may arise: promising
candidates for series with a sum may be those which fulfil two conditions: (i) the
summands, their terms are ALWAYS decreasing and can become as small as one
likes them to be; (ii) their signs alternate.
Consequently, the first term is greater than the sum, as the second term is taken
away from it; the sum of the first two terms is smaller than the sum, as the third
term is added to it: condition ii., etc. Finally: the difference (a change!) between the
successively calculated sums decreases and becomes arbitrarily small: condition i.
This phenomenon that “the sum becomes increasingly more accurate if the
differences between sums, which can truly be calculated, decrease below each given
quantity” is what we call “convergence” today. Yet, during Leibniz’ time this name
was not established.
Leibniz was the first ever to describe these novel objects of mathematics (“convergent
series”) systematically.
24 3 Lines and Variables

Leibniz’ Formulation of His Theorem

In his manuscript Leibniz wrote:


If a quantity A is equal to a series b − c + d − e + f − g, etc.,

A = b − c + d − e + −... ,

which decreases infinitely in such a way that the terms eventually become smaller than an
arbitrarily given quantity, it will be
+b greater than A, so that the difference is smaller than c
+b − c smaller . . . . . . is smaller than d
+b − c + d greater . . . . . . smaller than e
+b − c + d − e smaller . . . . . . smaller than f .
And in general, the part of the decreasing series with alternating additions and subtractions,
which ends with an addition, will be greater than the sum of the series, the part which ends
with a subtraction will be smaller; but the error or the difference will always be smaller than
the term of the series, which follows the part at once.

With the exception of the line “A = b − c + d − e + − . . .” and the last


three “is smaller than”, this is exactly what Leibniz wrote in his manuscript. It is
as precise as possible. Leibniz describes in all detail what an “infinite series” with
“alternating signs” and “steadily decreasing” terms which “decrease below each
assumed quantity” is.
It is permissible to read this statement of Leibniz as follows:
Theorem. If the terms of a series

b − c + d − e + −...

decrease indefinitely (i.e. they eventually get less than any assumed quantity) than this series
has a finite sum.

Leibniz did not leave it at this description but added a fairly detailed proof.

Leibniz’ Proof of His Theorem

Leibniz’ proof consists of a preliminary consideration and four steps.


Preliminary: as the amount of the terms steadily decreases, there is altogether
more added than subtracted. Added were b + d + f + . . ., subtracted were c + e +
g + . . . and we have b > c, d > e, f > g, etc. In this manuscript Leibniz does not
use a “greater” sign, but in others he does.
Especially we have for the sum A:

A < b.
Leibniz’ Theorem: Fresh from the Creator! 25

1. The first step of the proof: using the last inequality we may consider the number

b − A.

(We see: Leibniz tries to deal with positive, true numbers!) Therefore,

b − A = b − (b − c + d − e + f − g + − . . .) = c − d + e − f + g − + . . . < c ,

where the last inequality < holds for the same reason as was given for A < b
in the preliminary!
2. The second step of the proof: we found that A < b. In the same way we derive
from (1) that

A > b−c.

Is it?—Really!—Therefore, it is allowed to take A − (b − c) (which is also a


true number) and we get

A − (b − c) = d − e + f − g + − . . . < d ,

by the same argument.


3. Leibniz’ third step of the proof: as before, through rearranging the result above,
we get as the next starting point:

A<b−c+d.

So we can build (b − c + d) − A. And by the same procedure we arrive at

b − c + d − A = e − f + g − +... < e.

Leibniz adds a fourth step in the same manner, but we need not repeat it here, as
the pattern is clear by now.
As a result Leibniz gets a sequence of inequalities:

A<b
b−A<c
A − (b − c) < d
(b − c + d) − A < e
A − (b − c + d − e) < f
...
26 3 Lines and Variables

From the second line onwards we have on the left side the error that arises,
if instead of the whole series only its beginning till the nth term is taken and
the right hand side shows that this error is always less than the (n + 1)st term.
But it was presupposed that the terms’ magnitude “eventually decrease below any
assumed magnitude”, and, therefore, this assumption—as can be seen in the above
inequalities—leads to the conclusion that the the error caused by breaking off the
series also becomes less than any assumed magnitude.
What more do we want?

It is impossible to give a more precise, stronger argument.

And that even today. Leibniz did so in his manuscript of late summer 1676.

Reflection on Leibniz’ Achievement

Through these notes Leibniz became the first person in the history of mathematics
to describe, as precisely as possible, what is nowadays called the “convergence” of
an “(infinite) series”. (The attribute “infinite” is plainly superfluous but it sounds so
breathtakingly bombastic.)
The expression “convergence” was not used by Leibniz. However, the importance
is the content of the finding, the name is quite irrelevant. What counts is that Leibniz
stated his result in absolute clarity.
The philosophical question: what did Leibniz do? What was the object of his
inquiry?
My answer: obviously, Leibniz explored a variable quantity, i.e. the successively
calculated sum of a series. Today this is called the “partial sum” and we have a
standard symbol for it: sn . We write:

s1 = a
s2 = a − b
s3 = a − b + c
s4 = a − b + c − d
s5 = a − b + c − d + e
...
sn = a − b + c − + . . . ± n

It is evident that this object, the “partial sum” or sn , is a variable (i.e. changing)
quantity. Leibniz only had the right hand sides of the above equations and, therefore,
Leibniz’ Theorem: Fresh from the Creator! 27

they can be named without any reservations (“sn ” will do as well as any other name).

Without explicitly saying so, Leibniz studied variable quantities.

Leibniz’ way of arguing is so fascinating, so deeply mathematical that nobody


dared to characterize it as “unmathematical” and to criticize it: According to
the standards of classical mathematics this is not mathematics!—At least to our
knowledge nobody dared to make such an accusation, written or otherwise.

An Idea Which Leibniz Could not Grasp and the Reason for His
Inability

Today we illustrate the above facts by the following picture:

At a single glance we grasp the situation. However,

Leibniz was unable to draw this picture.

For this picture demands two things: (a) lengths have a direction, and (b) “negative”
numbers do exist.
Actually we have seen earlier that a length has no direction! Moreover, if negative
numbers were true “numbers”, some of the more than two thousand years old laws
(known at least since Euclid!) would be invalidated.
One of these laws reads as follows:
a c
If = , and if a > c, then b>d.
b d
But if −1 is a “number”, this law requires:

1 −1
As = , and if 1 > −1 it follows − 1 > 1.
−1 1

and thus a contradiction. Contradictions are absolutely forbidden in mathematics for


otherwise everything whether false or correct can be proved.
28 3 Lines and Variables

The mathematicians of the late seventeenth century and the beginning of the
eighteenth century had to make a decision: should those time-honoured laws be
preserved, or should −1 become a true number?
Leibniz was an astoundingly creative thinker but not a revolutionary; he was a
diplomat. He shunned revolutions (the dismissal of the validity of these classical
laws) but praised the news by couching it in unctuous words:
Nevertheless, I do not want to deny [. . . ] that −1 is a quantity smaller than nothing; this only
has to be understood right-minded. Such statements are what I call passable true (following
the renowned Joachim Jungius); [. . . ] However, they would not bear a severe verification,
but yet they are of great help for the calculation and of immense value for the inventive
genius as well as for universal concepts.

The classical phraseology of diplomats: “. . . / as well as . . . ”; “−1 is smaller than


1 / but this has to be correctly understood”; “to be on the safe side, I cite an
authority, however, unknown or vague, or hint at my obedience: strictly speaking
this is forbidden / but it is of huge utility”.
Thus, the paper published by Leibniz in a famous scientific journal in April 1712
can be understood as an act of great political diplomacy.
The quintessence is:
Accepting “negative” numbers as true ones goes hand in hand with the invalidity of some
time-honoured laws.

If he or she may say something like: “this law is valid only for positive numbers,
and all is fine”, the known contradictions are outlawed. (Hopefully, no others will
appear we have not yet thought of!)

The Precise Calculation of Areas Bounded by Curves: The


Integral

The Beginning Is Easy

Lasting for thousands of years, the Babylonians operated quite differently compared
with classical Greek scholars, but in Greek culture the following art of planimetry
was taught:
It only allowed for rectangles. The area is

length times width.

width

length
The Precise Calculation of Areas Bounded by Curves: The Integral 29

From this all further calculations had to be deduced, e.g.: the area of a rectangular
triangle is

1
times base times height.
2

height

base

The Problem

What happens, if the boundary of one side is curved?

The Solution of Leibniz—The Original Way

For such an instance Leibniz has the following astounding idea:


I present Leibniz’ original figure, omitting what is not essential to the principal
idea. Even so, the figure is still sufficiently complicate (Fig. 3.1).
The object of the construction is the area between the curve D1 , D2 , D3 , D4 and
the three line segments D1 B1 , B1 B4 , and B4 D4 . The points Dn are any points on
the curve.
1. Leibniz takes the stairway B1 , N1 , P1 , N2 , P2 , N3 , P3 , B4 , B1 as a first
approximation of the area. In the following and for the sake of simplicity we
will indicate the single steps of the stairway by their dotted “upper” lines, the
“step zones”, i.e. N1 P1 , etc.
2. Now let us assess the error of this approximation! It consists of the sum of
partial errors. Compared with the actual area we get:
(a) The first step N1 P1 is too big by the (curved) triangle D1 N1 F1 —as well
as too small by the triangle F1 P1 D2 . Leibniz is saying that at any rate this
first partial error is less than the whole rectangle in between D1 and D2 .
This is really generous, isn’t it!
30 3 Lines and Variables

D4

N3
P3
F3

D3

N2
P2
F2

D2
N1
P1
F1
D1

A
B1 B2 B3 B4

Fig. 3.1 Leibniz’ figure to calculate the area (purified), 1676

(b) The same with the following step N2 P2 , again the second partial
error caused by the approximation is clearly less than the rectangle in
between D2 and D3 .
(c) And so forth.
3. So what is the total error at most? Most likely it is less than the sum of the
partial errors; undoubtedly it is less than the sum of the rectangles D1 , D2 ;
D2 , D3 and D3 , D4 .
4. Therefore, what is our estimate of the total error? The total height is obviously
(if the curve is always ascending or always descending; otherwise it must be
divided in such way) the height from D1 to D4 . And, to be on the safe side,
Leibniz chooses as width the maximal width of the steps B1 B2 , B2 B3 and
B3 B4 . In the example it is B3 B4 .
Outcome: the total error of the first approximation is clearly less than the
product of the height from D1 to D4 and the maximal width of the Bn Bn+1 .
5. All boils down to the following conclusion:
The points D on the curve are completely arbitrary. We may choose as many of
them as we please. Let’s say, we choose k equidistant points such that the steps
have width B1 Bk/(k−1).
Consequently, the total error of this k th approximation will be certainly less than
the rectangle with width B1 Bk/(k−1) and height D1 Dk (more precisely, B1 Dk ).
The Precise Calculation of Areas Bounded by Curves: The Integral 31

Whereas the height of the rectangle remains the same, its width continuously
decreases if further points D are chosen. And the total error of the k th estimate
is clearly less than the area of the corresponding rectangle.
Subsequently, the product of these two values will also drop: if in the product
bk · h the factor bk is constantly decreasing while h (the height) remains the
same, the product decreases, too.
Thus the total error of the estimate reduces further.
6. How small does it become?
Obviously, there is no lower limit to the area’s magnitude (actually, its
“smallness”): with the exception of the limit zero, of course. According to
Leibniz’ own words:
The points D may be thought of as near and in such a great number that the straight-
lined step-shaped area differs from the four-lined area D1 B1 B4 D4 D3 etc. D1 itself
by a quantity which is less than an arbitrarily given.

This means: the total error which emerges from the calculation of the area
below the steps instead of the area limited by the curve, can be decreased to
any desired degree of exactitude.
Did we hit the jackpot? Do we have the area?—Yes and no. On the one hand, we
have a method of calculation: Leibniz is capable of calculating the area as precisely
as he wants to.
On the other hand, being able to calculate something does not mean to have
a concept for doing it. Engineers may be satisfied with a method of calculation,
but mathematics needs concepts! In this case, an appropriate concept of numbers.
However, Leibniz could not offer one. Understandably so, as it turned out mathe-
matics needed two further centuries to coin such concepts (Chaps. 13 and 12) and
in a certain sense, these solutions contradict Leibniz’ thinking, for they demand the
acceptance of the “actual infinite” in mathematics (see Chap. 5).

Outlook

Leibniz’ idea from 1676 provided the foundations for what was later called
“integration”, especially for the “integral” of a “function”, the graph of which
Leibniz still called “curved line”.
However, the names “integral” as well as “function” were already used by Leib-
niz, but both with other meanings.
(a) “Function” was Leibniz’ title for a multitude of line segments which can be
constructed to make up a curve when set in relation to straight coordinate axes:
abscissa, ordinate, tangent, normal, sub-tangent, sub-normal, resecta, . . . , a lot
of special geometrical constructions, which were frequently studied in his times.

Leibniz used the name “function” without a definite meaning.


32 3 Lines and Variables

(b) The name “integral” was invented by Johann Bernoulli and in 1690 his brother
Jacob used the notation for the first time in a printed paper. The first printed
document in which Leibniz used the sign “ ” dates back to the year 1686. He
used it in connection with his sign for the “differential”: “ dx ”—a concept which
will be treated in the next section. As “d” is an operator, “ dx ” as well as “ dy ”
are to be read as a single quantity in the following text.
If one transforms the differential equation p dy = x dx into the “summatorial” [by
building the sums on both sides], one has p dy = x dx . From what I have shown
in the method of tangents, we clearly have d( 12 xx) = x dx; therefore, the reverse is
1
2 xx = x dx (because like powers and roots in usual calculations, we have sums and
differences or and d as reciprocal).

The first evidence of the integral sign, as it is still used today, “ ”, goes back to
1691. Leibniz had encountered the name “integral” for the first time in 1690, in an
article by Jacob Bernoulli
However, the topic that was presented above, i.e. the exact calculation of an
area with a curved boundary, was made the subject of mathematics, as precisely
as 1676, only 178 years later. According to its later inventor Bernhard Riemann, the
mathematical object is nowadays known as the “Riemann integral”. Shortly before
Riemann, Cauchy had come up with a similar idea (pp. 145f).
To sum up: Leibniz had already developed this notion as precisely as possible—
but without today’s concepts of “function”, “infinite series”, and “convergence”. It
worked without these notions!
If Leibniz’ plan had succeeded and his manuscript had been printed, mathematics
would have developed differently.

Leibniz’ Neat Construction of the Concept of a Differential

The First Publication: A False Start

Leibniz published his idea of “differential” from the 1670s in October 1684. Albeit,
his explanation remained very vague. Even worse, Leibniz made a mistake such
that the intelligent reader had to decide whether the whole treatise was wrong or
merely the definition of its fundamental notion. In case of the latter, the reader had
to reconsider the correct concept of “differential” all by himself.
Clearly, this publication was almost a complete failure. Where to find a clever
and astute reader? Jacob Bernoulli (1654–1705) had tried hard to understand this
text since 1687. In the end he needed two or three years to understand the main
idea in order to use the new method himself. What’s more, he was able to develop it
further, in dialogue with his younger brother Johann—and with Leibniz.
Leibniz’ Neat Construction of the Concept of a Differential 33

Another False Start: The New Edition

The differential calculus consists of the concept of “differential” as well as laws of


calculation for those differentials. Without these laws the concept is of no use.
Leibniz owed the explicit formulation and the detailed foundation of these laws
to the public of his time. Just as before, he wrote a manuscript thereon but did
not publish it. In 1846, when it was finally published, together with a lot of other
manuscripts on this topic, nobody took notice. The same happened in 1920, when
the English translation appeared. This fact has been documented only fairly recently:
by Henk Bos in 1972 and by Richard T. W. Arthur in 2013, who are both historians
of mathematics.

The Neat Construction, Part I

In regard to the above presentation it comes as no surprise that Leibniz took also
the “differential” to be a geometrical notion. More precisely, with help of the
“differential” it should become possible to draw a tangent to an arbitrary curved
line (Fig. 3.2).
A “tangent” is a straight line that touches the curve, i.e. that snuggles up to it. To
touch usually means there is no cut, the curve remains on the one side of the straight
line. However, there are unusual curves and sometimes the “tangent” still cuts those
curves.
Leibniz came up with the following geometrical construction of the differential.
The abscissa x is drawn upwards, the ordinate y to the right (today we usually do it
the other way round). Take a parabola

x2
y= ,
a

Fig. 3.2 Leibniz’ calculation A


of the tangent T
X1 Y1

dx

X2 dy Y2
D
34 3 Lines and Variables

which is represented by the curved line in the figure. Leibniz chooses AX1 = x and
X1 Y1 = y. From the point Y1 the perpendicular line Y1 D to the larger horizontal
line X2 Y2 (the ordinate) is drawn. The difference of AX2 and AX1 is called by
Leibniz the “differential” dx , similarly, the difference of X2 Y2 and X1 Y1 is called
the “differential” dy . These are the notations, now the calculations:
The equation of the curve reads

x2
y= .
a
Leibniz starts by changing x to x + dx and subsequently y to y + dy . This modifies
the equation to:

(x + dx )2 x 2 + 2x dx + dx 2
y + dy = = .
a a
We subtract the original equation and get:

2x dx + dx 2 2x + dx dy 2x + dx
dy = = · dx or = .
a a dx a
Of course, dx and dy are changing quantities. And as we would expect they
decrease indefinitely: below any given quantity.
Therefore, the numerator of the fraction will approach 2x. But then, as zero is
not allowed as denominator, we encounter a problem regarding the dx on the left
side of the equation. Yet, if dx does not truly reach its limit zero, the numerator on
the right will not become = 2x.
What are we to do?

Interlude: The General Rule: The Law of Continuity

Leibniz needs an argument that enables him to extend the validity for cases only
holding for dx = 0 to the case dx = 0.
And indeed, Leibniz really had such an argument at his disposal: his Law of
Continuity. According to the historian of mathematics Herbert Breger, this cognitive
law has a similar significance and power for Leibniz as later the Method of
Dialectics has for Hegel. It is a “universal scheme of thought and cognition”. And as
we might expect, coming from Leibniz, its main characteristic is a diplomatic rather
than a logical one. The Law of Continuity unites opposites instead of separating
them as logic does.
Generally speaking, Leibniz’ Law of Continuity says: different things, even if they are
contradictory, can emerge (or be understood) from a common principle.
Leibniz’ Neat Construction of the Concept of a Differential 35

This is the principal idea. Naturally, the law may be given some more precise
formulation, in accordance with the concrete requirement.
Thus, in order to give his concept of “differential” a rigorous foundation, Leibniz
formulated his Law of Continuity in his manuscript as follows:
If some continuous transition ends in some limit, it should be allowed to state a common
law of thought that includes the last limit.

This bails him out.

The Neat Construction, Part II

Leibniz takes the differential triangle Y1 D Y2 . It is his idea to oppose it to an


auxiliary triangle with two features: (a) It is similar to the differential triangle.
(b) One of its sides is fixed.
This auxiliary triangle originates by producing the side Y2 Y1 to T on the axis
A X2 .
The auxiliary triangle T X1 Y1 has the same angles as the differential triangle and
is, thus, similar to it.
Then point Y2 moves on the curved line (in our example: the parabola) toward
the point Y1 . What is going to happen?
Point T moves along the vertical axis A X2 up and down, whereas the side X1 Y1
remains fixed. However, the differential triangle Y1 D Y2 and the auxiliary triangle
T X1 Y1 remain similar (if the latter triangle degenerates, one has to think anew).
Next, if point Y2 coincides with point Y1 , we have exactly the situation which
is covered by the Law of Continuity in its above specification: the coinciding
of the points Y2 and Y1 represents the limit; the case in which the differential
triangle has vanished. However, the auxiliary triangle survives the coinciding of
Y2 and Y1 —for its side X1 Y1 is fixed and cannot vanish. Therefore, this auxiliary
triangle T X1 Y1 represents the COMMON PRINCIPLE which covers BOTH CASES
the vanishing differential triangle AND its disappearance.
Both triangles are similar. Consequently, we have

D Y2 dy X1 Y1
= = .
Y1 D dx T X1
dy
Leibniz already knows: dx = 2x+ dx
a . So all together he has:

dy X1 Y1 2x + dx
= = .
dx T X1 a

Actually, Leibniz can apply his Law of Continuity: on the left there is a fraction
in which numerator and denominator vanish together; the middle is a fraction with
changing value but both, numerator and denominator, remain and stay finite; on the
36 3 Lines and Variables

right there is a fraction with vanishing dx in the numerator but all other quantities
remain fixed.
Next, Leibniz lets the dx vanish, i.e. dx → 0. The outcome is obvious:

dy 2x
= ,
dx a
which is with the help of his Law of Continuity absolutely neatly derived—although
on the left is a fraction where numerator as well as denominator tend toward zero!
This is exactly the point! Leibniz does not divide 0 by 0 but analyzes a ratio
dx
dy with simultaneously (or, at the same time—but we should keep time out of
dy
mathematics!) vanishing terms. In today’s notation: lim = 2x
a . The auxiliary
dx→0 dx
triangle provides Leibniz with a fixed point to unhinge the world. The Law of
Continuity, invented by him, allows him the precise calculation of this ratio dy/ dx
even in the case where dx → 0 as well as dy → 0.

What Is x (and What Is dx) for Leibniz?

Leibniz develops the differential as a geometrical concept. He denotes the respective


length on the x-axes by “x”. Of course, this length is not invariable—just the
opposite! For Leibniz the length x is changing and this change is described by
“x + dx ”.

For Leibniz both x and dx are varying lengths.

We know, Descartes denoted by “x” a certain fixed number or length. Leib-


niz revolutionized Descartes’ world of concepts completely! Only the letter “x”
survived—as if nothing had happened. But an upheaval took place: the continuum
was introduced in the calculations on the quiet.
Of equal importance is that Leibniz created dx as a changing quantity which
decreases below any given quantity. This circumstance was named by him an
“infinitely small quantity”.
Consequently, an “infinitely small quantity” is for Leibniz nothing supernatural,
inconceivable—but only a special case of a commonly used changing quantity: just
one which decreases indefinitely (although Leibniz had no “negative” numbers).
Using the concept “limit” one may say: for Leibniz an “infinitely small quantity”
is a changing quantity with zero as its “limit”.

Literature

Arthur, R. T. W. (2013). Leibniz’s syncategorematic infinitesimals. Archive for History of Exact


Sciences, 67, 553–593.
Literature 37

Bernoulli, J. (1690). Op. XXXIX. Analysis problematis antehac propositi. In Opera, S. T. (Vol. 1),
pp. 421–426. Genevæ: Hæredum Cramer & Fratrum Philbert, 1744. Acta Erud., Mai 1690,
pp. 217–219; cf. Bos 1974, p 21.
Bos, H. J. M. (1974). Differentials, higher-order differentials and the derivative in the Leibnizian
Calculus. Archive for History of Exact Sciences, 14, 1–90.
Breger, H. (2016). Kontinuum, Analysis, Informales – Beiträge zur Mathematik und Philosophie
von Leibniz. In W. Li (Ed.). Berlin: Springer Spektrum.
Buchenau, A. (1966). Gottfried Wilhelm Leibniz – Hauptschriften zur Philosophie (Vol. 2). In
E. Cassirer. Vols. 107, 108 of the series Philosophische Bibliothek. Hamburg: Felix Meiner,
1904–1906.
Gerhardt, C. I. (Ed.) (1846). Historia et Origo Calculi Differentialis a G. G. Leibnitio conscripta.
Hannover: Hahn’sche Hofbuchhandlung. https://archive.org.
Hofmann, J. E. (1966). Leibniz als Mathematiker. In W. Totok & C. Haase (Ed.), Leibniz, sein
Leben – sein Wirken – seine Welt. Hannover: Verlag für Literatur und Zeitgeschehen.
Knobloch, E. (Ed.) (2016). Gottfried Wilhelm Leibniz: De quadratura arithmetica circuli ellipseos
et hyperbolae. Klassische Texte der Wissenschaft. Berlin: Springer. German by Otto Hamborg.
Latta, R. (1898). Leibniz. The Monadology and Other Philosophical Writings. Oxford: Clarendon
Press.
Leibniz, G. W. (1982). Monadologie. In Gottfried Wilhelm Leibniz: Vernunftprinzipien der Natur
und der Gnade. Monadologie. Vol. 253 of the series Philosophische Bibliothek (pp. 26–69).
Hamburg: Felix Meiner. German by Herbert Herring 1956. Also in: Buchenau 1904–06, Vol. II,
pp. 435–456, 1714
Leibniz, G. W. (1993). De quadratura arithmetica circuli ellipseos et hyperbolae cujus corollarium
est trigonometria sine tabulis. In E. Knobloch (Ed.), Abhandlungen der Akademie der Wis-
senschaften in Göttingen. Göttingen: Vandenhoeck & Ruprecht.
Leibniz, G. W. (2004) Gottfried Wilhelm Leibniz: Quadrature arithmétique du cercle, de l’ellipse
et de l’hyperbole et la trigonométrie sans tables trigonométriques qui en est le corollaire. Lat.
text: Eberhard Knobloch (Libraire Philosophique. Paris: J. Vrin), ed.: Marc Parmentier.
Leibniz, G. W. (2011). Die mathematischen Zeitschriftenartikel. In H.-J. Heß & M.-L. Babin
(Eds.). Hildesheim, Zürich, New York: Georg Olms Verlag.
Leibniz, G. W. (2012). De quadratura arithmetica circuli ellipseos et hyperbolae. In Leibniz-
Archiv Hannover (ed.), Gottfried Wilhelm Leibniz – Sämtliche Schriften und Briefe. Vol. 6 –
1673–1676, Arithmetische Kreisquadratur, of the series VII Mathematische Schriften, pp. 520–
676. N 51 (1676). https://www.gwlb.de/leibniz/digitale-ressourcen/repositorium-desleibniz-
archivs/laa-bd-vii-6.
Scholtz, L. (1934). Die exakte Grundlegung der Infinitesimalrechnung bei Leibniz. Dissertation
thesis, Hohe Philosophische Fakultät der Philipps-Universität Marburg. Görlitz-Biesnitz: Ver-
lagsanstalt Hans Kretschmer.
Chapter 4
Indivisible: An Old Notion (Or, What Is
the Continuum Made of?)

A Modern Theory?

Gottfried Wilhelm Leibniz invented a theory and the language of this theory is still
used in modern mathematics.
 Today, just like him, we write differentials as “ dx ”,
“ dy ” and integrals as “ y dx ”.
However, today these signifiers do not have the same meaning  as they had
for their daring inventor. (Let’s also remember that the sign “ ” was not at all
invented by Leibniz but by his younger colleague Johann Bernoulli.) Whereas
Leibniz thought of areas and lines as purely geometrical quantities, these are seen
today as much more general concepts. Nowadays, the differential “ dx ” is often
only a symbol of calculation without any material significance—whereas Leibniz
thought it, as we have seen, to be a “variable (geometrical) quantity”, “decreasing
indefinitely”, “below any given quantity”, and “eventually vanishing”.
The transition from Descartes to Leibniz showed us a complete semantic change
of the “x” in the formulae! Descartes thought the “x” to be an unknown “number”
which was to be calculated—but essentially for him it signified a definite length of
a line segment as he held arithmetic at bottom to be the same as geometry (although
in some new version, e.g. a “unit” had to be defined). To sum up, Descartes thought
of the “x” as a “number” or as a “line segment”—whereas Leibniz made “x” to be
a “variable quantity”, or in short: a “variable”.
In the following chapter we shall learn in which way Leibniz’ concept of
differential was developed and changed by Johann Bernoulli. We shall come to this
later and note for now:

The foundational concepts of mathematics, with which we are dealing, change over time.

So one has to be careful not to superimpose our modern understanding on an old


text. At least, we should try not to, as it is all too easy to fall into ingrained habits.
Let’s see!

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 39


D. D. Spalt, A Brief History of Analysis,
https://doi.org/10.1007/978-3-031-00650-0_4
40 4 Indivisible: An Old Notion (Or, What Is the Continuum Made of?)

Leibniz Knew His Theory Was Descended from an Old Tradition

Leibniz formulated the foundations of his new theory during the years 1674–76 in
Paris. Naturally, he continued working on them. Because of his first publication
on the topic in 1684, which was hard to understand (we mentioned this on p. 32),
two very capable mathematicians had become alert: the brothers Jacob and Johann
Bernoulli. Subsequently, this triumvirate developed Leibniz’ fundamentals to a great
extent.
As yet in the year 1692, Leibniz called his system “our analysis of indivisibles”.
Thus, nearly twenty years after his invention he stuck to his original methodology:
the “method of indivisibles”. What this is about will be indicated in this chapter.
Only indicated, for the details are too intricate to be explained here. But the basic
ideas are of great importance: we need to realize that Leibniz did not create his
theory from scratch. The studies of former scholars opened up the path toward Leib-
niz’ creation. Nevertheless it was he who developed these studies further and into
another direction which turned out to be so very successful. In a similar way, the
same holds for Newton.
Before we turn toward the concept of “indivisible”, we have to focus on an issue
which is even older, and which has the name “continuum”. Today we still work with
the notion of “continuum” but do not use “indivisible”.

The Continuum and Why It Does Not Consist of Points

What Is the Continuum?

The continuum is cohesive, unbroken, connected. Prototypes are the line, the area,
the volume as well as the course of time—Thus far, very well and easy.
But notice, although the continuum is cohesive, it can be divided.
The course of time is divided by the present into past and future. The line is
divided by the point in “left of” and “right of” it. The area is divided by a whole
line. This, too, is self-evident and obvious.
For the future we register:

The continuum can be divided.

And it is also evident, what the continuum divides is inside the continuum. The
present is a moment in time. The point, the area are limits within the continuum. In
short, what divides the continuum belongs to it.
The Continuum and Why It Does Not Consist of Points 41

How Do Continuum and Point Interact?

Now the question: What is the proper relation between point and continuum?
Clearly, the point belongs to the continuum. The moment belongs to the course
of time, etc. But the other way round? Do only points make the continuum?
Maybe you will answer this question with “Obviously!”. What else should exist
within the continuum?
But it is not that simple! Caution: point and continuum differ from each other in
an essential aspect: the continuum can be divided, as we have just pondered upon—
but the point cannot be divided.

A “point” is that which has no part.

This is the definition of “point”. It is already in Euclid. For him it is the first
definition, the very first sentence at all.
But what has no parts, clearly cannot be divided.
Therefore, continuum and point are essentially different: the continuum can be
divided; the point cannot.
What follows from this?

The Continuum Does Not Consist of Points

The pair of concepts “part/whole” was already the subject of the earliest and
most elementary aspects of philosophical thought in occidental culture. Early
philosophers used these notions to speculate about the metaphysical nature of the
world. One of the bedrocks of these ideas is the following truism:

“Part” can only be what has the essential qualities of the whole.

This principle remained unchallenged in Western philosophy up to the beginning


of the twentieth century, or until the development of set-theory.
As long as we accept this foundational axiom of philosophical thought (at least
for Western culture), we arrive at the result above and have to conclude the following
theorem:

The continuum does not consist of points (or “nows”).

Let us recapitulate the proof! It is made up of four steps:


Fact 1. The continuum can be divided.
Fact 2. The point cannot be divided.
42 4 Indivisible: An Old Notion (Or, What Is the Continuum Made of?)

Fact 3. The continuum has a quality which the point does not have: divisibility.
This quality is an essential one because for the continuum it is essential that it
can be divided. Everything which is extended in space or persisting in time can
be divided.
Fact 4. Consequently, the point cannot be “part” of the continuum—End of proof!
We have shown conclusively that: “The continuum does not consist of points (or
nows)”. It was not terribly difficult to prove our initial statement! Or was it?
For thousands of years this way of thought has been accepted within occidental
culture until, some 150 years ago, it became outdated. It was held no longer suitable
for the new times and thrown onto the rubbish heap. We shall return to this later,
starting from Chap. 12. However, one fact can already be mentioned here: the later
rise of set-theory quickly did away with this age-old way of thinking.
But we are not done yet! There are still some further developments of analysis
and, most importantly, a short retrospective regarding the late Middle Ages, to which
we will turn now.

The Indivisible

We remember: Leibniz, as late as 1692, spoke of his theory as “our analysis of


indivisibles”. This Latin-based name was taken over into the English language; in
German it sounds a little arcane today. As long as Latin remained the language of
scholars (and theologians), such notions were widely accepted.

Thomas Aquinas

Thomas Aquinas (c1225–74) was a philosopher and theologian and one of the most
famous and influential scholastics of the late Middle Ages. Aquinas used the notion
“indivisible” when he spoke of a point or an instant of the spatial or temporal
continuum: the indivisible is the point on the line or the present moment in time.

Nicholas of Cusa

There were other philosophers for whom the concept of the indivisible played an
important role. One of the most illustrious is Nicholas of Cusa (1401–64) whose
Latin name is Nicolaus Cusanus.
Although Leibniz lived about 250 years later than Cusanus, there are many
similarities between the two men. Both were scholars of jurisprudence as well as
diplomats and both were keen travellers. Nicholas of Cusa was a modern scholar of
The Indivisible 43

his times. Just as Leibniz much later, Cusanus was a pioneer of new thought. One of
his revolutionary mottoes was: “Man is the measure of all things!” In reawakening
this doctrine (usually attributed to Protagoras) he opposed the pious tradition of his
contemporaries.
He formulated a principle of thought which is called “Doctrine of Coincidence”.
Somewhat shortened, it says: reason is the wholeness of those opposites (including
contradictions!), which are incompatible to our understanding. At once we are
reminded of Leibniz’ Law of Continuity (p. 34).
Those mathematicians who think that a wholeness of contradictions is an evil
trick may be referred to the quotation of a contemporary philosopher (who is a great
authority on Cusanus), Kurt Flasch:
Who once got acquainted with the contradiction that our thinking is, will grasp that the Law
of Noncontradiction cannot be a philosophical criterion of truthfulness.

Because, says Flasch:


Thinking is rest as well as motion; both are its qualities; who tries to make distinctions
between them, in order to get rid of the contradiction, damages the elementariness of
thinking.

(Caution: This “elementariness” does not mean “simplicity” but is to be under-


stood as the Leibnizian designation: “elementariness” means “without parts”—
p. 21.)
Let me present at least one sentence from Nicholas of Cusa on the indivisibles,
to be found in his Conjectures written ca. 1440/44:
However, reason is of such a lucid nature that it grasps, so to speak, the whole sphere in its
indivisible center.

The “indivisible centre” is the “puncto centrali indivisibili”.

Buonaventura Cavalieri

The (younger) contemporary of Galilei, Buonaventura Cavalieri (1598?–1647),


made the “indivisible” a principal notion of his mathematical theory of the
calculation of areas. This theory is explained in two books. Cavalieri came too early
and could not make use of the language of formulae which Descartes was to publish
in 1637, and consequently, his writings are not easily understood by us.
However, the historian of mathematics Kirsti Andersen took on the challenge of
thoroughly deciphering Cavalieri’s writings. Her representation will be my source
in what follows.

1. In a letter to Galilei from 2nd October 1634, Cavalieri wrote clearly: “I


absolutely do not declare to compose the continuum from indivisibles”.
2. It is Cavalieri’s principal idea to compare areas and to draw conclusions from
these comparisons regarding the magnitudes of those areas.
44 4 Indivisible: An Old Notion (Or, What Is the Continuum Made of?)

Fig. 4.1 Cavalieri: straight


and oblique traverse
(Exercitationes 1647, p. 15)

He traverses a “ruler” through two areas and compares what happens. While
doing so he is interested in a concept invented by him and called “all the lines”. (In
case of an area it is the line which is the decisive “indivisible”.)
The line I K above is the ruler. The above plane moves downwards. Then the
two hatched rectangles signify “all the lines”, the rectangle KM in the case of the
“straight traverse” and the rectangle KO in case of the “oblique traverse”. And these
two collections “all the lines” were considered by Cavalieri to be equal (Fig. 4.1):

OKM (l)straight traverse = OKO (l)oblique traverse

Of course, this does not mean that the areas of KM and KO are equal. The
reason is that the “traverse” of the “ruler” differs in both cases: in case of KM it is
“straight”, but in case of KO it is “oblique”. It is eminently important to compare
both “traverses” with each other, that is to say, their ratio.
Nowadays we describe this ratio with the help of the sine of the angle of
inclination of the rectangle KO. Cavalieri does not do this.
It is crucial that Cavalieri does not say, the collection “all the lines” makes up the
area. This would be nonsense. (We will prove this below!) Instead, Cavalieri takes
the ratio of two of those collections:

OKM (l) : OKO (l) or more general OF1 (l) : OF2 (l) .

He compares only this ratio to the ratio of the considered areas and that only in
case both areas belong to the same plane. Consequently, the case of the “oblique
traverse” is excluded. Then he has:

OF1 (l) : OF2 (l) = F1 : F2 .

It follows that the ratio of the collections “all the lines” of the two surfaces F1 and
F2 is the same as the ratio of their areas.
Cavalieri proves this last equality in all detail, i.e. according to Euclid’s stan-
dards.
The Indivisible 45

We will not go into these details here and just accept the result of Kirsti
Andersen’s research: by this means Cavalieri succeeded, with all mathematical
rigour, in calculating some intricately formed areas. (We quietly pass over the fact
that he had to make up his mind anew when faced with differently shaped areas.
Leibniz enabled us to deal with this much better.)
Our only aim was to present the principle invented by Cavalieri. The picture
shows the uninteresting case where the areas of two rectangles are to be determined.
In this case we obviously do not need a new method, for we know: this area is length
times width. Cavalieri’s method is only of interest if more complicated areas are to
be found.

Evangelista Torricelli

Another contemporary of Galilei and Cavalieri is Evangelista Torricelli (1608–47).


At first, Torricelli refused to accept Cavalieri’s method; but later on he was thrilled
by it and thus subscribed to it.
However, Torricelli misunderstood Cavalieri’s method. For, contrary to him,
he asserted that Cavalieri’s strange collections “all the lines” were thought to
be identical to the area. In other words, Torricelli used just that equation which
Cavalieri painstakingly shunned, namely

F = OF (l) (banned equation!)

And as Torricelli intensely promulgated this “adopted” method as Cavalieri’s—be it


out of ignorance or on purpose— he brought Cavalieri into discredit.

Why Are “All the Lines” Not the Area?

That the last equation is nonsense (and, therefore, was rightfully avoided by
Cavalieri) can easily be proved.
We take a rectangle ABCD with its diagonal and conclude:

D C

E F

A B
G
46 4 Indivisible: An Old Notion (Or, What Is the Continuum Made of?)

1. Each line EF corresponds to a line F G.


2. Each line EF and each corresponding line F G have the ratio EF : F G, i.e.
the ratio AB : BC.
3. But if we have OF (l) = F, it follows that the areas of the two large triangles
ADC and ABC must have the ratio AB : BC.
4. But they are obviously equal and thus we reach a contradiction!
As the first two statements are established facts, the error must arise in the third
step. Consequently, the equation OF (l) = F must be wrong.
Torricelli’s method is of no use! But the reason is not his handling of “indi-
visibles”. (For Cavalieri relied on indivisibles and got a working method.) Instead,
Torricelli made the wrong usage of indivisibles. Wrong means: he identified “all
indivisibles” with the “area”. Torricelli simply ignored the fact that the continuum
does not consist of indivisibles!
Torricelli’s mistake disappears, if one does not compose the area from lines but
from (very small) pieces of an area instead. The piece EF GG F  E  consists of two
parts with equal areas, divided by F F  . (It is the difference of the larger right angle
AG F  E  and the smaller one AGF E, both halved by a diagonal.) Now, if the sides
of these parts, EE  and GG , are infinitely small, i.e. if they are taken as indivisibles,
then the equality of the two large triangles does follow!

D C

E F
E F

A B
G G

By the way: Torricelli knew about this problem!

Newton’s Method of Fluxions

It would be completely unjustified to give an overview of Leibniz’ calculus and not


to say a word about Isaac Newton (1643–1727), especially as Newton invented his
method about ten years prior to Leibniz.
However, Newton’s formulation of his method was far less clear than Leibniz’.
Besides, it is much more specific than the calculus. Therefore, we will only deal
very briefly with Newton’s method.

Newton’s Method

Newton, too, worked in his manuscripts with “indivisibles”. Usually he called them
“infinitely small lines” and used them to get new results, as a method of invention.
Newton’s Method of Fluxions 47

But in case of proving his results in his publications he carefully tried to avoid
these notions.

An Example

Since 1981, all the working papers of Newton have been published, including those
which he did not wish to be published. Therefore, everybody has the chance today to
witness how he was working: The Mathematical Papers of Isaac Newton, volumes 1
to 8.
In the following I present an example from his originally unpublished papers that
illustrates how Newton truly worked. To make things easier, I simplify Newton’s
example; but his method is preserved. Newton starts with an equation like

x 2 − ax + a 2 = 0 .

Then he continues: let x be a “fluent” quantity with velocity m. During the infinitely
small interval of time o, x will become x + mo. (Because length is velocity times
duration.) In the equation, x + mo may be substituted in place of x:

(x + mo)2 − a(x + mo) + a 2 =


x 2 + 2 · x · mo + (mo)2 − ax − a · mo + a 2 = 0 .

When the terms of the first equation are erased, we get:

2 · x · mo + (mo)2 − a · mo = 0 .

Newton divides all by o and gets

2 · x · m + m2 o − a · m = 0 .

And then he writes boldly:


Since o is supposed to be infinitely small, terms which have it as a factor will be equivalent
to nothing in respect to the others. I, therefore, cast them out and there remains

2·x·m−a·m=0 or more simply: 2x − a = 0 .

Every physicist knows: Newton’s result is correct. But in regard to mathematical


or philosophical standards, Newton’s method leaves much to be desired. At first he
divides by the quantity o—and, therefore, implicitly assumes it to be = 0— because
a division by 0 is forbidden, and thereafter he acts as if o = 0!
According to mathematics or logic, such reasoning cannot be justified; it is
acceptable only by its success.
48 4 Indivisible: An Old Notion (Or, What Is the Continuum Made of?)

Nearly one hundred years were to pass, till mathematics succeeded in rendering
this crooked reasoning logically sound. Today it is known that Leibniz faced the
same problem—and that he solved it unobjectionably with the help of an intricate
idea (pp. 32f).

Fluxions and Fluents

Newton calls his variable quantities “fluxions” and “fluents”. The velocity of the
variable x he describes as “fluxion” and sometimes he wrote this fluxion as “ẋ ”.
Generally speaking, Newton’s first problem is to determine the velocity of
a known variable quantity. Newton’s second problem is the reversal of this: to
determine the distance covered when the velocity of the moving object is known.
It is easy to express this in Leibniz’ own language:
1. If y is given, determine dy
dt .
dy
2. If dt is given, determine y.
This is due to the fact that Leibniz writes down all the involved quantities. (Just
as Descartes then demanded!) Clearly, “velocity” comes along with “time”, and,
therefore, Leibniz explicitly wrote it down as t. But Newton did not!

Newton’s formulae concerning motion do not show time as a variable.

At best, time is hidden deep in the notation “ẋ ”. Each physicist today can
immediately translate this in Leibniz’ language:

dx
ẋ = .
dt
However, Newton has no knowledge of this language and instead writes “m” (as we
have just seen) or “ẋ ”.
This reveals the second shortcoming of Newton’s method: it is conceptually too
restricted.
Newton’s method demands that every variable depends on time. But not each
problem in the world is of this kind. Sometimes, a variable does not depend on time
alone but in addition to some other quantity: temperature, pressure, height, etc.
It happens that one variable depends on two quantities. For Leibniz there is no
difficulty: he simply writes all the variables of the problem down (as was demanded
by Descartes in his particular problems).
Without this approach, Newton gets into severe difficulties: in the case of two
variables he has to trace both of them back to time, then to solve the problem and
finally, he has to eliminate time again.
Literature 49

This is possible. But it is not simple. It is of little surprise then that British
mathematics fell behind in the further development of calculus: Leibniz’ notations
were more general and thus more applicable.

Literature

Andersen, K. (1985). Cavalieri’s method of indivisibles. Archive for History of Exact Sciences, 31,
291–367.
Becker, O. (1954). Grundlagen der Mathematik in geschichtlicher Entwicklung. Freiburg,
München: Karl Alber, 2 1964.
Cavalieri, B. (1647). Exercitationes geometricae sex. Bologna: Iacobi Montji. https://www.e-rara.
ch/zuz/content/titleinfo/14611669. http://dx.doi.org/10.3931/e-rara-53675
de Gandt, F. (1995). Force and geometry in Newton’s Principia. Princeton, NJ: Princeton
University Press.
Flasch, K. (2001). Nicolaus Cusanus. München: C. H. Beck.
Gerhardt, C. I. (Ed.) (1849–1863). Leibnizens mathematischen Schriften, 7 vols, Berlin: Weid-
mannsche Buchhandlung. http://www.archive.org
Leibniz, G. W. (2011). Die mathematischen Zeitschriftenartikel. Ed.: Heinz-Jürgen Heß & Malte-
Ludolff Babin. Hildesheim, Zürich, New York: Georg Olms Verlag.
von Kues, N. (2002). Philosophisch-theologische Werke. Darmstadt: Wissenschaftliche Buchge-
sellschaft.
Wallner, C. R. (1903). Die Wandlungen des Indivisibilienbegriffs von Cavalieri bis Wallis.
Bibliotheca Mathematica. Zeitschrift für Geschichte der mathematischen Wissenschaften. third
series, 4: pp. 28–47.
Whiteside, D. T. (Ed.) (1967–1981). The mathematical papers of Isaac Newton (vols 1–8).
Cambridge: Cambridge University Press.
Chapter 5
Do Infinite Numbers Exist?—An
Unresolved Dispute Between Leibniz and
Johann Bernoulli

A Correspondence

Leibniz and the brothers Jacob and Johann Bernoulli were a dream team. On friendly
terms and encouraging each other, they further developed the system first invented
by Leibniz and turned it into a productive tool for science and technology. In
time, the considerably younger of the brothers, Johann, became Leibniz’ closest
mathematical friend.
Normally, Johann Bernoulli and Leibniz complemented each other. If one of
them made a mistake and the other became aware of it, it went without saying that
it was discussed and corrected.
However, there was one big mathematical issue upon which they were unable to
come to an agreement. The hotly disputed point was: do infinite numbers exist?
This dispute continued for nine months, from June 1698 until February 1699,
alongside various other topics in their correspondence. Nevertheless, these two
celebrated mathematical minds never reached unanimity regarding this issue.
Isn’t this astounding? Two very well harmonizing mathematicians are unable to
settle a mathematical controversy? It is! And, therefore, it is worth taking a closer
look at the nature of their dispute.

The Subject of the Controversy

The subject of this friendly controversy was the endless halving of one:

1 1 1 1 1
, , , , , ... (L)
2 4 8 16 32

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 51


D. D. Spalt, A Brief History of Analysis,
https://doi.org/10.1007/978-3-031-00650-0_5
52 5 Do Infinite Numbers Exist?—An Unresolved Dispute Between Leibniz and. . .

Today such an infinite object is called a “sequence”, in those times it was called
a “series”. But names do not matter and, therefore, we shall choose the modern term
“sequence”.

Harmony

It is clear and undisputed between both opponents that:

This sequence has infinitely many terms.

For, if there were only finitely many terms, there would be a last one. Yet,
according to our rule this term is to be halved—and obviously this would be
possible.
A more sophisticated argument is a proof by contradiction: suppose, we only
have a finite number of terms. These could be counted, let us say we count to m.
Consequently, the last term would be 21m :

1 1 1 1 1 1
, , , , , ... .
2 4 8 16 32 2m
According to the rule there would exist a next term:

1 1 1 1 1 1 1
, , , , , ... , ,
2 4 8 16 32 2m 2m+1
1
and, therefore, contrary to our assumption, 2m would not be the last term! That is
why our assumption must be false.

Johann Bernoulli’s Exciting Position

In his letter of 26 August 1698 Johann Bernoulli sets forth the following exciting
chain of reasoning:
1. This sequence L has infinitely many terms. Therefore, there exist infinitely
many terms.
2. But if there exist infinitely many terms, “infinite” is a number—and conse-
quently there exists a most infinite term.
3. But if there exists one most infinite term then there exist others too.
That is why the sequence of halves is to be written properly containing an
infinitely large number i:
. . . with Different, Even Contrary Consequences 53

1 1 1 1 1 1 1 1
, , , , , ... , , , ... (B)
2 4 8 16 32 i 2i 4i

A completely new idea! As far as we know, nobody ever thought in this way
before!
Johann Bernoulli did not express his idea in this formula, but we have enough
evidence to assume that this is what he thought. (This can be seen in the way his
most prominent pupil, Leonhard Euler, wrote these ideas down much later.—Euler’s
treatment of the matter will be the subject on p. 77.)

Johann Bernoulli’s Prudence

Johann Bernoulli had a good reason not to write formula B down, for he knew
his friend Leibniz to be a philosophical thinker and, as such convinced that infinite
“wholes” do not exist. Therefore, he had to be careful in talking to Leibniz about
“infinitely large” numbers. (“Whole” means not only an aggregate but also a true
object that embodies its parts in a specific way.)

Another Shared (Mathematical) Point of View . . .

To both it was crystal-clear: for pure mathematical reasons a “largest number”


cannot exist.
For if we could call it i and add 1—whoops, we have a larger number! The
assumption “there exists a largest number” has led to a contradiction—consequently,
the assumption must be false: a largest number does not exist.
We record:

There exists no largest number.

. . . with Different, Even Contrary Consequences

From this well established and proven mathematical theorem Leibniz and Johann
Bernoulli draw conclusions which are not only different but even completely
contrary:
54 5 Do Infinite Numbers Exist?—An Unresolved Dispute Between Leibniz and. . .

Leibniz says: Therefore, infinite numbers do not exist at all.


Johann Bernoulli says: infinite numbers do exist—however, not a largest
one but infinitely many, always more.

That means: Leibniz thinks the sequence of all halvings to be (only) L, while for
Johann Bernoulli it is more, namely B.

Johann Bernoulli’s Position in Dispute

Johann Bernoulli was convinced that he could prove his position mathematically.
On 8 November he wrote to Leibniz: “If the infinitely small terms of the sequence
do not exist, then not all terms exist.” That is to say: for Johann Bernoulli only
sequence B is right but not sequence L, because L does not contain all terms.
On 18 November Leibniz countered: “These infinitely small terms are impossi-
ble.”
However, that is strong stuff! Leibniz not only claimed to be right (i.e. the
sequence under consideration has shape L, not B), but moreover that he was able to
prove that he is right.
Unfortunately, Leibniz did not disclose his proof. That’s really a pity! For no
such proof is known to date. (And I venture: it is impossible.)

Johann Bernoulli Argues

On 6 December Johann Bernoulli argued for his mathematical viewpoint as follows:


“If the terms of the sequence are numerically infinite, then necessarily a most infinite
term must exist, and it must be infinitely many times smaller than a finite term.”
Apart from the final sentence it was the same argument he had already advanced
on 26 August.

Leibniz Holds Against

Leibniz did not agree. On 17 December he called the conclusiveness of Johann


Bernoulli’s argument into question: “Moreover, I could not say what hinders us to
think of a sequence in which each term is finite in size, but which has an infinite
number of terms.”
Johann Bernoulli Provides the Evidence for His View 55

Johann Bernoulli Provides the Evidence for His View

On 7 January 1699 Johann Bernoulli formulated his response:


I can prove this theorem easily in the following way: if we have ten terms, there exists
necessarily the tenth, if a hundred, then the hundredth, if a thousand, then the thousandth,
and consequently, if we have numerically infinitely many terms, then there exists the most
infinite (infinitesimal) term.

We see: thus Johann Bernoulli supports his pure assertion by an argument. But
is his reasoning conclusive?

Leibniz Is Doubtful

Leibniz was not convinced and replies on 23 January 1699:


Yet against this, one is allowed to object that in this case the inference from the finite to
the infinite is not conclusive, and if one says there exist infinitely many terms that does not
mean a certain number, instead it is only stated that there exist more than can be expressed
by any limited number.

That is to say, Leibniz was convinced: not every multitude is a number and that
“infinitely (large)” is no number.
Johann Bernoulli must have been totally flabbergasted: Leibniz frequently
reasoned by relying on just this principle, which this time he does not want to accept.
We remember the Law of Continuity (p. 34), which for Leibniz is of great value and
power. Why should this principle not work in this case?
Leibniz, possibly anticipating this, confirmed his doubt by the following addi-
tional statement:
In just the same way I could take it upon myself to argue: among ten numbers one is the
last, and it is also the greatest of them; consequently, in the totality of quantities there is a
last one, and it is the greatest of all, too. Nevertheless, it is my opinion that such a number
comprises a contradiction.

That is clever because it obscures the view. Johann Bernoulli’s argument was:
If there exist n terms, then there exists the nth term. Leibniz plainly side-steps this
reasoning by saying: the “greatest” term of an infinite sequence is a contradiction.
Well, this just avoids Johann Bernoulli’s argument, as he speaks about the nth term,
not about the “greatest”. Thus, Leibniz’ reasoning is correct but does not apply to
Johann Bernoulli’s reasoning.
Summary: on 7 January 1699 Johann Bernoulli proved his view, though not
conclusively. On 23 January 1699 Leibniz argued against Johann Bernoulli, though
not conclusively.
56 5 Do Infinite Numbers Exist?—An Unresolved Dispute Between Leibniz and. . .

The End of This Debate: The Disagreement Continues to Exist

Johann Bernoulli’s proof is not conclusive—nor is Leibniz’ reply. As a consequence,


Johann Bernoulli has no objective reason for a reply and does not respond. (Another
reason might have been his respect for the older man.)
But Leibniz interpreted the silence in his favour and believed he could confirm
his victory. On 21 February 1699 Leibniz wrote:
You do not answer in response to my argument that the logical inference “if there exist
infinitely many terms, consequently, also a most infinite term” is not conclusive.

Did Leibniz really give a reason that “the inference from the finite to the infinite
is not conclusive in this case”?—No, he didn’t; he only obscured the argument.
It seems as if Leibniz was somewhat aware of the weakness of his position, for
he continued:
I do accept an infinite multitude, but this multitude does not constitute a number nor a
uniform whole.
This means nothing else than the existence of more terms than can be signified by any
number, just as there exists the multitude—or the notion—of all numbers; but this multitude
itself is neither a number nor a uniform whole.

Leibniz gave two arguments: (1) Infinity is not a number; (2) and no whole.
The philosophical argument (“no uniform whole”) we will not discuss here.
The mathematical argument is this: Leibniz thinks a “number” to be an object
which could be given a certain ratio to “one”: 51 , or ds , where d is the diagonal and
1− 1 + 1 − 1 +−...
s = 1 the side of the square. (Or: 3 5 7
1 as on p. 27.) Obviously, infinity
does not fit this concept of number.
That is the mathematical reason why Leibniz does not accept “infinity” to be a
number, and this is plainly correct—according to his way of thinking. And he kept
the upper hand in regard to Johann Bernoulli, for his friend did not present another
concept of “number” which includes infinity as a number.
We can conclude:
1. The existence or non-existence of “infinite” numbers cannot simply be deduced
from the notion of “infinity”.
2. Leibniz’ concept of number does not allow for infinite numbers; they do not
exist for him.
So, anybody who wants to stand up to Leibniz has to offer a new and different
notion of “number”.

Looking Ahead

As we are yet to see, the concept of number which we use today was invented only
in 1872, about 200 years after this dispute between Leibniz and Johann Bernoulli
Considering the Real Significance of This Problem: An Inconsistency in the. . . 57

(first attempts were made already in 1838 and 1849 and remained largely unnoticed:
see below, pp. 201f). Its basic precondition, however, was the acceptance of the
mathematical “infinite” in some way, at least as a concept of number; i.e. that it was
mathematically legitimate to accept “infinite” aggregates in mathematics as well-
defined objects.
This modern state of mathematics, which was reached, say, in 1872, is neither the
position of Leibniz nor that of Johann Bernoulli. This new view demands more than
Leibniz was willing to accept, viz to handle “infinite” mathematically legitimized
objects. But this new view also demands less than Johann Bernoulli presupposed
for his mathematics, for it does not create “infinite” natural numbers.

Considering the Real Significance of This Problem: An


Inconsistency in the Actual Mathematical Thinking

Decimal Numbers Today: Like Johann Bernoulli Then

This debate is not horribly theoretical; at least not for mathematicians. What is the
number π = 3.14159 26535 8979 . . . about?
We do have a completely precise definition of this number: it is the ratio of the
semi-circle to the radius of the circle. But do we have it also as a decimal number?
This decimal number has infinitely many digits, but they are not determined by a
general rule—although they are well-defined, each of them.
Are we consequently and according to its decimal representation allowed to
accept the number π as a well-defined, definite mathematical object?
Leibniz must deny this, while Johann Bernoulli agrees. Leibniz had to accept that
each digit of this decimal number can be determined—but this does not determine
the whole number, the aggregate of its digits. Johann Bernoulli thought this to be too
sophisticated; in his view (see above) the whole number is determined. (We realize:
the notion “determined” is not determined clearly and distinctly.)
Nowadays we are used to accepting Johann Bernoulli’s view. (Some minorities
do not.)

Natural Numbers Today: Like Leibniz Then

However, if we change from the decimal representation of the number π to the


totality {1 , 2 , 3 , . . .} of the natural numbers, Johann Bernoulli can find there
“infinitely large” numbers too: {1 , 2 , 3 , . . . , i , i + 1 , i + 2 , . . . }. But Leibniz
denies the existence of those passionately.
Yet today, in the case of natural numbers, we stick to Leibniz’ view.
58 5 Do Infinite Numbers Exist?—An Unresolved Dispute Between Leibniz and. . .

Upshot: Anything Goes in Today’s Mathematics!

Leibniz and Johann Bernoulli clearly had opposite views concerning the “infi-
nite” numbers: Leibniz denied their existence and even believed them to be
“impossible”—whereas Johann Bernoulli took them as self-evident and easy to
manage.
However, today’s mainstream of mathematics casually embraces both views:
• The existence of decimal numbers and their unrestricted usage is general
accepted—which is the mathematical view of Johann Bernoulli. (Smaller
circles, e.g. the constructivists, do not agree.)
• Concerning the natural numbers matters are opposite: today’s mainstream
refuses to take “infinite” natural numbers into consideration—just as Leibniz
did. (Smaller circles choose the opposite position, last but not least the
proponents of nonstandard-analysis.)
The mainstream is the historically grown view. We realize: taken in total, the
mathematical thinking of today is not consistent. Not even mathematics! (Of course,
the expression “in total” is taken here not in the sense of Leibniz’ “whole”.)

Literature

Buchenau, A. (1904–06). Gottfried Wilhelm Leibniz – Hauptschriften zur Philosophie, 2 vols, ed.:
Ernst Cassirer. Vols 107, 108 of the series Philosophische Bibliothek, Hamburg: Felix Meiner,
3 1966.

Gerhardt, C. I. (Ed.) (1849–1863). Leibnizens mathematische Schriften. (7 vols). Berlin: Weid-


mannsche Buchhandlung. http://www.archive.org
Chapter 6
Johann Bernoulli’s Rules for
Differentials—What Does “Equal”
Mean?

Johann Bernoulli’s Rules for Differentials—Part 1:


Preparation

Review of Leibniz’ Ideas

Earlier we have seen how Leibniz developed the geometrical notion “differential”
(pp. 32f). The Leibnizian differential dy is a line segment, where its length is
determined as the ratio to a given (finite) line segment with the help of a limiting
process (in the figure on p. 33).
This calculation with geometrical objects relies on Leibniz’ Law of Continuity.
The decisive aspect is to apply this law to ratios (or fractions) (see calculation on
p. 35).

Johann Bernoulli Generalizes

Johann Bernoulli threw off the shackles of Geometry. Descartes was discharged, or
more exactly, he was declared outdated. The Leibnizian differential calculus was
about geometrical variables x, y, dx , dy etc. Johann Bernoulli, in dispensing with
Geometry, deals more generally with “variable” quantities.
(We have seen that Leibniz had already used geometrical quantities as variables:
p. 36. With respect to this, Johann Bernoulli offers nothing new.)

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 59


D. D. Spalt, A Brief History of Analysis,
https://doi.org/10.1007/978-3-031-00650-0_6
60 6 Johann Bernoulli’s Rules for Differentials—What Does “Equal” Mean?

From Leibniz’ Law of Continuity to Johann Bernoulli’s First


Postulate

As Johann Bernoulli dispenses with geometrical concepts, he no longer has Leibniz’


Law of Continuity (see p. 35) at his disposal. Clearly, he needs a replacement.
And indeed, he has an alternative. He sets down three postulates. The first goes
as follows:
1. Postulate.
A quantity which is increased or decreased by an infinitely smaller quantity is neither
increased nor decreased.

A surprise will remain, even if we read very carefully: a quantity which is


changed, should not be changed? Is there a greater contradiction? Mathematics must
be free from contradictions!—What is Johann Bernoulli thinking about?
This is a very hard nut to crack. We need an intermediate stage.
The two other postulates concern geometry, and so we will ignore them.

What Does “Equal” Mean?

The Evident Facts

In mathematics “equal” does not mean “identical” or “the same”. Of course not, for
otherwise we could not get off the ground: identities (like 4 = 4 or a = a) are of no
benefit to mathematics.
Mathematics exists because of equations which are NOT identities.
The mathematical “equal” means, more exactly, “equal in a well-defined respect”.

When Descartes writes “x = 3” that means: “The root of the equation is 3.”
That is: the “root of this equation” and the number “3” do relate to each other.—
Or possibly: “This line segment has length 3.” That is: the “length of this line
segment” and the number “3” do relate (in these circumstances).—It depends on the
concrete circumstances: are equations to be solved? Are geometrical figures to be
constructed? Often the problem is to assign a number to specific geometrical objects
(e.g. a line segment) or in algebra (e.g. an unknown of an equation). Or it is about
different signifiers for the same thing, which are set “equal”, e.g. (a + b)2 and y.
Descartes does not write equations like “3 = 3” or “x = x”, for they are of no use
for his mathematics. (In logic, things are different; but logic is not under discussion
here.)
What Does “Equal” Mean? 61

What Johann Bernoulli’s First Postulate Is All About

Now back to our puzzle: Johann Bernoulli’s First Postulate. We may possibly guess
what he means. It should be read like this:
If the quantity x is increased or decreased by the quantity dx (i.e. x ± dx ) than this may be
treated—sometimes—like x.

Or more shortly:
x ± dx is in some respect equal to x.

Remarkable and important is the following: Johann Bernoulli does NOT write
this as an equality! And that is sensible, because such an equation would read:

x ± dx = x . (forbidden equation!)

Why should this equation be forbidden?—Simply, because it implies conclusively


and according to the general laws of calculation:

± dx = 0 . (false equation!)

But this must absolutely be avoided! For, if dx = 0 the entire calculus was to
collapse: calculations with added (or subtracted) noughts are fruitless, just like the
use of identities. The first boils down to the second: if dx = 0 it follows x ± dx =
x ± 0 = x, and the First Postulate simply comes down to x = x. Nothing of any
use.

How This Could Be Written

So far we have understood Johann Bernoulli’s First Postulate to be an equation in a


certain sense (“equal in some respect”) and at the same time NOT to be an equation
in the usual sense.
What are we to do about this?
Since Descartes, mathematics has needed three more centuries for the idea to
arise that it is possible to use several DIFFERENT equalities at the same time besides
each other (preliminary Weierstraß: p. 177; then Cantor: p. 195; the exception
proves the rule: p. 120).
These equations must of course be distinguished. Especially, it is not permissible
to designate the new, second equality introduced in Johann Bernoulli’s First
Postulate by the old “=”. This we just examined.

Consequently, we choose another equality sign.


62 6 Johann Bernoulli’s Rules for Differentials—What Does “Equal” Mean?

Which one? In principle the shape of the sign does not matter. We could write
“=∗ ”, or “=2 ”, or “≈”, or anything else.—Let’s agree on “≈”. Consequently, from
now on we shall not write “x ± dx = x” (for this is wrong!) but

x ± dx ≈ x

in Johann Bernoulli’s First Postulate. Then we will get no problems. (We may read
this as: “is approximately equal”.)

Johann Bernoulli’s First Postulate is, in modern notation:


x ± dx ≈ x.

For us this is not a huge step, but it was within mathematics of the time and,
indeed, well beyond: it took about three centuries for this success. (Naturally, there
were hidden forerunners: see p. 120.)

What Is This Huge Step About?

Descartes tought us the pure formal equation. Mathematics of the twentieth century
(more precisely: structural mathematics—the Bourbaki group and its followers)
brought us the idea of using several distinct signs for equality besides each other.
Consequently we have been able to write really intricate equations like

(x + dx )2 − x 2 = x 2 + 2x dx + dx 2 − x 2
= 2x dx + dx 2 = (2x + dx ) dx
= (x + [x + dx ]) dx
≈ (x + x) dx
= 2x dx .

For the transition from line three to four we used Johann Bernoulli’s First
Postulate—visible through the other equality sign; and only there.
We record:
Mathematics developed in such a way in the twentieth century that more (i.e. different)
equality relations could be used together.

Those other “equalities” (besides =) are called “equivalences”; instead of talking


about “other equations” one speaks, more pompously, about “equivalence relations”.
A foreign term does not make things better or easier. At face value, an equivalence
relation is another equality, another relation of the form “equal in respect to . . . ”.
Johann Bernoulli’s Rules for Differentials—Part 2: Execution 63

The Equalities Must Be Consistent

If one is to calculate with both equalities together, as we just did, those must be
consistent. (We should have checked this already!) Consistency in this case means:
one of the equalities must imply the other. And only one! This is met here:

If a = b then also a ≈ b. (Of course, the opposite is not valid!)

In this case, the second of these two “equals” is called “finer” than the other: ≈
is “finer” than =.—And naturally, = is called “coarser” than ≈.

Johann Bernoulli’s Rules for Differentials—Part 2: Execution

The power of the “differentials” is: their calculation is subject to laws. Oddly
enough, these laws are called “rules”.
Leibniz had already laid down these rules and proved them. Of course, he argued
in geometric terms, and he did this meticulously. We passed over this procedure.
Instead, we look at Johann Bernoulli: which were his Rules for Differentials?
And how did he prove them?

Rules 1 and 2: Addition and Subtraction

Johann Bernoulli writes:


Rule 1
The differential of a sum is the sum of the differentials of each single summand.

His proof goes as follows:


The differential of the quantity a + x is dx , because a denotes a “constant”, as is assumed
now and in the following.
To wit, if a + 0 and x + e are added, the sum is a + x + e; if the subtrahend a + x is
deducted, the remainder becomes e = dx . Q. E. D.

(in both meanings of the phrase: which was to have been proved and that a
statement which was to be proved, finally had been proved).
There is not much to say about this proof. Because a is assumed to be a
“constant” it does not change, and consequently da = 0. On the contrary x changes:
dx = 0 (for otherwise all would be nothing: d(a + x) = da + dx = 0 + 0 = 0).
Regarding our detour to the twentieth century we notice one aspect: Johann
Bernoulli does not write any equality sign, except at the end, where it is completely
unnecessary.
64 6 Johann Bernoulli’s Rules for Differentials—What Does “Equal” Mean?

Today we are used to writing this proof as a sequence of equalities, and we only
need the common equal sign (we do not need the First Postulate:)

d(a + x) = [(a + da ) + (x + dx )] − [a + x] = da + dx = 0 + dx = dx .

The same idea applies if both summands are “variables”: d(x + y) = dx + dy .


This does not need to be elaborated here.
It is the same with subtraction: d(x − y) = dx − dy .

Rules 3 and 4: Multiplication and Division

We have already treated the special case d(x 2 ) = (x + dx )2 − x 2 above, where we


used two different equality signs! Let us simply add: Johann Bernoulli presents yet
again a proof without writing one single formal equation, except, as before, at the
end, where it is completely superfluous.
Thereafter Johann Bernoulli shows: the differential of x 3 is equal to 3 · x 2 · dx ;
again without writing any equation.
Johann Bernoulli proves the general case next, where he, for the first time, does
a sloppy job, because he actually writes the result as an equation:

d(x p ) = px p−1 dx . (false equation!)

Correct would be:

d(x p ) ≈ px p−1 dx ,

but this was not part of mathematics during the eighteenth (or even during the
nineteenth) century.
When dealing with the rule for division, interestingly Johann Bernoulli writes the
decisive step as follows:
ey − f x ey − f x y dx − x dy
= (following Postulate 1) =
yy + fy yy yy

(that is to say: besides e = dx he has f = dy ). Today we would write this:


dx · y − dy · x dx · y − dy · x y · dx − x · dy
≈ = .
y · y + dy · y y·y y·y
For our “≈”, Johann Bernoulli wrote “= (following Postulate 1)” and thus he
actually made use of another sign for equality!
The First Book Containing the Rules for Differentials Stems from de l’Hospital 65

Rule 5: Roots

Finally, Johann Bernoulli deduces his Rule 5 for roots. Written correctly, it is:

√  dx
d p
x ≈ √
p p−1
.
p x

Needless to say: Johann Bernoulli writes again the incorrect sign “=”.

The First Book Containing the Rules for Differentials Stems


from de l’Hospital

Johann Bernoulli did not publish the “Rules for Differentials” himself but taught the
French nobleman Marquis de de l’Hospital. (Of course, he was paid for his teach-
ing.) As a consequence of this, this nobleman, in 1696, printed the first textbook
on the differential calculus: Analyse des infiniment petits pour l’intelligence des
lignes courbes; under his own name, because he was the author. However, at least
in the preface he named Johann Bernoulli for having been his teacher. Naturally, de
l’Hospital was only interested in geometry (curves) and not in algebra—the latter
being Johann Bernoulli’s invention.
(Actually, in the early 1920s a copy of a manuscript was found in the library of
Basel University which most certainly stems from Johann Bernoulli’s lecture notes
for de l’Hospital.)
The Marquis de de l’Hospital was an eminently astute scholar. In his book the
Rules for Differentials are scrupulously presented and proved—and that without the
use of equality signs! Even when his teacher Johann Bernoulli was careless in his
manuscript (as it was not written for being printed!) de l’Hospital argues correctly.
This does him credit.

A Precursor of de l’Hospital’s Book!

The story above is not complete(ly correct). One year earlier than de l’Hospital’s
book, in the year 1695, a book by Bernard Nieuwentijt had appeared in Amsterdam:
Analysis infinitorum, i.e. “Analysis of the infinite”. In its last chapter (starting on
p. 278) this author presents Leibniz’ Rules for Differentials, too. He had come across
them in the Acta eruditorum, the first article in which Leibniz was to present his
invention in 1684 (p. 32).
Especially interesting is the following: Nieuwentijt does not only state these rules
but also justifies them, whereas Leibniz in his article had given no justifications of
his rules! Nieuwentijt was thus forced to find such justifications by himself. The
66 6 Johann Bernoulli’s Rules for Differentials—What Does “Equal” Mean?

problem: Nieuwentijt’s way of thinking was fundamentally different from that of


Leibniz.

An Unsuitable Justification of the Rules for Differentials

How does Nieuwentijt prove the product rule then? At first, Nieuwentijt proceeds
in the usual manner: like Johann Bernoulli (both followed the English Isaac Barrow
who invented this idea), he calls the increase of x shortly e and that of y shortly f
and states:

(x + e)(y + f ) − xy = xf + ye + ef .

In the last expression Nieuwentijt simply omitted ef —and arrived at the product
rule:

d(xy) = x dy + y dx .

What was his explanation for simply ignoring ef ? He argued completely


differently from Leibniz!
In the very first definition of his book, Nieuwentijt states the principle of a strong
empiricist philosophy: he solely took such quantities as given which do not have
a magnitude that exceeds human imagination. Consequently, for Nieuwentijt those
quantities which are too small for human imagination are just nought.
In Leibniz, Nieuwentijt reads that the differentials are not nought. But if the
differentials are diminishing further—as it is the case when they are multiplied:
ef = dx · dy —Nieuwentijt decided this product to be necessarily zero. All this
follows Nieuwentijt’s philosophy and proofs given at the start of his book; he simply
refers to that.—End of Nieuwentijt’s proof of the product rule.
Nieuwentijt endeavoured to create a kind of mathematics which included only
such quantities which do not exceed human imagination. But a thinking like this
cannot cope with differentials.
Actually, single differentials are already “infinitely small” quantities. For Leibniz
this meant: they get smaller than any given quantity. That is why it is arguable
whether or not differentials fit to “the limits of human intelligence” at all.
If not, they have to be taken to be nought—and then there is no differential
calculus. But in case they (just about) fit within the “limits of human imagination”,
undeniably their products (and other, even smaller quantities) are nought—and again
a differential calculus like that of Leibniz becomes impossible.
History is, as often, full of surprises and contradictions: the first book of Leibniz’
Rules for Differentials, together with their justifications, was presented within a
system of notions that renders it absurd. Luckily, de l’Hospital’s book appeared the
following year and offered a reasonable and accurate presentation.
Literature 67

Literature

Leibniz, G. W. (1684). Nova methodus pro maximis et minimis, itemque tangentibus, quae
nec fractas, nec irrationales quantitates moratur, & singulare pro illis calculi genus. Acta
eruditorum, pp. 467–473. cited after: Leibniz 2011.
Leibniz, G. W. (2011). Die mathematischen Zeitschriftenartikel. Ed.: Heinz-Jürgen Heß & Malte-
Ludolff Babin. Hildesheim, Zürich, New York: Georg Olms Verlag.
Marquis de l’Hospital 1696. Analyse des infiniment petits pour l’intelligence des lignes courbes.
Paris: François Montalant, 2 1716. http://gallica.bnf.fr/ark:/12148/bpt6k205444w, http://www.
archive.org/details/infinimentpetits1716lhos00uoft
Nieuwentijt, B. (1695). Analysis infinitorum, seu curvilineorum proprietates ex polygonorum
natura deductae. Amsterdam: Joannem Wolters. http://dx.doi.org/10.3931/e-rara-10853
Schafheitlin, P. (1924). Die Differentialrechnung von Johann Bernoulli aus dem Jahre 1691/92,
Vorlesungen über Differentialrechnung. Leipzig: Akademische Verlagsgesellschaft.
Chapter 7
Euler and the Absolute Reign of Formal
Calculation

The Absolute Monarch of Eighteenth Century Mathematics

Leonhard Euler (1707–83) dominated mathematics during the eighteenth century.


The royal homes of this enduring absolute sovereign of mathematics were the
scientific Academy in St. Petersburg (the new capital of the newly restructured
Russian state)—from 1731 till 1741 and again from 1766—and in between, the
Berlin Academy. (It had been Leibniz’ suggestion to cover Europe with a network
of scientific academies.) Euler never gave a lecture at university.
Leonhard Euler left behind the most comprehensive collection of mathematical
writings known today. They are documented in more than a hundred large sized,
bulky volumes. This is the more astounding as he went completely blind in 1766.
Subsequently he just dictated his treatises. This shows an exceptional mathematician
endowed with an exceptional memory.
From the age of 13, Euler studied mathematics for five years at the university of
Basel under the tutelage of Johann Bernoulli—and this besides his main subject of
theology. The very busy Johann Bernoulli recommended certain works to his student
Euler, but found no time to lecture him privately. However, on Saturdays Euler was
allowed to put questions to Johann Bernoulli. Being very respectful, Euler always
tried to ask as little as possible in order not to bother Johann Bernoulli unnecessarily.

The Invention of the Principal Notion of Analysis: “Function”

Euler made function the principal notion of analysis.

Leibniz had already used the word “function”, but not in a specific, well-defined
way (see p. 40). Only Johann Bernoulli provided a precise meaning for this word.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 69


D. D. Spalt, A Brief History of Analysis,
https://doi.org/10.1007/978-3-031-00650-0_7
70 7 Euler and the Absolute Reign of Formal Calculation

In a treatise, printed in 1718 (i.e. two years before Euler started being supervised by
him) he stated:
One calls function of a variable quantity any quantity which is arbitrarily composed from
variable quantities and constants.

Johann Bernoulli had detached the foundational concepts of Leibniz’ theory from
geometry. Instead, he dealt more generally with “variable quantities” like “abscis-
sas”, “ordinates” etc. (see p. 59). Consequently, he built from these “quantities” a
new concept: the “function”. The “quantities” are to compose a “function”.
His disciple Leonhard Euler accepted this idea. In the year 1748 Euler’s
Introductio in analysin infinitorum (Introduction to the analysis of the infinite)
appeared, the first textbook on analysis. For quite a few decades it set the standards
for the entire field of analysis.

The Components of Functions: Quantities

What Is a Quantity?

Johann Bernoulli did not disclose what a non-geometrical “quantity” ought to be.
Euler is hardly more informative. Casually, Euler once came up with the statement
that “each quantity can be increased or decreased endlessly”, thereby expressing his
conviction that this is due to the nature of a quantity. (By the way, this definition
is not contained in his textbook on analysis but in that of the differential calculus,
which appeared in 1755.)
This example shows: Euler was not a gifted philosopher. The citation names
properties of a “quantity” but does not state, what a quantity ought to be exactly.

The First Kinds of Quantities. Euler’s Characterization of


Quantities Is Insufficient

Consequently, we do not know what Euler took “quantities” to be. What we do


know is: Euler conceived a “quantity” to be something which can be “increased or
decreased endlessly”.
All the more surprised we then read the very first definition of his textbook of
analysis:
A CONSTANT quantity is a definite quantity which always keeps the same value.

The concepts “quantity” and “constant quantity” are definitions of Euler. Log-
ically, a “constant quantity” is a “quantity”. But Euler, contrarily, demands that a
“constant quantity” should lack exactly those properties which he named as a char-
The Components of Functions: Quantities 71

acterization of a “quantity”—viz that it can be “increased or diminished endlessly”.


Euler is caught in a logical error: there should be no “constant quantities”.
After all, Euler set these concepts himself. Therefore we are left with the
somewhat unfriendly conclusion:
With the help of his notation “constant quantity” Euler circumnavigates his failure to
define a general concept of quantity.

However, our criticism is of no help for our understanding of Euler. Euler just
ignored this fundamental flaw of his concepts and argued as if it did not exist. We are
committed to check whether this defect of Euler’s concepts has any (mathematical)
consequences. (It does not seem so.)
A “constant quantity” always retains “the same value”. We would rather not
ask Euler what a “value” ought to be. Instead, we will confine ourselves to the
observation of what Euler used as values. First of all he takes the “numbers.” But
there is more to come—we should take care!

The Second Kind of Quantities

Besides the “constant quantities” (which are those that retain “always the same
value”) there exists a second kind of quantities. This is stated by Euler in the second
definition of his textbook of analysis:
A VARIABLE QUANTITY is an indefinite or general quantity which can take any definite
value whatsoever.

Consequently, “variable” means for Euler: indefinite or general. Euler takes


the “variable” quantity as the general quantity. Consequently the “constant” (or:
“definite”) quantity is for him a special quantity. To sum up: a “variable” quantity
takes all the “constant” (or: “definite”) quantities.
Following Descartes, Euler denotes the variable quantities with the last letters of
the alphabet, sometimes written small: z, y, x etc., sometimes not: Z, Y , X, etc. (see
p. 75).

Euler’s Algebraic Concept of Function

Now we turn to Euler’s definition of that notion of analysis, which has been central
ever since:
A function of a variable quantity is a variable quantity which is described (represented) by
an analytical expression which is in some way or other composed from the variable quantity
and numbers or from constant quantities.

To be meticulous: the above is not the exact formulation of Euler, but if one
strives for philosophical precision it amounts to that.
72 7 Euler and the Absolute Reign of Formal Calculation

Instead of “analytical expression”, following a proposal from Rüdiger Thiele, one


can more precisely say: “calculatory expression” (“Rechenausdruck”). Its meaning
is what was invented by Descartes: a formula composed from letters (in twofold
signification: definite or indefinite quantities), numbers and signs of calculation.
Examples for such expressions are:

a + 3z; az − 4zz; az + b aa − 4zz; cz etc.

These calculatory expressions are not yet the “functions” but only denote, represent
them. The “calculatory expression” is Euler’s specification of what his teacher
Johann Bernoulli called “composed from”.

Simple but Important Consequences from Euler’s Notion of


Function

Let us notice three simple but important consequences to be drawn from Euler’s
notion of function.
1. Whether a calculatory expression really denotes a “function” has to be checked
in each particular case. For example: if we let c = 1 in the calculatory expression
cz we get 1z . But 1z is always 1 whatever the value of the variable quantity z will
be. Consequently, the calculatory expression 1z is not a “variable” quantity but a
“definite” one, and therefore no function! The calculatory expression 1z does not
designate a “function” (but a “definite” quantity). Similarly for the calculatory
expression z0 : whichever value we assign to z, we always get z0 = 1. (Except
for the case z = 0 for we have 00 = 0—but even an exception from the rule
does not matter.) That is why Euler does not take 1z and z0 as functions but as
constant quantities—We conclude:
Following Euler, not every calculatory expression denotes a function!

2. Euler explicitly demands that a “variable” quantity takes “all definite values
whatsoever”. To illustrate this, he explains the calculatory expression in detail

9 − zz .

As under the square root sign we have something not greater than 9, this
calculatory expression seems to denote numbers not greater than 3. But this
consideration is wrong! The “variable” quantity is to take “all definite values
whatsoever”—and for Euler this means: all imaginable numbers; √ and therefore
also, e.g. the
√ (not “actual” but) at least imaginable number 4 −1! Yet, if you
take z = 4 −1, you get
The Components of Functions: Quantities 73


√  √ 2  √ 2 √
9 − zz = 9 − 4 −1 = 9 − 42 · −1 = 9 − (−16) = 25 = 5 > 3 .

9 − z2 = u is equivalent to 9 = u2 + z2 , which is an expression where z and
u have the same status. As z shall denote all “definite values”—and takes even
all “imaginable”
√ numbers—, u shall do so, too. The function which is denoted
as u = 9 − z2 thus takes all numbers, which are “imaginable” according to
Euler. √
Today, numbers with the component −1 are called “complex” numbers. As Eu-
ler expressly demands to include those numbers in the domain for his functions,
we can state in modern terminology:
In principle, Euler admits complex functions.

3. Finally, a very important and for us surprising view:


Euler does not distinguish between finite and infinite calculatory expressions.

Surely a surprise—but if inspected more closely, not unfounded. No doubt, it


would be possible to consider the number calculated by Leibniz (p. 23):

1− 1
3 + 1
5 − 1
7 + 1
9 − +... = π
4

to be something special, but this peculiarity is not solely founded on the infinite
character of the calculatory expression. This is shown by the calculatory expression
on the left which is also infinite

1+ 1
2 + 1
4 + 1
8 + ... = 2,

for it equals a completely common and in no sense peculiar number.


The equation

1
1 + x + x2 + x3 + x4 + . . . = , (∗)
1−x

shows even more clearly that a distinction between finite and infinite calculatory
expressions is meaningless.
This infinite series emerges from the usual procedure of division if applied to the
calculatory expression instead of two numbers, i.e. akin to long algebraic division:

1−x )1
1−x
x
74 7 Euler and the Absolute Reign of Formal Calculation

In detail: if we divide 1 by the difference 1 − x we at first obtain 1 with the rest x


(because from 1 there is the product 1 · (1 − x) to be subtracted: 1 − (1 − x) = x):
x
1 : (1 − x) = 1 + .
1−x
Check:
x 1−x x 1
1+ = + =
1−x 1−x 1−x 1−x
and consequently the equation before is correct. Continuing this process we get:
1 x
=1+
1−x 1−x
1 x2
=1+x+
1−x 1−x
1 x3
= 1 + x + x2 +
1−x 1−x
1 x4
= 1 + x + x2 + x3 +
1−x 1−x
etc.

How Did Euler Denote Functions?

At the beginning but usually also later, Euler denotes a general “function” of a
“variable” quantity x (or y or√z) by the respective capital letter: X (or Y or Z).
For example, he writes “Z = 9 − zz”.
Sometimes he uses the letter “f ” and appends the variable quantity: for Euler,
“f x” indicates a function of the variable quantity x. In these cases Euler does not
write brackets, as we do today—we write “f (x)”. Such brackets are used by Eu-
ler only if the respective variable quantity is composed, e.g. if it is a sum of two
quantities: “f (x + y)”.
For Euler “ x” denotes a variable quantity.

A Standard Form for Functions

Euler proves in his differential calculus that each function of a variable quantity can
be presented according to the general scheme:
A Daring Calculation with Infinite Numbers 75

1 1 1
b + X(x − a) − P (x − a)2 + Q(x − a)3 − R(x − a)4
2 2·3 2·3·4
1
+ S(x − a)5 − + etc.
2·3·4·5
where a is any definite (finite) value; and also b, X, P , Q, R, S etc. are definite
values which have to be calculated by means of differential calculus from the given
function.
(We realize that Euler is not consistent in his notations, for strictly speaking the
capital letters X, P , Q, R, S should designate functions, whereas in this case they
designate values.)
For example b is the value of the given function for x = a (as all other summands
are then equal to zero). X is the value of what we call today the “first derivative” of
the given function, again at the value x = a. (If you know differential calculus, you
see this at once.) Today, this is called the “Taylor series” of the given function.
In particular cases this sum can have a finite number of terms.

Our Problems with This Theorem of Euler

From today’s point of view we have some objections to the validity of this theorem.
One can say: in some cases (i.e. for some values a) the equation is false!
But such a judgement does not hold for Euler. For Euler it is clear (as in the
previous case, where he does not count z0 to be a “function”): the exception proves
the rule! Euler’s theorems are valid in general, for his “variable” quantities are
“indefinite” or “general” quantities. In particular cases there might be an exception.
Euler’s analysis focuses on “variables” (not on “values”), and, consequently, so do his
theorems.

A Daring Calculation with Infinite Numbers

Euler, in his textbook on analysis from 1748, offers a completely surprising kind
of calculation. Its result is still undisputed to this day. However, the manner in
which Euler reaches his result is judged questionable, problematic, or even wrong
in today’s mathematics. Let’s take a look!
76 7 Euler and the Absolute Reign of Formal Calculation

From the Powers of Ten to the Exponential Quantity

The way we usually write numbers is founded on the powers of 10; the decimal
system. It is the reason why we can write every number in standard index form.
The value of a digit is determined by its place, its distance from and location to the
decimal point.

3; 30 ; 300 ; ... ; 0.3 ; 0.03 ; ...

This is how the exponential quantity

10z .

works. The larger the value of the exponent z the larger is the value of the
exponential quantity.
Needless to say, Euler took the most general perspective and represented the
number 10 by some “constant” quantity a:

az .

We limit ourselves to the analogy of a to 10 and take only a > 1.

The Exponential Function

Now we turn to Euler’s startling way of dealing with the exponential function. He
says:
1. We have a 0 = 1.
2. If the exponent increases, so does the value of the number.
3. If the exponent only increases very little—more precisely: only “infinitely
little”—then the value of the number also increases only very little, more
precisely: only “infinitely little”.
Euler writes this in a very clever way:

aω = 1 + ψ .

ω as well as ψ denote two infinitely small numbers. Next, Euler connects them:
4. ψ = k · ω.
Consequently, he gets

aω = 1 + k · ω .
A Daring Calculation with Infinite Numbers 77

The arbitrary exponent (z) of a z is chosen to be an infinitely small value (z = ω);


and a fundamental insight into the nature of the exponential function is expressed
by the form of its value, i.e. by 1 + k · ω.
Finally, we need one more idea and the rest will simply be mathematical routine.
5. We have to return from the infinitely small exponent ω to the finite exponent z.
This we achieve by multiplying ω by an infinitely large number i: z = ω · i.
Because of (a ω )i = a ω·i , we get

a z = a ω·i = (1 + k · ω)i .

The rest is mathematical routine. On the right we have a formula which we can
develop according to the Binomial Theorem as an infinite series:

i i(i − 1) 2 2 i(i − 1)(i − 2) 3 3


a ωi = 1 + kω + k ω + k ω + ···
1 1·2 1·2·3

Through rearranging z = ω · i, we obtain ω = zi and through cancelling in each


fraction the i (e.g. in the second summand: 1i k · zi = 11 kz etc.), we have

1 1(i − 1) 2 2 1(i − 1)(i − 2) 3 3


a z = 1 + kz + k z + k z
1 1 · 2i 1 · 2i · 3i
1(i − 1)(i − 2)(i − 3) 4 4
+ k z + ···
1 · 2i · 3i · 4i
And then comes the final clue, Euler’s trick of cancellation: each fraction has the
infinitely large number i as often in the numerator as in its denominator. As i − 1
is nearly equal to i (for i is infinitely large) and the same with i − 2 etc., Euler
keeps cancelling i against i − 1, i against i − 2 and so forth, and only the simplified
expression

kz k 2 z2 k 3 z3 k 4 z4
az = 1 + + + + + ···
1 1·2 1·2·3 1·2·3·4
remains. If we chose a special value for k, we fix the value of a. The a which results
from k = 1 we often call “Euler’s number” today and denote it by “e”. So we get

z z2 z3 z4
ez = 1 + + + + + ···
1 1·2 1·2·3 1·2·3·4

Even the mathematical layman might find this comprehensible: this infinite series
does not cause any problem. More precisely, it has for any value of z (clearly, we
only speak about finite values!) a definite value which can, at least in principle, be
calculated. For, the denominators of the fractions on the right increase very quickly
78 7 Euler and the Absolute Reign of Formal Calculation

and will surpass any value of z whatsoever. Once the denominators “overtake” the
numerators, the fractions will decrease rapidly and will become arbitrarily small.
Since Leibniz we know: quantities with “arbitrarily small” changes are alright,
because we may calculate them to any degree of accuracy.
Let me add a technical detail. Unfortunately, the “rapidly” above is important! If
the summands decrease too slowly—take e.g. 1+ 12 + 13 + 14 + 15 +. . .—then this will
not be enough to provide the series with a finite value. The considerations above are
not sufficiently precise: it matters how rapidly the remaining terms diminish.
(Surely, the sum of the n terms n+1 1
+ n+2
1
+ . . . + 2n
1
will be greater as if we only
take the smallest, i.e. the last one, n times—and consequently greater than n · 2n 1
; on
1
p. 86 we calculate this in more detail.) However, the latter is 2 ! In other words: if
we proceed in the harmonic series 1 + 12 + 13 + 14 + 15 + . . . far enough, the terms
again and again add up to more than 12 ; therefore the sum of this series is greater
than 1 + 12 + 12 + . . . and, consequently, not finite.)

The Ingenious Trick—Or: Euler’s Cheat

However, in Euler’s calculation there is a bold step! Euler cancels i against i − 1;


than this and i against i − 2; than all this and i against i − 3, and so forth.
But this is false! For, it equates (a) i − 1 and i, which means: −1 = 0; (b) and
in addition i − 2 and i, which means: −2 = 0; (c) and just as much −3 = 0 and so
forth! These equations are not only false but they also contradict each other!
Well, you may say: that is true—yet these errors are infinitely small! Today, we
may write this in the following way (as i is infinitely large):

i−1≈i, i−2≈i, i−3≈i etc.

and all is well.


By no means! For Euler has to concede: his trick of reduction produces an error
in every fraction—although an infinitely small one. Nevertheless, there are infinitely
many fractions and, consequently, infinitely many such errors! And the integral
calculus teaches how to add infinitely many infinitely small amounts to a finite value
(we followed Leibniz’ calculation on pp. 29f).
Outcome: Euler’s calculation is not false. (Here we must trust the mathematical
specialists: if one assigns any value to x, the values on both sides of Euler’s
equation completely coincide.) However, the equation is not conclusively proved.
An argument is missing that in this case infinitely many infinitely small errors do
not add up to a finite amount.
Today’s readers may take comfort from the following. Let k = 1. The general
n
term αn (z) = 1(i−1)(i−2)···(i−(n−1))
1·2i···ni zn is smaller than the general term βn (z) = zn! of
the ez -series, taken absolute values. Though, the last series is known to converge—
take the ratio test. Consequently, also the series of αn (z) converges for finite z, let’s
Euler’s Concepts of Numbers 79

say to A(z). As both series


are converging we

get for finite ε >

m0 a finite m0
such


m

that if m > m0 we have:

αn (z) − A(z)

< ε as well as

βn (z) − ez

< ε.
0 0
m
However, both partial sums having only finitely many summands are (like their
0
summands) infinitely close and, consequently, their difference is certainly < ε.
Hence |A(z)−ez | < 3ε and therefore it follows that Euler’s error is indeed infinitely
small: A(z) ≈ ez . Nevertheless, remember: Euler does not have ANY concept of
“convergence” in his textbook on analysis!

Euler’s Concepts of Numbers

Leonhard Euler performed an inconceivable amount of calculations (by hand!),


surely more than any other mathematician has ever done. He calculated with
numbers as well as with formulas. One of his calculations with formulas we have
just seen. We shall not present any of his calculations with numbers, for this is not
part of any consideration of the foundational concepts of analysis, which are at stake
here.
It has already been stated that Euler was not a gifted philosopher. Therefore,
we should not be astonished that Euler keeps silent in regard of the essence of
number and its concept. This topic he handles pragmatically. He is not afraid
to found mathematical laws by recourse to everyday life. (Philosophically this
is unacceptable.) For example, he justifies the multiplication rules for negative
numbers (among them: “minus times plus is minus”) by interpreting negative
numbers as debt. The most difficult of these rules (“minus times minus is plus”)
he just decrees: “I say, the opposite must be the answer.”—as minus times plus is
already minus.
The “irrational numbers”
√ √Euler √ presents plainly under the aspect of required
calculations, as roots: 2, 3, 5 (irregularities
√ √ of this kind do not √ bother him
at all) etc. Besides, there exist of course 3 2, 3 3 etc., clearly also 4 2 . . . and so
on and so forth. A huge zoo, without a general view or theory. For the sake of
calculating that
√ will do.
Similarly, −a may be subjected to the usual rules of calculation without further
ado. The arithmetic genius and “calculator”, Euler, is amply satisfied with this.
Euler’s treatment of the exponential function has shown that he also uses infinite
(natural) numbers. Euler’s teacher Johann Bernoulli was reluctant, at least in his
exchange with Leibniz, to denote such numbers in formulas and confined himself
to mention them only verbally (pp. 53f). Euler has no such qualms—and was also
able to show which fruitful mathematical results could be gained with the help of
formulas.
80 7 Euler and the Absolute Reign of Formal Calculation

Infinitely large natural numbers are plainly objects of mathematical thinking


which cannot easily be characterized clearly and distinctly. This we have seen in
the dispute between Leibniz and Johann Bernoulli (pp. 52f).
Euler’s often happy-go-lucky, really careless, presentation of “infinite” numbers
is therefore completely disastrous and plainly awful. So he actually writes:
Consequently, one can justly say that 1 divided by 0 indicates an infinitely large number or
∞.

And, shortly after, he goes even further and writes:


Because 10 indicates an infinitely great number and, undoubtedly, 20 is two times as great, it
is obvious that even an infinitely great number may be enlarged 2 times.

His intention is obvious: 2i > i (“i” for Latin “infinitus”). However, not even Euler
should get away with such an inconsistent way of thinking like 10 < 20 , because
when taking reciprocals we inevitably get nonsense: 01 > 02 or 0 > 0, which is
absolutely forbidden in mathematics.
Even the greatest geniuses have their weaknesses! Thus we are left with the
following judgement:
Euler did not deliver any contribution to the clarification of the notion of number.

Anyway, one positive outcome we get from this completely crazy passage in
Euler: obviously, besides the “numbers” there exists something else for him that
stands in as a “value”—and that is ∞. Thus, let us finish this consideration with the
optimistic result:
Euler counts, besides the “numbers”, also ∞ as a “value”.

Analysis Without Continuity and Convergence

Hopefully, today’s mathematicians are wondering about the fact that Euler’s analysis
gets along without those two concepts which are of the greatest importance in
modern analysis: continuity and convergence. Although nobody will have read every
textbook of analysis in use, we can be fairly sure: each of them deals with these two
notions. Yet, Euler’s textbook does not.
This is one of the rather rare cases when the question “why?” can be answered
easily and clearly. The question reads: why does Euler get along in his analysis
without the notions of continuity and convergence? Answer: because both concepts,
convergence and continuity, deal with mathematical facts which are about values—
nonetheless, singular values are not the subjects of Euler’s analysis; they are
considered only in their entirety.
We remember: Euler demands from a “variable” quantity to take “all definite
values without exception” (p. 71). His analysis deals at most with a “constant”
Continuity According to Euler 81

quantity (and, of course, with “numbers”) as “values;” but not as today, with values
as constituent and foundational notions.
Euler’s analysis ignores the notion of “value.”
Euler delivers no arguments for that.
With this, his analysis is fundamentally simpler than today’s analysis. It lacks
these two notions, “continuity” and “convergence”—and they are both, each in
itself, somewhat intricate. “Continuity” and “convergence” are not concepts suitable
for small talk. “Continuity” and “convergence” are really difficult technical terms.
These logically somewhat tricky notions were only theorized by the mathematicians
of the nineteenth century. We will trace this story later (pp. 103f).
To sum up Euler’s analysis: its only central concept is “function”. It is based
on numbers and the concept of “quantity”. The last exists in twofold manner: as
“constant” and as “variable”.
In the nineteenth century, the concepts “continuity” and “convergence” were
introduced and became more and more fruitful. That is why we finally ask what
Euler thought about these concepts. For yes, he himself actually used them!

Continuity According to Euler

The “calculatory expression” is the central notion in Euler’s analysis. Therefore, it


makes sense to transfer this concept to other mathematical fields, e.g. to geometry.
Euler did this in the following way:
The principal object of geometry in the eighteenth century is no longer—
as once for Euclid—the straight line, but, far more generally, the “curved line.”
Descartes invented the pure formula and showed how to denote certain straight
lines (“abscissas” and “ordinates”) by “x” and “y” and from these, together with
the signs denoting arithmetic operations, he built a “formula”. Leibniz made these
“x” and “y” fluid. He overthrew Descartes’ decree to understand them as definite
“numbers” and, instead, took them as “variable quantities”.
Subsequently Euler fused both, Descartes’ formula and the then principal notion
of geometry, the “curved lines”. He decided: the proper curved lines are those which
can be described by a “formula.” As the formula, more precisely, the “calculatory
expression”, is Euler’s favourite concept, he introduced the definition:
A “curved line” is called “continuous” if it can be described by a calculatory expression.

An example is: (x + 2)2 + 3.


In Euler’s conceptual world this means: a “continuous” curved line is the
geometrical image of a “function”, for a “calculatory expression” describes in
Euler’s eyes a “function” which is therefore its algebraic form.
Without being explicitly stated, Euler’s emphasis on this definition of “continu-
ity” is that it must be one single calculatory expression. As we already stressed,
Euler demands a “variable” (that is the “x” in the formula) to take “all definite
82 7 Euler and the Absolute Reign of Formal Calculation

Fig. 7.1 Example 1 y

Fig. 7.2 Example 2 y

x
2

values without exception”. Consequently, it was for Euler unthinkable to supply a


“function” with a RESTRICTED domain of values and to say, e.g.: “take the function
y = 2x + 1 for all values of x, starting from x = 3.” or even: “take y = 1 for all
x  0—by which we nowadays describe the half-line” (Fig. 7.1).
Similarly, we describe the following broken straight line (Fig. 7.2)

by

1
2 (x + 2) if −2x <0
y=
1 if 0x,
and we are using two “calculatory expressions”:
1
(x + 2) and 1.
2
For us the picture shows the geometric image of a “function”—and more: the image
of a (in today’s understanding) “continuous” function!—Euler would not have
understood this. For him, not one of the two examples can be described by “one
calculatory expression” and therefore they are not “continuous” curved lines for
him.
And Euler is right! None of these lines can be described by one single calculatory
expression; additional particular distinctions are needed, e.g. (in)equalities or a
principal change of the notion of function which renders it possible to implement the
method of “Fourier analysis”. However, according to the notions of his own analysis,
Euler is absolutely right: according to his own concept of function the lines given in
these examples are not “continuous”.

Euler’s Second Notion of Function

As only some curved lines can be described by a single calculatory expression, it


is impossible for Euler to deal with all curved lines in this way. Of course, he was
aware of this.
Euler’s Second Notion of Function 83

Fig. 7.3 Euler’s second notion of function (1748/1759)

Nevertheless, Euler is the last to feel restricted because of mathematical concepts.


On the contrary, whenever he is faced with the problem of describing any “curved
lines” mathematically—e.g. those of Fig. 7.3—, he changes his view. In these cases
Euler does not fall back on his teacher Johann Bernoulli’s notion of function as a
calculatory expression. Instead, he describes the “curved line” in just the same way
as Leibniz had already done: he expresses the dependency of the “ordinate” on the
“abscissa”. This way, Euler specified a second concept of function and even created
a special notion for it:
As it is generally possible to represent the ordinate of some curved line by f : z, where
z is the abscissa, we take A M B to be the curved line, where the ordinates P M are the
functions[1] of √
the abscissas A P and denoted with help of the sign f : z, the result of which
is P M = f : (t b).

More clearly: the respective “abscissa” z, e.g. A P, refers to the “ordinate”


PM. These “ordinates”, altogether, constitute the “geometrical” function; given as
formula: P M = f : (z).
Of course, this way Euler is out of the woods. For he is now able to deal with
each “curved line” and make it a “function” in this sense, denoted by “f : (·)”.
(The “ordinates” are represented through their end points: n  , b, m, A, M, n, B, N,
a, M  .)
This approach—to resort to a second notion of function depending on the
problem—is unusual. Of course, scrupulous readers of Euler became aware of
this and set out to study Euler’s second (geometrical) concept of function, as for
instance the Alsatian mathematician Louis Arbogast (1759–1803) in his treatise
Dissertation on the nature of arbitrary functions, which enter into the integrals of
partial differential equations (original in French, in the year 1791). Later this fell
into oblivion and today’s books on the history of analysis do not mention it. (Only
Prof Dr Thomas Sonar in his book with the pompous, but completely misleading
title 3000 years of analysis (in German), hints at it, but unfortunately he forgot to
name his source.)

1 Emphasis added.
84 7 Euler and the Absolute Reign of Formal Calculation

Outlook

Nowadays the concept of continuity is fundamental to analysis. It is a notion of


analysis, not of geometry. Nonetheless, Euler’s concept of continuity refers to
“curved lines” and is therefore a property of geometrical entities. In other words:
The concept of “continuity” which is common today, must differ from that which Euler
invented.

Its definition and its development will be shown later (pp. 107f, pp.125f).

Convergence According to Euler

The two notions of “convergence” and “sum” (of a series, more precisely “true
sum”) are for Euler strongly related. We will explain them separately.
Initially, the notion of “convergence” refers to infinite sums, also called “series”.
We remember one of Leibniz’ problems (p. 23):

1− 1
3 + 1
5 − 1
7 + 1
9 − +...

In the case of this infinite sum Leibniz had shown by way of a completely accurate
proof that it has a finite value. We may say: he proved that it “converges.”
Euler came up with a spectacular deduction of the exponential series

z2 z3 z4
ez = 1 + z
1 + 1·2 + 1·2·3 + 1·2·3·4 + ···

(p. 77). Indeed, this series always gives a finite value, whichever (finite) value z
might take. (And we have also seen that an infinite series not necessarily has a finite
sum.)
Nowadays we say that a series is “convergent” only if it has a finite sum,
otherwise it is not.

Convergence and Divergence

For Euler, things are different! Euler, too, coined a notion of convergence—but one
which differs from the one we use today. In a treatise titled On divergent series Euler
presents the following definition:
We call a series convergent if its terms keep getting smaller and finally disappear
completely, like with

1+ 1
2 + 1
4 + 1
8 + 1
16 + 1
32 + etc.,
Convergence According to Euler 85

whose sum is indubitably = 2.


For, the more terms one adds, the more one approaches the 2. If one adds just a hundred
terms, the error to the two is already extremely small: it is a fraction with numerator 1 and
a denominator with thirty digits.

Unfortunately, this definition is not as clear as it should be. Euler seems to argue
that it is important that a series, which “converges” according to his definition must
have a finite sum, too. But this is not the case!
One has to read his words very carefully. If so, one realizes:
1. It consists of two sentences.
2. The first sentence begins with an explanation, the definition; which is
3. immediately followed by an example, introduced by the words “like with”.
4. The second sentence refers exclusively to the given example.
Consequently, the end of the first sentence together with the second sentence do
not relate to the definition of “convergence” any more. They are a consequence of
it, or properly speaking: an example thereof.
In short, Euler’s definition reads:
We call series “convergent” if their terms continually decrease and finally disappear
completely.

That this painstaking interpretation leads to the correct result becomes immedi-
ately clear when we read Euler’s definition of the opposite of “convergence”, namely
his definition of “divergence”:
Series are called divergent if their terms do not tend to nothing, but either never decrease
below any given limit or increase to infinity. Such are

1 + 1 + 1 + 1 + 1 + etc.

as well as

1 + 2 + 3 + 4 + 5 + 6 + etc.,

wherein many terms added give an always greater sum.

Again, Euler comes up immediately with examples. However, this time the first
sentence is unambiguous and clearly separated from the second one.
To memorize the difference between Euler’s “convergence” and ours, we state:

In Euler’s conceptual world the harmonic series

1+ 1
2 + 1
3 + 1
4 + 1
5 + ...

is “convergent”, although it has no finite sum!

In more detail: whatever natural number k one proposes, it will always be


possible to designate a number p which gives:
86 7 Euler and the Absolute Reign of Formal Calculation

1+ 1
2 + 1
3 + 1
4 + 1
5 + ... + 1
p−2 + 1
p−1 + 1
p >k.

Even if we have k = 10100 this will be possible—although this number very likely
exceeds the numbers of the atoms in the universe! And the same will be true for
k = 101000 , a number which is clearly beyond good and evil.
As already explained on p. 78, the decisive point is that you will always find
another section of the series which adds up to more than 12 :

1
n+1 + 1
n+2 + ... + 1
2n >n· 1
2n = 1
2 .

And 12 , added often enough (namely 2k times), increases to any given natural
number k (for: 2k · 12 = k)! For the p above you may choose the number 22(k−1)
to be on the safe side. Realization: the number of the terms to be added, increases
exponentially, like 2x . Notwithstanding, one always obtains the fraction 12 —isn’t it
strange? And moreover, the particular summands decrease below any given limit—
actually an entirely astonishing wonder of calculation!

The True Sum

Euler is an outright consequent formalist! Of course, he knows about the signifi-


cance of the sum of the terms of a series, especially in the case of a “convergent”
series. In case of the series 1 + 12 + 13 + 14 + 15 + . . . this sum is infinite, for which
Euler writes “∞”.
But Euler is not satisfied with that! Referring to the example of the series ∗ from
p. 73 and the associated calculations, Euler introduces in his Differential calculus
the notion of “true sum”:
We will avoid these difficulties and apparent contradictions completely, if we give the word
sum another, unconventional meaning. So we will call the expression which is developed to
give rise to the infinite series, the true[2] sum of this series.

That means: the “true sum” of the series 1 + x + x 2 + x 3 + . . . is 1−x


1
. And that
is completely independent of the value of x! Also for the case x = 1 (as for Euler
holds 10 = ∞), and even for x = 2:

1 + 2 + 22 + 23 + 24 + . . . = 1 + 2 + 4 + 8 + 16 + . . . = 1
1−2 = 1
−1 = −1 .

According to Euler, the formula triumphs over common arithmetic! He even accepts
the “true sum” of infinitely many numbers, which increase beyond any limit, to
be −1. It seems impossible to show a greater degree of disdain for numbers and
common arithmetic!

2 This word added, for evidently Euler forgot it here.


Convergence According to Euler 87

Of course, in being a human “calculator” , Euler takes the “common” sum of


1+2+22 +23 +. . . to be nothing else than ∞. However, his theory called “Algebraic
Analysis” is something very different. It does not deal with common arithmetic, with
the “common” sum—but with the “true” sum!
For Euler the truth is in the formula, in the calculatory expression

The cited formulation from the Differential Calculus is not a one-off but it is
Euler’s true mathematical conviction. Twenty years later he repeated this view in its
entirety in his Complete Manual to Algebra and explained it, this time in the German
n+1
language by referring to the “remainders” x1−x that originate from the procedure
akin to long division (see p. 74):3
At the first glance this looks unfitting.
But we should notice that if one wants to stop somewhere in the above series, one needs
always to put an additional fraction there as well.
Thus, if we stop for instance at 64, we have to attach to

1 + 2 + 4 + 8 + 16 + 32 + 64

−1 = −128, wherefrom 127 − 128 emerges, which is −1.


128
also this fraction: 1−2 that is 128
If one moves on without end the fraction falls away, but one also never stands still.

To illustrate this further, let’s take a second example for a “true sum” in Euler’s
sense:
1
1 − 2x + 3x 2 − 4x 3 + 5x 4 − + . . . = (†)
(1 + x)2

The “true sum” of the series on the left is the calculatory expression on the right.
(Those who are used to some university maths will easily deduce this result with
the help of the method of undetermined coefficients—others will readily accept this
without proof; maybe they could just check the result?)
If we have the calculatory expression, it is not too difficult to write down the
corresponding power series. However, there is no strategy for the reverse process:
to find the “true sum” of a series is always a mathematical problem.
Euler faces a problem, which nowadays we confront only in case of “converging”
series: even if we have proved the convergence—in today’s meaning!—of a series,
we do not know its value by a long way! For Euler it is just the same: it is not easy to
find the “true sum” of a series, i.e. the corresponding calculatory expression. (This
is quite independent of the “convergence” of the series, be it in ours or in Euler’s
meaning.) Today, we are used to calculating the true “value” of a “converging”
series, whereas Euler does not offer a theory about such particular entities as
“values.” He deals with algebra, with the general: with the “formula.”
But one thing seems clear: with his concept of “true sum” Euler for once
prefers a finite calculatory expression to an infinite one. (Of course, it would be

3 as clumsy and old-style as possible.


88 7 Euler and the Absolute Reign of Formal Calculation

truly surprising to find a universal principle among the rules created by the highly
pragmatic Euler.)

To Sum up Euler’s Algebraic Analysis

Euler’s concept of analysis was called “Algebraic Analysis”. This name fits very
well. Our summing up of his theory is this:
The courtly etiquette of the absolutist royal ruler corresponds in mathematics to Euler’s
calculatory expression, the formula. Both etiquette and formula, outperform common-
sensical behaviour and calculations, respectively—be the outcome in the particular case
as absurd or irrational as humanly possible. Both are done according to fixed and rigid
rules and that at any cost, even if the obvious facts speak against it.

(Whoever wants to learn the details about the courtly etiquette in absolutist
monarchies may consult the authoritative work of Norbert Elias The Court Soci-
ety, written in 1969, or more recently the book Das Europa der Könige from
Leonhard Horowski.)

Literature

Arbogast, L. F. A. (1791). Mémoire sur la nature des fonctions arbitraires qui entrent dans les
intégrales des équations aux différentielles partielles. St. Pétersbourg: Académie Impériale des
Sciences.
Bernoulli, J. (1991). Op. CIII. „Remarques sur ce qu’on a donné jusqu’ici de solutions des
Problêmes sur les isoperimetres, avec une nouvelle methode courte & facile de les resoudre sans
calcul, laquelle s’étend aussi à d’autres problêmes qui ont rapport à ceux-là“. [Goldstine (1991).
Die Streitschriften von Jacob und Johann Bernoulli. Ed.: David Speiser. Die gesammelten
Werke der Mathematiker und Physiker der Familie Bernoulli (pp. 527–568). Basel, Boston,
Berlin: Birkhäuser Verlag].
Elias, N. (1969). Die höfische Gesellschaft. Vol. 423 of the series Suhrkamp Taschenbuch
Wissenschaft (Frankfurt am Main: Suhrkamp, 1981).
Euler, L. (1788a). Leonhard Eulers Einleitung in die Analysis des Unendlichen. Erstes Buch.
(Berlin: Carl Matzdorf). http://books.google.de/books?id=VwE3AAAAMAAJ&pg=PP2&ots=
LNs0o_YK04&lr=, German: Johann Andreas Christian Michelsen.
Euler, (1788b). Leonhard Eulers Einleitung in die Analysis des Unendlichen. Zweytes Buch.
(Berlin: Carl Matzdorf). http://books.google.de/books?hl=de&lr=&id=EwI3AAAAMAAJ&
oi=fnd&pg=PA5&ots=zSybL4AxZo&sig=G-Nioin1m_1c5p6-cGI91jpldWs. German: Johann
Andreas Christian Michelsen.
Euler, L. (1828, 1829, 1830). Leonhard Euler’s vollständige Anleitung zur Integralrechnung, 3 vols
(Wien: Carl Gerold), German: Joseph Salomon.
Euler, L. (1836). Leonhard Eulers Einleitung in die Analysis des Unendlichen. second volume,
new unchanged corrected edition. (Berlin: G. Reimer), German: Johann Andreas Christian
Michelsen.
Euler, L. (1983). Einleitung in die Analysis des Unendlichen von Leonhard Euler. (Berlin: Springer,
new edition: Berlin etc: Springer). http://gdz.sub.uni-goettingen.de/no_cache/dms/load/toc/?
IDDOC%=264689, German: H. Maser 1885.
Literature 89

Euler, L. (1745/1748). E 102, Introductio in analysin infinitorum. Tomus secundus. In: A. Speiser
(Ed.), Leonhardi Euleri Opera Omnia. Vol. 9 of series I (Zürich; Leipzig, Berlin: Orell
Füsli; B. G. Teubner, 1945). http://dx.doi.org/10.3931/e-rara-8740, Deutsch: [Euler, L. (1836).
Leonhard Eulers Einleitung in die Analysis des Unendlichen. second volume, new unchanged
corrected edition. (Berlin: G. Reimer), German: Johann Andreas Christian Michelsen].
Euler, L. (1746/1760). E 247, De seriebus divergentibus. In: C. Boehm & G. Faber (Ed.), Leonhardi
Euleri Opera Omnia, vol 14 of series I, pp. 585–617 (Leipzig, Berlin: B. G. Teubner, 1925).
http://math.dartmouth.edu/~euler/
Euler, L. (1748/1750). E 140, Sur la vibration les cordes. In: F. Stüssi (Ed.), Leonhardi Euleri
Opera Omnia, vol 10 of series II, pp. 63–77 (Leipzig, Berlin: B. G. Teubner, 1947). http://
math.dartmouth.edu/~euler/
Euler, L. (1763/1768). E 342, Institutiones calculi integralis. Vol. 1. In: F. Engel & L. Schlesinger
(Eds.), Leonhardi Euleri Opera Omnia. Vol. 11 of series I (Leipzig, Berlin: B. G. Teubner,
1923). http://urn:nbn:de:bvb:12-bsb10053432-6
Euler, L. (1768/1769). E 387, Vollständige Anleitung zur Algebra. In: Heinrich Weber (Ed.),
Leonhardi Euleri Opera Omnia. Vol. 1 of series I (Leipzig, Berlin: B. G. Teubner, 1911). http://
math.dartmouth.edu/~euler/
Euler, L. (1790–1793). Leonhard Euler’s vollständige Anleitung zur Differenzial-Rechung. 3 vols,
from Latin translated into German and annotated. (Berlin, Libau: Lagarde und Friedrich,
1790, 1790, 1793). http://dx.doi.org/10.3931/e-rara-8624, German: Johann Andreas Christian
Michelsen
Goldstine (1991). Die Streitschriften von Jacob und Johann Bernoulli. Ed.: David Speiser. Die
gesammelten Werke der Mathematiker und Physiker der Familie Bernoulli (Basel, Boston,
Berlin: Birkhäuser).
Horowswki, L. (2017). Das Europa der Könige. (Reinbek bei Hamburg: Rowohlt).
Laugwitz, D. (1986). Zahlen und Kontinuum. Eine Einführung in die Infinitesimalmathematik.
(Mannheim: Bibliographisches Institut).
Sonar, T. (2011). 3000 Jahre Analysis. Heidelberg, Dordrecht, London, New York: Springer.
Spalt, D. D. (2011). Welche Funktionsbestimmungen gab Leonhard Euler? Historia Mathematica,
38, 485–505. https://doi.org/10.1016/j.hm.2011.05.001
Thiele, R. (1982). Leonhard Euler. Vol. 56 of the series Biographien hervorragender Naturwis-
senschaftler, Techniker und Mediziner. Leipzig: BSB Teubner Verlagsgesellschaft.
Chapter 8
Emphases in Algebraic Analysis After
Euler

d’Alembert: Philosophical Legitimation of Algebraic Analysis


as Well as His Critique of Euler’s Concept of Function

Jean-Baptiste le Rond d’Alembert (1717–83), another first-rate mathematician, was


a slightly younger contemporary of Euler. He was also active as a philosopher
and involved in politics. His name, together with that of Denis Diderot (1713–
84) personifies the French Encyclopédie, published in the years 1751–80, a work
which greatly influenced the French Enlightenment. (However, d’Alembert left the
scientific editorial staff of the Encyclopédie as early as 1758.)

d’Alembert’s Reflections on the Notion of Quantity

In volume 7 of the Encyclopédie, published in 1757, d’Alembert starts his entry


about Quantity with the following sentence: “It is one of those words, the entire
world believes to have a clear idea of, but which is nevertheless very difficult to
define precisely.”

d’Alembert’s Critique

D’Alembert starts with the very same notion of quantity which we have already
found in Euler: quantity is, what “can be increased or decreased without end”.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 91


D. D. Spalt, A Brief History of Analysis,
https://doi.org/10.1007/978-3-031-00650-0_8
92 8 Emphases in Algebraic Analysis After Euler

At first, d’Alembert explains the importance of the “or” in this definition. Would
it instead read “and”, both zero and infinity would not meet the requirement of
the definition: for zero cannot not be decreased and infinity cannot be increased.
However, d’Alembert thinks it cannot be denied that both are “quantities”.
We realize: d’Alembert’s idea of the infinite is not like Euler’s or Johann
Bernoulli’s, but more like Leibniz’—because Euler as well as Johann Bernoulli
allows himself to increase the infinite (i): i + 1, i + 2, . . . ; in opposition to this,
Leibniz and d’Alembert do not.

d’Alembert’s Notion of Quantity

D’Alembert switches abruptly to his own understanding of “quantity” and declares:


“It appears to me that quantity can be defined well as being something which is
composed of parts.”
This brings d’Alembert back to Euclid, who had declared two thousand years
earlier: “A number is a multitude composed of units.” And: “A number is a part of
a number, the less of the greater, when it measures the greater.”
And by the way, the philosopher d’Alembert now explains the foundational
legitimacy of this Algebraic Analysis:

The quantity[1] exists in each finite being and it expresses itself in an indefinite number
which can only be known and understood with the help of a comparison and in relation
to another homogene magnitude.

Descartes’ universal unity is dismissed and the Law of Homogeneity restored,


albeit with a new meaning.

Assessment: d’Alembert’s Philosophical Legitimation of


Algebraic Analysis

Unfortunately we cannot know d’Alembert’s intention behind this formulation.


Maybe he did not want to say something different from what umpteen generations
of mathematicians have said already before him: only the fixing of a ruler (the
“unit”) allows the measuring of a quantity. However, closer inspection of the matter
(which includes the historical moment of d’Alembert’s writing) gives rise to read
this sentence as the philosophical legitimation of Algebraic Analysis:
It is not about definite values, but about all possible values, without any exception; only
that counts.

This was Euler’s demand, at the beginning of his textbook (p. 79).

1 d’Alembert first uses “quantité” for “quantity” but in the end he uses “grandeur”.
d’Alembert: Philosophical Legitimation of Algebraic Analysis as Well as His. . . 93

d’Alembert’s Critique of Euler’s Notion of Function

For many years Euler and d’Alembert competed for the best mathematical descrip-
tion of the vibrating string. Unable to agree—each of them thought his own answer
to be superior. As so often in everyday life: they talked past each other. Neither of
them understood the other. They had different ideas of the notion of a function.
What Euler understood by “function” in this case, was shown on p. 83, but what
was d’Alembert’s idea?
D’Alembert did not have Johann Bernoulli as his teacher (as Euler did) and was
therefore further removed from the geometrical foundations of analysis. D’Alembert
stuck to the algebraic notion of function, given by Euler: a function is described by
a “calculatory expression”. Nonetheless, it was clear to d’Alembert that this was not
enough to deal with all the “curved lines” encountered in practical problems—but
only with the nice ones: those which Euler had called continuous (p. 81).
D’Alembert tried to liberalize this algebraic notion of function in two steps. His
principal idea was to replace the one calculatory expression by two of them.
1. At first d’Alembert allowed a “function” to be determined by an equation, that is
to say by two calculatory expressions, related by an equal sign.
Unfortunately this idea is of no great help. If we have an unknown on both sides
of the equation it is very doubtful whether we will be able to make y the subject
of the equation. (Since the nineteenth century it has been known that this can
already fail with equations of the fifth degree. During the eighteenth century one
was far more optimistic.) If one does not succeed in isolating the unknown? How
should the “function” be further inspected, if it is not given explicitly?
2. D’Alembert tried to modify the notion of function after 1761 and from 1780 he
became more specific.
His writing is not terribly clear, but two points can be discerned:
a. It is clear that he connects two functions and their “equations” with adjoining
domains to make up one “continuous” (because of the two expressions!)
function. That is to say, he changes his original notion of function decisively
by limiting its scope. For Euler this was unthinkable!
b. He carries over the notion of “continuous” from “curved line” to “function”.
Contrary to modern concepts, d’Alembert calls a compound “function”
“continuous”, if it describes a single “curve” and if these two different
“functions” at their meeting point (i) take the same “value” and (ii) have the
same (one-sided) derived function (increase or decrease of the tangent).
Thus, d’Alembert aims to give the notion of function more flexibility within the
conceptual realm of Algebraic Analysis, because the algebraic notion of function
as demanded by Euler is not very useful in practical circumstances.
94 8 Emphases in Algebraic Analysis After Euler

d’Alembert’s Impulse: Condorcet

If one digs even further into the mathematical sources of the late eighteenth century,
more precisely into the treatise On the continuity of arbitrary functions of Marie
Jean Antoine Nicolas de Caritat, Marquis de Condorcet (1743–94) from 1774, one
gets some evidence that d’Alembert’s last step was inspired by a much younger
contemporary with that illustrious name. However, such subtleties exceed the scope
of this book.

Lagrange: Making Algebra the Sole Foundation of Analysis

The last first-rate mathematician who completely subscribed to the spirit of Euler’s
analysis was Joseph Louis Lagrange (1736–1813). His textbook was published in
1797, the “year V” following the calendar of the revolution, and in 1813 in a second
edition. A German translation by Johann Philipp Grüson (1768–1857) appeared as
soon as 1798 under the title (in German) Theory of analytic functions, wherein the
principles of differential calculus are given, independently from considerations of
the infinitely small or vanishing quantities, the limits or fluxions, and grounded
in Algebraic Analysis. Eventually this text led to the establishment of “Algebraic
Analysis” as a title for the new theory. It is an apt name.

Lagrange’s New Foundation of Analysis: The Base

The lengthy title of the work encapsulates its content and plan. Lagrange aims to
present the theory of functions and the differential calculus as simply as possible—
and, consequently, independently of such complicated notions as “infinitely small”
quantities, “limit” or (as with Newton) “fluxion”. Analysis as easy as possible, a
commendable plan.
Lagrange’s idea is very consequential and can be understood in three steps.
First Step. Lagrange starts at the very beginning, with Euler’s notion of function.
Accordingly to Euler, a “function” is a “variable” quantity, which is described by a
calculatory expression. Lagrange radicalizes this and says:
A function is a calculatory expression.
Following this, any calculatory expression is a “function” for Lagrange—be it
the description of a variable quantity or not.
Here, he differs from Euler. Euler only considers a calculatory expression as a
“function” which describe a changing quantity. Lagrange no longer cares whether
the calculatory expression can take on only one or more values. For him every
calculatory expression is a “function”.
Lagrange: Making Algebra the Sole Foundation of Analysis 95

Let’s recapitulate Euler’s notion of “calculatory expression”.


1. The simplest calculatory expressions are the sums, finite ones like 2 + 3x − 5x 2
or infinite ones like 1 − x + x 2 − x 3 + x 4 − + . . .
2. However, a simple fraction can also be written as an (infinite) sum (see formula ∗
on p. 73):

1
= 1 + x + x2 + x3 + . . .
1−x

or (see formula † on p. 87)

1
= 1 − 2x + 3x 2 − 4x 3 + 5x 4 − + . . .
(1 + x)2

3. Therefore Euler is convinced that every function can be described as such an


(infinite) sum. This he noted in § 59 of his textbook on differential calculus:
Thereupon it should be beyond doubt that every function can be transformed into such
an expression that runs towards the infinite

Ax α + Bx β + Cx γ + Dx δ + · · · , (‡)

wherein the exponents α, β, γ , δ . . . stand for any numbers whatsoever.

Euler chooses Greek letters (alpha, beta, gamma, delta) as exponents. They can
also be arbitrary numbers.

The Idea of Lagrange

To implement his plan, Lagrange—in his second step—seizes on a fact which was
originally advanced by Johann Bernoulli (in the year 1694) and again, very clearly
and repeatedly, by Euler. In his Integral Calculus, Euler restated a theorem, which
he had already proven in his Differential Calculus (and earlier, see p. 75):
dy
Theorem. If y denotes a function of x, which changes to b, if z = a, and if we put dz = P,
dQ
dz = Q, dz = R, dz = S etc., then we obtain the general expression:
dP dR

y = b + P (z − a) − 12 Q(z − a)2 + 16 R(z − a)3

− 1
24 S(z − a)4 + 1
120 T (z − a)5 − etc.

For a better understanding of the main idea we simplify this calculatory


expression. If (z − a) is replaced by x and the differential quotients P , Q, R, etc.
are shortened to 1P = p, − 12 Q = q, 16 R = r, − 24
1
S = s, . . . we get the following
condensed expression:
96 8 Emphases in Algebraic Analysis After Euler

f (a + x) = f (a) + p · x + q · x 2 + r · x 3 + s · x 4 + t · x 5 + etc. (§)

Euler had shown that each function, each calculatory expression can be
represented in this way. The left side of the equation is new, instead of “f (x)” we
now have “f (a + x)”.
It is now Lagrange’s idea—his third step—to prove that the scheme ‡ is nothing
other than Eq. §. In other words, Lagrange proves the following theorem:

Theorem. EACH function f (x) can be written as a calculatory expression of the form §.

The proof of this theorem is Lagrange’s opening to his textbook on analysis.


If Lagrange were really able to prove this theorem, he would stage a coup d’état.
Because of Euler’s theorem which we have cited above, one then was able to deduce
that the coefficient p of the second summand in § is the first differential quotient of
the function (viz dy dz ); the coefficient q of the third summand on the right in § is
the second differential quotient (in Euler: dP 1
dz ) inclusive of the factor 1·2 and the
sign; the coefficient r of the forth summand on the right in § is the third differential
quotient of the function ( dQ 1
dz ) inclusive of the factor 1·2·3 ; etc. In other words: if
Lagrange can prove this theorem he would be able to obtain for every function all
differential quotients. In this case, the differential calculus is founded all at once
and entirely without the use of the commonly applied notions like “infinitely small”
quantities etc.
Lagrange introduces also a new notation which is still used today. Instead of
the “first differential quotient” dy 
dz of the function f Lagrange simply writes “f ”;
instead of the second differential quotient dz he writes “f ” etc. Until today, f  ,
dP 

f  etc. are called the “first”, “second”, . . . “derivatives” of the function f .

A Contemporary Criticism on Lagrange’s Plan

August Leopold Crelle (1780–1855), who does not count as a first-rate mathemati-
cian, edited Lagrange’s textbook in a new German translation in 1823. Crelle saw
fit to launch a fundamental criticism of Lagrange. He wrote:
In my opinion, the proof that the series expansion of an arbitrary function f (a + x) only
contains positive integer powers of the quantity x, is firstly defective or at least very weak
and much too complicated to found the principles of a whole science; and, secondly, I think
such a proof is completely superfluous.

That is strong stuff, and the latter judgement is very noteworthy: why is the proof
of this theorem in Crelle’s eyes “superfluous”?
Crelle makes it easy for himself and says: later in the book it is shown, how it
works—but if you show how it works, you need not prove that it works.
This is obviously self-deception. Of course, Lagrange does not show in his book
how to expand all functions in a series: the book is finite (the original has 296 pages),
Lagrange: Making Algebra the Sole Foundation of Analysis 97

but there are infinitely many functions! And of course, Lagrange (without any doubt
being a first-rate mathematician) did not resort to such kind of argument. He was
sure to have a watertight, general proof for his theorem.

How Does Lagrange Proceed?

Lagrange’s idea that his theorem is correct is simple. It goes as follows: the essential
argument is to show that alpha, beta, gamma, delta, . . . in Euler’s calculatory
expression ‡ (p. 95) can only be taken by positive integers 1, 2, 3, . . . However,
this is self-evident, for (and that’s the point!) in case of an exponent different from
a positive integer the corresponding expression is one-to-many, whereas the given
function, clearly, is one-to-one.—And this is all there is to it!
1 √ √
That is to say, Lagrange’s argument is this: x 2 = x has two value s ( √ 4 = ±2);
√ √
x 3 = x has three values ( 1 has the value 1 and the two values 2 ± 2−3 ) etc.
1 3 1
3

This is a real fact—if one is to accept the “complex” numbers.

The Fundamental Gap in Lagrange’s Proof

Nevertheless, Lagrange’s proof has a gap. This gap is of a fundamental nature and
1
can be described as follows: Lagrange resorts to the argument that the “function” x 2
has two “values”. However, only the following is true: there exist two “values” X
1 √
to be discovered from the instruction x 2 (= x), i.e. X2 = x. (Example: we have
22 = 4 as well as (−2)2 = 4—and consequently two values for X.) But calculating
is one thing—and mathematics another!
Lagrange did not set out to present an elementary calculation, but to prove a
theorem! And within his proof he clearly uses the notion “value of a function”—
1
e.g. by arguing that x 3 has three values. This argument is wrong, if one only
considers “real” functions—as we do today in our√first course at university. Three
real numbers (i.e. those without the component −1), which can be calculated
1 √
following the instruction x 3 = 3 x do not exist.
Well, you might object and say: Lagrange does not restrict himself to real
functions—and, consequently, this objection is irrelevant.
The objection is only true in so far as Lagrange, indeed, did not write a
theory of real functions, but (as we say today:) a theory of complex functions.
Nevertheless, Crelle’s argumentation is too weak, as it only deals with the given
example (restriction to a real analysis) but not the essence. The essence remains:
Nowhere does Lagrange specify the concept “value of a function”.
However, he implicitly uses the concept. This is definitely a deficiency in
mathematical rigour.
98 8 Emphases in Algebraic Analysis After Euler

You may say: what a tiresome thing, “value of a function”—isn’t it obvious, what
this means?
Counter-question: is it really so obvious? What is the “value of the function” for
x
1
in case of x = 0? Or in case of x = ∞? That is to say: what is 10 , what is ∞
1
? And
moreover: what is the value of the function log 0, or tan 2 ? Or of sin ∞? This we
π

do not know beforehand or through calculating, these answers need a theory—and


that means: we are to do some mathematics!
That is why we have to launch some criticism at Lagrange: nowhere do you
declare what you take the notion “value of a function” to be! And it is unacceptable
in mathematics to base a proof on an undefined notion!
Usually, Lagrange is criticized for something quite different. He is accused of
making a technical mathematical error: that he has not taken into consideration that
the representation § does not work in particular cases.
However, this criticism is completely unfounded. A first-rate mathematician such
as Lagrange does not make such a technical mistake. If one reads his book carefully,
one can see that he treated the said technical problem extensively and explained what
he thought about it. Therefore, he knew about this problem—but this aspect does
not touch his theorem and, in particular, it does not refute it. Because, just like with
Euler, Lagrange’s theorem is about “quantities”, but not about particular “values”.
Unfortunately, today’s mathematicians (as well as historians of mathematics) often
do not follow this argumentation.

Literature

d’Alembert, J.-B.R. (1757). Grandeur. In Encyclopédie, ou Dictionnaire Raisonné des Sciences,


des Arts et des Métiers, vol. 7, pp. 855. Paris: Briason, David, Le Breton, Durand.
d’Alembert, J.-B.R. (1761). Recherches sur les vibrations des cordes sonores. In Opuscules
mathématiques, vol. I, pp. 1–64. Paris: David. https://gallica.bnf.fr/ark:/12148/bpt6k62394p.
d’Alembert, J.-B.R. (1771). Quantité. In Encyclopédie, ou Dictionnaire Raisonné des Sciences, des
Arts et des Métiers, vol. 13, p. 655. Paris: Samuel Fauche.
d’Alembert, J.-B.R. (1780). Sur les fonctions discontinues. In Opuscules mathéma-
tiques, vol. VIII, pp. 302–308. Paris: David. https://www.bnf.fr/fr/collections_et_services/
reproductions_document/a.repro_reutilisation_documents.html.
de Condorcet, M. (1771). Sur la détermination des fonctions arbitraires qui entrent dans les
intégrales des équations aux différences partielles. Histoire de l’Académie Royale des Sciences,
Année 1771, pp. 49–74.
Euclid. (1908). The thirteen books of Euclid’s elements. In T. L. Heath, vol. I. Cambridge at the
University Press.
Euler, L. (1828, 1829, 1830). Leonhard Euler’s vollständige Anleitung zur Integralrechnung, 3
vols. Wien: Carl Gerold. German: Joseph Salomon.
Euler, L. (1885). Einleitung in die Analysis des Unendlichen von Leonhard Euler. Berlin: Springer.
Reprint 1983. German: H. Maser. https://gdz.sub.uni-goettingen.de/no_cache/dms/load/toc/?
IDDOC=264689.
Euler, L. (1763/68). E 342 Institutiones calculi integralis, vol 1. In F. Engel, & L. Schlesinger
(Eds.), Leonhardi Euleri Opera Omnia, vol. 11 of series I. Leipzig, Berlin: Teubner, 1923.
https://mdz-nbn-resolving.de/bsb00034024.
Literature 99

Grabiner, J. V. (1981). Changing attitudes toward mathematical rigor: Lagrange and analysis in the
eighteenth and nineteenth centuries. In Epistemological and social problems of the sciences in
the early nineteenth century, pp. 311–330. Dordrecht, Boston, London: D. Reidel.
Lagrange, J. L. (1813). Théorie des Fonctions Analytiques. In Œuvres, publiées
par les soin de J.-A. Serret, vol. 9. Paris: Gauthier-Villars, 1881. Reprint:
Olms, Hildesheim, 1973. https://books.google.de/books?id=XGQSAAAAIAAJ&
pg=PA4&ots=ZgppLb1box&dq=lagrange+22theorie+des+fonctions+analytiques22&
lr=, https://gallica.bnf.fr/ark:/12148/bpt6k86263h/f1.table. Text version (extracts):
https://gallica.bnf.fr/ark:/12148/bpt6k88736g. German: Lagrange 1823.
Lagrange, J. L. (1823). Theorie der analytischen Functionen. J. L. Lagrange’s mathematische
Werke, vol. 1. Berlin: G. Reimer. German: A. L. Crelle.
Chapter 9
Bolzano: The Republican Revolutionary
of Analysis

The Situation

From the Academies to the University

The Algebraic Analysis of the eighteenth century was created by mathematicians


at academies: Euler, d’Alembert, and Lagrange were prominent as members of
scientific academies, but not as university professors. This changed with the growing
industrialization of the economic life and with increasing mechanization at the
turn of the century. Society needed more engineers of all kinds, and, as modern
technology relies on mathematics, higher education for engineers became necessary.
What effect this had on the centre of mathematical development at that time,
Paris, will be shown in the next chapter. First we have to take a detour to Prague and
to the province of Bohemia.

Bolzano: The Public Enemy

Emperor Francis had ordinal number I as Emperor of Austria and ordinal number II
as the last Holy Roman emperor. By a ruling from 24th December 1819 and
notification on 20th January 1820, he dismissed the theologian, philosopher and
mathematician Bernard Bolzano (1781–1848) from his chair of philosophical
theology at the University of Prague. Bolzano was kept under surveillance and
banned from publishing. Aid of the local gentry, among them functionaries of
the Church, prevented the worst and allowed Bolzano to pursue his studies in the
absence of his university position.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 101
D. D. Spalt, A Brief History of Analysis,
https://doi.org/10.1007/978-3-031-00650-0_9
102 9 Bolzano: The Republican Revolutionary of Analysis

Bolzano’s thoughts were considered too dangerous. The Austrian Emperor


wanted to prevent an overthrow of the social order, as had happened in France,
at any price. Bolzano’s sermons as a priest were thought to be socially revolutionary
and the Emperor’s confessor suspected subversion. Had not a priest and Prague
University professor Jan Hus (c1370–1415) inspired ordinary people through his
sermons and made himself an enemy of the authorities some 400 years earlier?
That Bolzano was a first-rate revolutionary thinker is also evident from those
mathematical manuscripts, which he composed in seclusion from the scientific
world and which were finally printed in the late twentieth century. We shall come
back to that.

A New Meaning of Convergence

Today, “convergence” is a principal concept of analysis. And as we know already,


this had not always been the case.

Euler: A Reminder

We have already seen Euler’s understanding of “convergence” (p. 84): a series a1 +


a2 + a3 + . . . “converges” (in Euler’s sense), if the ai continually decrease and get
arbitrary small, as e.g. 1 + 12 + 13 + 14 + . . . (p. 85). However, in Euler’s analysis,
“convergence” did not play any part.

Today

Today things are different—at least in scientific textbooks. There, the notion of
“convergence” (of a series) is often introduced quite early, even before the notion of
“continuity” and even before the techniques of differentiation and integration.
In textbooks for engineers, things may be different, or in cases where the author
aims to present analysis as a very first encounter with mathematics and not as a
prelude to it.
An author of the latter kind is Michael Spivak. His textbook Calculus published
in 1967 comprises 29 chapters and some 500 pages. He presents the concept of
“convergence” as late as in chapter 21, on page 373, only after the introduction
and treatment of “continuity”, “derivative”, and “integral”. It appears that modern
analysis can be constructed without using the concept “convergence”.
A New Meaning of Convergence 103

The Convergence of Sequences: Two Notions

We saw: in the seventeenth and eighteenth century, the mathematicians worked with
series: a1 + a2 + a3 + . . . Today we are more used to begin with sequences—these
are the separate summands of a series in their succession: a1 , a2 , a3 , . . . ; they are
not added.
Today we define the concept “convergence” of such a sequence in two steps (we
read Spivak):
1. “Convergence towards a limit”:
A sequence a1 , a2 , a3 , . . . “converges to” l (in symbols: lim an = l) if for every ε > 0
n→∞
there is a natural number N such that, for all natural numbers n,

if n > N , then |an − l| < ε .

l is called the “limit” of the sequence.

Without worrying about all the details of this definition, one soon gets the
impression: this definition is very similar to that which Euler has given in his
conceptual treatment of series. He had demanded that the an decrease continually
and finally vanish. Speaking with Spivak: Euler demands that the “sequence” of
the an converges to the limit 0: |an | = |an − 0| < ε—written in today’s notation:
lim an = 0.
n→∞

2. “Convergence” without a “limit”:

A sequence a1 , a2 , a3 , . . . “converges” if for every ε > 0 there is a natural number N


such that, for all m and n,

if m , n > N , then |an − am | < ε .

(Today such sequences are usually called “Cauchy sequences”—although


Bolzano was earlier, as we shall see soon.)
This gets really complicated! How should we understand this?—We shall
postpone the answer until we have looked at Bolzano’s approach and carry on.
It is not allowed to define a concept twice! That is why we have to disentangle
these two steps. To begin with:
1. Each of the two formulations in the boxes is a possible definition of our modern
concept “convergence”.
2. The first formulation is more specific and the second more general. (Because the
first formulation has the “limit”, whereas in the second case one must prove that
such a limit l exists at all!)
104 9 Bolzano: The Republican Revolutionary of Analysis

3. Therefore, let us prove that both formulations are equivalent for Bolzano!
(a) The first formulation implies the second: we suppose that |an − l| can be
made less than any given (positive) value, starting from a natural number N.
Consequently, |an − l| as well as |am − l| can be made < 2ε , if only n > N
and m > N. The calculation is simple if we resort to the “triangle inequality”
(|x ± y|  |x| + |y|). Because for n, m > N, we have
ε ε
|an − am | = |an − l + l − am |  |an − l| + |l − am | < + = ε,
2 2
and that was to be proved.
(b) Now the reverse direction. We assume that |an − am | can be made arbitrarily
small from some natural number N onwards (formally, for n > N and
m > N). Where do we get l from without stealing it?—Bolzano has an
idea. He says that l can be “determined as precisely as one wishes”. And he
thinks this to be sufficient. Let us decide that for ourselves!
(i) It is, indeed, possible to determine l with any degree of accuracy.
We only have to determine a suitable ε. For an accuracy of a tenth,
we choose ε = 10 1
. Then it follows from our assumption that there
exists a natural number N such that for n > N and m > N we
have |an − am | < 10 1
. So we simply choose l = an and we obtain
|l − am | < 10 , as required.
1

(ii) Without a conceptual construction of “number”, no better argu-


ment is known so far.
(iii) As we shall see later, not until the year 1872 (or maybe 1849 or
1838) was such a conceptual construction of “number” forthcom-
ing in mathematics.
(iv) That is why Bolzano obtained the best possible result for his time.
In short, the reverse direction is also proved, and that by Bolzano himself in
the year 1817; we shall come to this in the next but one section!

The Convergence of Series Today

In his chapter 22, Spivak defines the concept “convergence” also for series. Spivak
is very precise and up to date. Therefore, he does not say in his definition that the
“series” a1 + a2 + a3 + . . . “converges”, but instead the “sequence” a1 , a2 , a3 , . . .
is “summable” (the meaning is the same):
A New Meaning of Convergence 105

The sequence a1 , a2 , a3 , . . . is “summable” if the sequence s1 , s2 , s3 , . . . converges,


where

sn = a1 + · · · + an .

In this case, lim sn is denoted by


n→∞



an (or, less formally, a1 + a2 + a3 + · · · )
n=1

and is called the “sum” of the sequence a1 , a2 , a3 , . . .

Is this completely clear? Maybe not?


You might be tempted to think that the “convergence” of the series a1 + a2 +
a3 + · · · is about (a1 + a2 + a3 + · · · + an ) for n → ∞—but this is not the case! To
understand the real meaning of “ lim sn ”, one has to examine the definition of the
n→∞
second step above more deeply and this becomes complicated!
Instead of worrying about this any longer, let us now turn to Bolzano, because he
has defined exactly the same notion (“convergence” of a “series”, the “summable” of
Spivak), but in a linguistically much more ingenious way which can be understood
much more easily.

The Convergence of Series by Bolzano

An almost unnoticed mathematical treatise from Bolzano appeared in 1817 in


Bohemia (far from Paris), for which Bolzano is still praised today. This praise is
completely justified, as we shall soon see.
Bolzano did not give the concept any name (such as “convergence”). He directed
the reader’s attention to this property by calling it “particularly remarkable”:
Particularly remarkable among such series is the class of those series with the property
that the change in value (increase or decrease), however far the continuation of its terms
is taken, always remains smaller than a certain quantity, which itself can be taken as
small as we please, provided the series has been continued far enough beforehand.

A lengthy and long-winded sentence (emphases are not at all added). Unfortu-
nately, this is characteristic of Bolzano. Nevertheless, this is a fascinating, clearly
made definition! Let us analyze this sentence in more detail.
1. The crucial passage is “however far the continuation of its terms is taken”.
Bolzano writes it as a formula in the following way:

an + an+1 + an+2 + . . . + an+r . (¶)


106 9 Bolzano: The Republican Revolutionary of Analysis

Spivak notes for this (caution: the convergence of series is about the estimating
of the magnitude of the sums sn but only indirectly about the an !):

|sn+r − sn−1 | .

Although it says the same, this is much less understandable. Really! (Is this true?)
2. “However far the continuation is taken” means: r is allowed to get arbitrarily
large; there are allowed arbitrarily many, very many summands (but of course,
only finitely many).
3. By the “however far the continuation is taken” (that is the sum ¶), Bolzano
demands it must “always remain smaller than a certain quantity, which itself can
be taken as small as we please”. Today we formulate the latter as seen already in
Spivak: “for all ε > 0 we have | . . . | < ε”, and the three dots stand in for almost
the entire sum ¶.
Today the pompous formulation “for all . . . ” is called the “universal quantifier”.
Bolzano gives a completely different formulation, namely this: “. . . a certain
quantity [i.e. ε], which itself can be taken as small as we please”. This is
much smoother than the brutish “all”: “a certain . . . , which itself can be
taken . . . ”. (That quantities do not have negative values—Spivak’s “> 0”—was
in the mathematics of the time of Bolzano a matter of course and, consequently,
not written down.)
4. The final touch: the phrase “can be taken” implies a basic assumption, namely,
the series “has been continued far enough” beforehand. What does Bolzano
consider here?—The n: with his formulation “has been continued far enough” he
means: in order that the sum an + an+1 + an+2 + · · · + an+r remains always—for
each r—smaller than the chosen ε, the n has to be taken large enough.
This condition is also formulated with the help of a quantifier today, namely
with the other one, the “existential quantifier”: “there exists a number N such
that . . . ”. Again, Bolzano formulates this more simply: “the n must be chosen
great enough”. The content is the same: what Spivak gives in a normed, pompous
formulation, Bolzano puts in simpler prose, but not in colloquial language.
Unfortunately, such prosaic language is impossible in case of intricate notions
such as “convergence”. Mathematics does not come without effort!
5. Recapitulation: if we now reread Bolzano’s sentence about the “particularly
remarkable” class of series, we would be able to enjoy it, wouldn’t we? In
this single sentence, Bolzano embraces everything for which Spivak needs three
stages of a definition (the last one on p. 105). Additionally, Bolzano presents
the most general definition of “convergence”. He does this in a single sentence
of five (in the German original, seven and a half) lines. Of course, Bolzano’s
sentence is somewhat challenging, in the vocabulary as well as in the syntax.
Nobody understands it at once. One has to read it repeatedly. But it can be
understood without any previous knowledge (especially of logic). This sentence
Continuity with a New Meaning 107

is not technical at all and it lacks modern quantifiers. Therefore it is open to the
laymen and the laywomen. Let us note:
The modern concept “convergence”, in its most general version, was first formulated
by Bolzano in the year 1817. (See the previous box as well as the proof on pp. 104f.)

In science it is common to take the year of publication as the year the text was
created because in many cases (not in this one!) it is hard to detect in which year
the manuscript was really finished. The publishing date is usually undisputed.

The Remaining Deficiency

Although Bolzano argued very subtly and conceived his notions very carefully, he
capitulated at the same point as once Leibniz did (see p. 31): just like his renowned
predecessor, he also, one and a half centuries later, lacked a concept of number
which served his needs. As stated before, to calculate a number “as precisely as
one wishes” is not the same as to have a number, to present it! To carry out
its proofs, mathematics needs a proper concept of number (which itself can be
constructed in a proof); otherwise it cannot do so. In 1817, Bolzano had none. (Only
half a generation later, he worked on this project, but this remained hidden in his
manuscripts until the last quarter of the twentieth century.)

Continuity with a New Meaning

Convergence Works with Discrete Objects

The notion “convergence” is made for discrete objects.


In the nineteenth century, those discrete objects were the summands of the infinite
series. The series a1 + a2 + a3 + . . . consists of the discrete objects: a1 , a2 , a3 , . . .
Towards the end of the nineteenth century, these discrete objects as they stand
were taken as genuine mathematical entities in their own right and they got a name,
which was never used before: “sequence”. (This way a new mathematical object
was created.) The “sequence” is a1 , a2 , a3 , . . . , and a new notation was invented:
“(an )n ”, or sometimes in full length: “(an )n∈N ”, whereby “ N ” denotes the set of
natural numbers. Some authors use braces instead of round brackets and sometimes
the running index is omitted, and then the sequence is just “{ an }”, more on this in
Chap. 13.
108 9 Bolzano: The Republican Revolutionary of Analysis

Continuity Is Analogous to Convergence

Anybody who understands the concept of “convergence” will recognize at once the
meaning of “continuity”.
The transfer of the modern concept of “convergence” into the continuum creates today’s
concept of “continuity”.

The notion “continuity” is more easily visualized. Therefore, if you have


difficulties with the notion “convergence”, you should study the notion “continuity”
and return thereafter to “convergence”.
The German language has two expressions, whereas the English as well as the
French only has one: the German calls the new technical term “stetig” as distin-
guished from “kontinuierlich”, whereas English and French only have “continuous”
and “continue”, respectively.

Continuity of Functions in Bolzano

Bolzano defines continuity to be the following property: a function is called


“continuous” at the value x0 , if
the difference f (x0 + ω) − f (x0 ) can be made smaller than any given quantity, provided ω
can be taken as small as we please.

Bolzano takes ω, the last letter of the Greek alphabet, as a variable which can be
“taken as small as we please”. Without stating it explicitly, Bolzano always assumes
a positive ω: ω > 0.
Let us construe the function f (x) not only as a calculatory expression but also
as a drawn curve in the coordinate system. If so, Bolzano’s definition says: at the
value x0 , the curve does not have a sudden change. The more the x-values approach
the value x0 (i.e. the smaller ω gets), the smaller is the difference of the respective
values of the function f (x + ω) from the value f (x0 ); expressed formally,

if |ω| < δ we have |f (x0 + ω) − f (x0 )| < ε .

Bolzano says: “The difference [namely: |f (x0 + ω) − f (x0 )|] can be made smaller
than any given quantity, provided ω can be taken as small as we please.”. It is a
logical challenge to grasp this. It says that the δ is to be defined after ε is given, at
first the ε and then the δ. Can we accept this?
As it turns out, at this point Bolzano is a little bit sloppy! He denotes the “value”
of the variable x in the same way as the “variable” itself, plainly by “x”. The notation
“x0 ” for a particular “value” of the variable x is my doing, it is not in Bolzano. (I
adopted it from Cauchy, as will be seen in the following chapter, p. 123.) As far as I
know, nobody was ever disturbed by Bolzano’s carelessness; everybody understood
Continuity with a New Meaning 109

him correctly. But I want to proceed here as pedantically as it is (meaningfully)


possible.

The Little Difference Between Then and Now

Bolzano’s definition of “continuity” is neat, clear, and still valid today. Nevertheless,
for his contemporaries it meant something other than for us today, because they
unspokenly assumed:
A quantity does not have negative values.
At the beginning of the nineteenth century, negative numbers were still not
accepted as true numbers in mathematics. (See also p. 27.)
In his definition, Bolzano unequivocally writes “x0 + ω” instead of “x0 ± ω”.
(And of course, with “+”Bolzano also means +, not “+ or −”.) The implication is
as follows: Bolzano studies only the right side of x0 , that is to say only, those values
that are greater than x0 . The smaller ones (i.e. x0 − ω) he ignores. Consequently,
Bolzano only defines one-sided continuity.
The scrupulous thinker Bolzano was clearly aware of this. In his later work
Theory of Functions, he distinguishes between “positive increments”’ and “negative
increments”, and similarly he differentiates between continuity “in the positive
direction” and “in the negative direction”. If both apply, he uses the words “plainly
continuous”. (However, in his formulae, he only uses the plus sign.)
Such a strange thing as alternating signs within the terms of the sequence, as for
instance in the sequence 1, − 12 , 13 , − 14 , 15 , − 16 , . . . , could not be imagined even
by somebody as revolutionary as Bolzano at that time. This was not due to a lack
of radicalism in Bolzano’s thought, not at all! The explanation is this: according
to Bolzano, the continuous variable is, of course, only capable of continuous
changes! Naturally, Bolzano thinks of the quantity ω, denoting an “increment”
of the continuous variable x, as being continuous, too. In all probability, his
contemporaries did the same. (The mathematical object “sequence” did not exist
at the beginning of the nineteenth century, and it was not invented yet.)

The Differences from Euler’s Continuity

Euler called a “curved line” “continuous” if it could be described by a calculatory


expression (see p. 81). Bolzano calls a “function” “continuous” if its geometric
drawing shows no sudden change.
These two concepts of continuity differ in two respects:
1. They relate to different objects: Euler deals with “curves” and Bolzano with
“functions”.
110 9 Bolzano: The Republican Revolutionary of Analysis

2. They denote completely different properties: for Euler it is the property that a
curve can be described as a whole by a single calculatory expression, and for
Bolzano it is the property of a function to change its values near to another
value only very gradually.
That is why Euler’s concept of continuity is sometimes called “global” (it relates
to the whole curve), while Bolzano’s concept is called “local” (because it only
operates in the neighbourhood of a single value x0 ). However, a helpful comparison
of both concepts is difficult, as they deal with totally different objects (curve resp.
function). It gets even worse if one tries to include d’Alembert’s idea about “not
continuous”: p. 93, item a.
Sometimes one is a bit less rigorous and identifies a “function” with its drawing
in the coordinate system; this lessens the differences of the two notions of continuity.
But mathematics should have no place for inaccuracies!—however, see p. 125. . . !
Euler’s concept of continuity, together with his “Algebraic Analysis”, perished a
long time ago. The concept of continuity has survived, as it was created for the first
time by Bolzano. This is not the merit of Bolzano. On the contrary, this shows: in a
certain sense our analysis today is the same as Bolzano’s—but completely different
from Euler’s.

Continuity and the Continuous

The new concept “continuity” is conceived for dealing with the continuous. It
transfers the notion “convergence” from the discrete to the continuous. This we
shall point out shortly.
“Continuity” demands for the function f (x):

for |ω| < δ we have |f (x0 + ω) − f (x0 )| < ε .

Now we concentrate on two analogies:


1. Instead of the continuous object “variable”, i.e. x, we take the discrete object
“natural number”; the individual may be named n.
2. Instead of the continuous object “function”, i.e. instead of f (x), we take the
discrete object “sequence”, i.e. (an )n .
Consequently, the fixed value of the function f (x0 ) corresponds to the limit l,
and the continuously varying values of the function f (x0 + ω) correspond to the
discrete terms an of the sequence.
Then |f (x0 + ω) − f (x0 )| < ε
changes to |an − l| < ε.
Bolzano’s Revolutionary Concept of Function 111

Finally, the condition to determine a sufficiently small δ to a given ε such that


|ω| < δ is guaranteed for the first inequality translates to the condition to determine
a number N great enough such that for n > N the second inequality is valid.
—Done!

Bolzano’s Revolutionary Concept of Function

Even after his relegation from university, Bolzano had the possibility of continuing
his studies. (Only a few succeed in this.) Some 17 years after the publication
of the treatise for which he is still praised today, his mathematical manuscripts
were completed sufficiently to provide a foundation for his work on a textbook of
analysis. The title of this manuscript reads simply: Theory of Functions.
Bolzano had already laid those foundations in earlier papers. Here the mature
thinker Bolzano reached a notion of function, which is quite removed from
everything that his contemporaries were able to think about.
Because this innovative work stayed unknown until the twentieth century, it could
not be of any influence within the world of mathematics. However, it is spectacular
to such a degree that we will give a brief comment on it below.

Bolzano’s Definition of the Concept of Function

As the reader might already expect, Bolzano expresses this definition also by a
single (somewhat lengthy) sentence:
The variable quantity W is a function of one or several variable quantities X, Y, Z, if there
exist certain propositions of the form: the quantity W has the properties w, w1 , w2 , which
can be derived from certain propositions of the form: the quantity X has the properties ξ ,
ξ  , ξ  —, the quantity Y has the properties η, η , η , etc.

If we translate this long-winded language of Bolzano’s Biedermeier period into


today’s words, it might read as follows:
A variable quantity w is called a “function” of the variable quantities x, y, z, if the properties
of w are determined by the properties of x, y, z via a law which depends on certain values.

Or, more succinctly:


A quantity w is called a “function” if its values are determined by a law which depends on
the values of other variables.

This is the most general notion of function (in analysis) possible and this even
by today’s conceptions and standards. It is crucial that the “calculatory expression”
(as in Euler) is no longer mentioned! Only a “law” is required, a “dependency from
certain values” (Bolzano uses the word “derivable”).
112 9 Bolzano: The Republican Revolutionary of Analysis

Bolzano’s Examples of Functions

Bolzano not only jots down this definition, he actually means it in this generality.
One can see that already in his Theory of Functions by studying his very first
example of a function:
If W denotes the prize that shooters are rewarded with for skill at target practice if we decide
that a shot in the bullseye should get 100 Reichsthaler, a shot which is a distance = x inches
from the bullseye, where x is not more than 2, gets 100 − 25x Reichsthaler, and a shot with
distance from the bullseye > 2 and < 5 gets 58 − 2x 2 Reichsthaler, etc.

Judgement

It is not the (very simple) calculatory expressions used by Bolzano which are
important, but the arbitrary piecewise manner of the definition of the winnings.
For Bolzano’s contemporaries (like Cauchy—we shall meet him in the next
chapter), such a way of specifying the mathematical concept of “function” was
completely unthinkable and downright out of place. Analysis should deal with such
objects? “Anything but that!” was the opinion of Bolzano’s contemporaries.
Bolzano was decades ahead of his time with this idea. Riemann’s definition from
the year 1854 clarified one aspect of Bolzano’s definition but actually did not add
anything new to it. However, Riemann’s formulation was printed and became well-
known and marks the state of affairs to this day. (We shall come to that on p. 155.)

Mathematical Consequences of Bolzano’s Notion of Function

Another demonstration that Bolzano really meant his notion of function to be so


radical can be found in a casual half-sentence in his Theory of Functions (there
Bolzano uses even the formulation “dependency”):
As it is permitted to think the law of dependency of a number from another as we wish . . .

Then Bolzano gets creative. At first he constructs a (simple) function, which is


not “continuous” for any chosen interval of x, be it as small as one wishes.
In contrast to this, his contemporaries were convinced that a function is usually
nearly everywhere “continuous”—with the exception of some particular values.
Bolzano constructed a radical counter-example to this way of thinking.
But this is still not enough! Bolzano constructs other functions with completely
unexpected properties—e.g. one which “in between the limits of x = 0 and x = 1
infinitely often increases or decreases”.
Finally, Bolzano constructs (by slightly overstretching his notion of function) a
function, which for any interval of x, however small, is neither only increasing nor
Literature 113

only decreasing during this interval. (Today we say: which is not “monotonous” in
any interval, however small it is taken.)
This function invented by Bolzano is a true mathematical monster. It is continu-
ous, but it is in no interval monotonous. Even worse, it cannot be differentiated at
any value.—Contemporaries and successors of Bolzano were thoroughly convinced
that a function could be differentiated (nearly) everywhere, some of them even as
late as in 1874: see p. 171. Bolzano’s thinking was far more radical—more like ours
today.

Literature

Bolzano, B. (1816). Der binomische Lehrsatz, und als Folgerung aus ihm der polynomische, und
die Reihen, die zur Berechnung der Logarithmen und der Exponentialgrößen dienen, genauer
als bisher erwiesen. Prag: C. W. Enderssche Buchhandlung. http://dml.cz/handle/10338.dmlcz/
400346. English in: Russ 2004, pp. 154–248.
Bolzano, B. (1817). Rein analytischer Beweis des Lehrsatzes, daß zwischen je zwey Werthen, die
ein entgegengesetztes Resultat gewähren, wenigstens eine reelle Wurzel der Gleichung liege.
In P. E. B. Jourdain (Ed.), Leipzig, 1905. http://dml.cz/manakin/handle/10338.dmlcz/400352?
show=full. English in: Russ 2004, pp. 249–277.
Bolzano, B. (1830–1834). Reine Zahlenlehre. In J. Berg (Ed.), Bernard-Bolzano-Gesamtausgabe,
vol. II A.8. Stuttgart–Bad Cannstatt: Friedrich Frommann Verlag (Günther Holzboog), 1976.
Bolzano, B. (1830–1835). Erste Begriffe der allgemeinen Größenlehre. In J. Berg (Ed.), Ein-
leitung zur Größenlehre und Erste Begriffe der allgemeinen Größenlehre. Bernard-Bolzano-
Gesamtausgabe, vol. II A.7, pp. 217–285. Stuttgart-Bad Cannstatt: Friedrich Frommann
(Günther Holzboog), 1975.
Bolzano, B. (1834–1842a). Einführung in die Funktionenlehre. In K. Rychlík (Ed.), Bernard
Bolzano’s Schriften, vol. 1. Prag: Königl. Böhmische Gesellschaft der Wissenschaften, 1930.
http://dml.cz/handle/10338.dmlcz/400333.
Bolzano, B. (1834–1842b). Functionenlehre. In: B. van Rootselaar (Ed.), Bernard-Bolzano-
Gesamtausgabe, vol. II A.10/1. Stuttgart-Bad Cannstatt: Friedrich Frommann (Günther
Holzboog), 2000. English: Russ 2004, pp. 429–589.
König, G. (Ed.) (1990). Konzepte des mathematisch Unendlichen im 19. Jahrhundert. Göttingen:
Vandenhoeck & Ruprecht.
Russ, S. (2004). The mathematical works of Bernard Bolzano. Oxford: Oxford University Press.
Spalt, D. D. (1990). Die Unendlichkeiten bei Bernard Bolzano. König 1990, pp. 189–218.
Spivak, M. (1967). Calculus. Amsterdam: W. A. Benjamin, Inc.
Chapter 10
Cauchy: The Bourgeois Revolutionary as
Activist of the Restoration

Cauchy: The Atipode to Bolzano

Augustin-Louis Cauchy (1789–1857) was taught mathematics between the ages of


15 and 18 at the École Polytechnique in Paris, by first-rate teachers of his time. At
the age of 26, he himself started teaching at this university. In the meantime, he
worked as an engineer, for instance, in Paris on the construction of canals and an
aqueduct and in Cherbourg on the construction of the new harbour.
This is an extraordinary biography: a singularly gifted mathematician works
in his earliest years for several years in the material construction of the world
through mathematics until he starts his creative mathematical career. This had
unique consequences on the mathematical disposition of the man Cauchy as well
as on the transformation of mathematics as a science through Cauchy.
For the sake of completeness, we may add: at the age of 26, after the downfall of
Napoleon and the beginning of the Restoration, Cauchy (together with a physicist)
was appointed by King Louis XVIII to the Académie des Sciences—this after
the exclusion of 62-year-old Lazare Carnot and 69-year-old Gaspard Monge for
political reasons.

The Heart of Cauchy’s Revolution of Analysis

As a mathematician, Cauchy was influential not so much as a member of the


Academy but as a university teacher. Accompanying his lectures in 1821, he pub-
lished a textbook on Analysis (Cours d’Analyse de l’École Royale Polytechnique),
in 1823 a textbook on differential calculus, and in 1829 one on integral calculus.
(Cauchy counts as one of the most productive scientists. In addition, he wrote plenty
of treatises; today’s edition of his Collected Works comprises 27 large sized, thick
volumes.)

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 115
D. D. Spalt, A Brief History of Analysis,
https://doi.org/10.1007/978-3-031-00650-0_10
116 10 Cauchy: The Bourgeois Revolutionary as Activist of the Restoration

In the introduction of his textbook on Analysis, he drew conclusions from his


experience as an engineer who did practical work by writing the following two
sentences (the emphases are added):
We must also observe that [the arguments drawn from the generality of algebra] tend to
grant a limitless scope to algebraic formulas, whereas, in reality, most of these formulas
are valid only under certain conditions or for certain values of the quantities involved. In
determining these conditions and these values and in establishing precisely the meaning of
the notation that I will be using, I will make all uncertainty disappear, so that the different
formulas present nothing but relations among real quantities, relations which will always be
easy to verify by substituting numbers for the quantities themselves.

There had never been a more radical declaration of war on traditional mathematics in
modern times.

The radicalism of this challenge was so comprehensive that nobody was aware
of it or even took it seriously, neither Cauchy’s contemporaries nor later historians
of mathematics.
This is true even though Cauchy said everything in these two sentences. For what
do they say?
The first sentence says that the formulae of the Algebraic Analysis of Lagrange
and Euler are completely general (they “tend to grant a limitless scope”)—while
practical mathematics (the “reality”) shows that these formulae are valid only “for
certain values”. In short, the mathematics from the Academies is vague, aloof, and
removed from reality—its application to practical problems requires the precise
determination of their validity.
The second sentence says two things: (i) the meanings of the mathematical terms
must be clear and unequivocal and (ii) through the assignment of the VALUES in the
expressions, we obtain crystal-clear numbers which are suitable in practice. And the
numbers are such “values” by all means.
In the years he was working with others as an engineer, Cauchy experienced at
first hand:
The analysis only becomes useful and effective when its general formulae are adapted in
such a way that they result in numerical values.

What was Cauchy to do when he was appointed to teach Analysis at the École
Polytechnique to future engineers? He taught them to adjust the general formulae
by substituting numbers in order to make them useful for the practice.

Mathematical View of Cauchy’s Revolution of Analysis

Euler and Lagrange founded Algebraic Analysis on a most general concept of


“variable”. By wanting to go beyond this, Cauchy, as a matter of principle,
declared that he wished to materialize the “variables” by assigning “values” to them.
However, this amounts to nothing else than that:
Cauchy’s Concept of Variable Is Determined by “values” 117

Cauchy wished to add to the original foundational concept of Algebraic Analysis


“variable” a second and new one: the foundational concept “value”.

Cauchy’s motive for this innovation was a non-theoretical one: his practical
experience in applying the formulae of Algebraic Analysis.
However, the consequences of Cauchy’s move are deeply theoretical. Even the
layman will understand that:
A change in the foundational concepts of a theory necessarily changes the theory itself.

The reason is that “theory” in mathematics involves proofs. Yet the foundational
concepts are the starting points of those proofs. Subsequently, it follows that new
foundational concepts must change the proofs!
In other words, by announcing, as a matter of principle, the making of general
formulae of the “Algebraic Analysis” fit for praxis by introducing values into the
expressions, Cauchy complements the foundation of the theory (the “variable”) by
concretizing it to take specific “values”.
Such a foundational reconstruction of a theory is nothing other than a conceptual
revolution. The use of the term “revolution” is not just empty talk or a way of
attracting attention; it shows the significance of Cauchy’s innovative idea.
Science needs clear concepts. A clearly determined fact needs a special, unique
name.
This also holds for the science history of mathematics. The analysis of Euler and
Lagrange has been called “Algebraic Analysis”—we propose the name “Calculus
of Expressions” (see p. 94).
Cauchy abolishes this kind of analysis and establishes a new theory instead. He
states this explicitly in the Introduction of his first textbook of analysis. Therefore,
Cauchy’s analysis clearly needs a new name of its own!
Fascinatingly, history of mathematics has not come up with such a name until
today: the analysis as introduced by Cauchy has not been christened yet. Therefore,
we are free to do so. My proposal is as follows: let us call it “Calculus of Values”
or “Analysis of Values”, a name that draws attention to the structural change in
analysis. Our result is:
Cauchy transformed the “Calculus of Expressions” (or “Algebraic Analysis”) into the
“Calculus of Values” (or “Analysis of Values”).

Cauchy’s Concept of Variable Is Determined by “values”

Cauchy defines “variable” this way:


We call a quantity variable if it is assumed to take on many different values.

This is not alarming. But it clearly differs from Euler’s declaration (p. 71). Euler
demanded that “all definite values whatsoever” should be taken by the variable.
Thus, since Cauchy a “function” may be defined only within a finite interval. We
118 10 Cauchy: The Bourgeois Revolutionary as Activist of the Restoration

now operate within a “Calculus of Values” and no longer within a “Calculus of


Expressions”.
The graduate of mathematics realizes that Fourier Analysis is only possible since
Cauchy—i.e. the method of finding a suitable calculatory expression in the form of
an infinite trigonometrical series for an arbitrarily given function. The reason is that
the construction of this series requires the calculation of definite integrals and this
is only possible if the domains of the respective functions are finite intervals.
Euler, the indefatigable calculating mind nevertheless managed to construct the
aforementioned “Fourier Integrals”. However, as Euler was unable to conceive of
a “function” with a finite interval as its domain, he was unable to recognize the
wider scope of his calculation. This interpretation goes back to Jean Baptiste Joseph
de Fourier (1768–1830) who wrote it down in the first quarter of the nineteenth
century and who subsequently lent his name to this technique.
What cannot be seen in this definition of “variable” is its actual usage by Cauchy.
At first sight, Cauchy’s designation of the “variable” seems to be hardly different
from that of Euler—apart from the already noted (and very important!) aspect that
Cauchy no longer demands the variable to take “all definite values whatsoever”. But
beyond that?
The difference cannot easily be seen from Cauchy’s definition of the notion of
variable. The meaning of “value” introduced by Cauchy has further importance
and lends a greater degree of precision than it had for Euler or Lagrange and their
“Calculus of Expressions”.

Cauchy Derives “number” from “quantity”

“Quantity”

Modern analysis is founded on the concept of “number”. Yet, Cauchy could not
build on this concept because he did not have a suitable notion of “number”. He
knew that the decimals are suitable and useful for practical calculations. Through
the lack of an alternative idea, Cauchy stuck to tradition and relied, as did his
predecessors, on the notion of “quantity”. (For the proofs, mathematicians need
concepts.)
Cauchy starts, in a similar way to his teacher Sylvestre François Lacroix (1765–
1843), with “magnitudes” and focuses on their changes, more precisely on their
increases or decreases. These changes of the “magnitudes” he calls by the traditional
name “quantity”. (The word “Zahlgröße” was coined in the German language for
this type of quantity at that time.) We summarize:
The “quantity” is the increase or the decrease of a “magnitude”. The “opposite” quantity is
the decrease or the increase which the second quantity undergoes to reach the first.

That means that “quantity” is a change, which is a notion that comprises motion.
This is clever because the concept itself is fixed, motionless. It only comprises
Cauchy Derives “number” from “quantity” 119

motion (and we already know from p. 27 that according to the standards of classical
mathematics motion is not accepted in mathematics).
Let us look at the philosophical aspect: Cauchy’s definition, just like Euler’s (see
p. 70), does not say what “magnitude” essentially ought to be.

“Number”

Cauchy bases the concept “quantity” on the notion “magnitude”. In the same way,
he deals with the concept “number”. He writes:
The measure of the second magnitude, if compared to the first, is a number which is
represented by the geometrical ratio of one to the other.

If, e.g., the two measures are “6 hours” and “2 hours”, their geometrical ratio is
the “number” 3:
6 hours 6
= = 3.
2 hours 2
This does not seem convincing: doesn’t the measure “6 hours” already contain
the number 6?—Strictly speaking, this is not the case! The ancient inhabitants
of Mesopotamia arranged their civic and economic activities as well as their
fortifications, canal planning, and accounting (the documents in clay have survived
until today) without knowledge of the abstract concept of “number”. For anything
and everything, they had their own system of measure; these systems were totally
different from each other. The signs for “6 days” and for “6 goats” were, for instance,
completely different.
Here is not the place to go into further detail, but those thousand years of
early civilization show clearly: the operation with measures does not necessarily
require the knowledge of the abstract object of “number”. Measuring and (abstract)
counting are completely different actions. Accounting does not need numbers! His-
torically, measuring preceded counting by specialists (with numbers) by thousands
of years.
Cauchy dealt with the concept of number only to a certain degree. (He asked, for
example, what about 0? Following the existing concept, zero is no number!) We are
content with this answer: this way, Cauchy was at least able to give a clear concept
of fractions (or rational numbers).
However, for doing analysis, Cauchy needs more, at least the “irrational” num-
bers. How did Cauchy think about them?—For this, we need some preliminaries.
The central conceptual tool for the construction of analysis introduced by Cauchy is the
concept of “limit”.

The mathematical abbreviation is “lim”, standing for the Latin “limes” and the
French “limes” and “limite”.
120 10 Cauchy: The Bourgeois Revolutionary as Activist of the Restoration

The Basic Definition of “limit”

The object is not new. We have already found it in Leibniz, with regard to his
convergence criterion (as we call it today, p. 24) and his calculation of areas with a
curved boundary (p. 31), then again in great detail in Bolzano (pp. 104 and 105).
These authors did not create a name for their new object. (Others did, e.g.
d’Alembert.) Cauchy too, now gives it a name:
When the values successively attributed to the same variable approach a fixed value
indefinitely, in such a manner as to end up differing from it by as little as we will wish,
this final value is called the limit of all the others. It is written

lim x = X .

Undoubtedly, the concept “limit” is difficult. It is as difficult as the concepts


“convergence” and “continuity”! Cauchy does not worry about these difficulties.
(Pedagogical abilities do not always go hand in hand with subject knowledge.) His
contemporary and textbook author Johann Tobias Mayer (1752–1830) tackles this
problem in his book Complete Teaching of Higher Analysis, published in 1818, in
quite another way. Mayer does not use the concept “limit” at all (which does not
prevent him from writing at some point about a “ratio of limits”), but he treats the
subject on no less than 20 printed pages!
This topic occupies Mayer’s mind to such a degree that he even introduces a
special sign for it. Interestingly, it is not the operator “lim” but the binary relation
“≡”:
If one were to introduce a special sign in analysis to indicate an infinite approximation of
some quantity to another, e.g. the sign ≡, nobody would take offence that, if the equation
read
1 1 1 1
T = + 2 + 3 + x
log x x x 2

where the quantity x increases without end, i.e. becomes infinite, we would get

1
T ≡ .
log x

As we remember, Johann Bernoulli did not come up with the idea of a second
sign for another type of equality: p. 61. Times have changed. (As a mathematician,
Mayer cannot hold a candle to Johann Bernoulli!)
Cauchy needs less than four lines in his preface for the same fact. This is followed
by another four to five lines with two examples: irrational numbers as limits of
fractions and the circle as the limit of inscribed polygons. His students will have
found this hard to swallow, but this is quite another matter.
Cauchy Derives “number” from “quantity” 121

The Unspoken Luxury Version of the Concept of Limit

Cauchy’s first example for his concept of limit, which was just mentioned above,
is that of the irrational numbers as limits of fractions. However, there is a snag in
this idea! Strictly speaking, Cauchy is not allowed to introduce an irrational number
as the limit of an approaching sequence of fractions: it does not meet his notion of
limit at all!
Just picture the problem: we do know what fractions are and we want to know
what irrational numbers are. For us, at this stage, the irrational numbers do not yet
exist at all.—Let us reread Cauchy’s text: “When the values successively attributed
to the same variable approach a fixed value indefinitely . . . ” It clearly mentions
a “fixed value” which is to be “approached” by a sequence of values. But if the
irrational number does not yet exist (quite to the contrary, it is yet to be constructed),
then Cauchy’s concept of limit is of no help—for there does not yet exist this “fixed
value” which is to be approached by this sequence of fractions. This “fixed value”
has, first of all, to be created by this construction.
However, Cauchy does not seem to be aware of this problem. Directly afterwards,
just after having given his definition, he puts forward the irrational numbers as
examples of “limits”.
And this is no slip, for later on when he defines the operations for the calculation
of the numbers in detail, he proceeds in the same way. Cauchy takes yet again the
irrational number B just to be B = lim b where the (variable) quantity b is only
allowed to have fractional values.
Today we have a problem with this strategy. We strongly discriminate between
whether the “limit” spoken of already exists and whether it does not yet exist and is
only to be constructed or defined with the help of this sequence. For us, these are two
different notions of “limit”. In the first case already existing facts are described: a
given number is called a “limit”, whereas in the other case a new mathematical
entity is created through being a “limit”. This second case could be named the
“luxury version” of the notion of limit for it offers more than the other one, it
produces a new mathematical entity (e.g. an irrational number).
In the true sense of the luxury version, an “irrational number” is nothing else than just
this sequence of fractions, for which it is the “limit” (which is approaching it as nearly
as we wish).

They exist only in this way. For example, π is nothing else than the sequence:

3; 3.1; 3.14; 3.141; 3.1415; ...

Cauchy’s contemporary Johann Traugott Müller was well aware of this (see
pp. 200f).
122 10 Cauchy: The Bourgeois Revolutionary as Activist of the Restoration

What Is the Difference?

Cauchy writes as if there were no difference between the two (for us today) different
versions. Was he conscious of this cheating?
I do not know. I did not find any formulation by him showing any awareness of
this problem.
Let us remember Bolzano’s view of this problem (see p. 104). According to
him, a number is already well-defined if it could be calculated “as precisely as one
wishes”. As his contemporary, Cauchy could have thought in just the same way.
If this were the case, then for Cauchy the difference of these two versions of the
notion of the limit was not discernible: it is the difference between those two “steps”
which we made in presenting the modern notion of convergence (following Michael
Spivak: p. 103)! We can state:
Neither the brilliant thinker Bolzano nor the outstanding mathematician Cauchy in the first
third of the nineteenth century made this distinction. However, Johann Traugott Müller did
in 1838.

Consequently, this difference did not exist in their analysis. (But it exists today in
our analysis.)
Some may judge this to be wrong. They might say that this difference between
these two versions of the notion of limit (as well as the notion of convergence) is
eternal.
But the proponents of this view are obliged to spell out what is meant by the
“existence of a notion”—when even the most reknown mathematicians did not
realize it. The hidden implication of this view is that mathematical notions (and
truths) exist independently of their discovery and even if they are not used at all in
what is well-known of mathematics.

“Function” and “value of a function” in Cauchy

Let us return to the foundational concepts of analysis, to one of the central notions.
What is Cauchy’s understanding of “function”?

The Concept of Function in Cauchy

Cauchy expresses very clearly and at length what he understands by a “function”,


namely a “changing” quantity:
When variable quantities are so related among themselves that, the value of one of them
being given, we are able to deduce the values of all the others, we usually consider these
various quantities expressed by means of one among them, which then takes the name of
the independent variable, and the other quantities, expressed by means of the independent
variable, are what we call functions of this variable.
“Function” and “value of a function” in Cauchy 123

That is to say, given the four quantities x, x 3 , 5ex+2 , and sin x, the last three are
called “functions” of the first, which is the “independent variable” x.
Cauchy is very scrupulous and defines the same again—in case there are more
than one “independent variable” (as in 5x 2 + 2y 3 ). We can pass over this here.

The New in Cauchy’s Concept of Function and a New Style of


Notation

His detailed talk of “values” is easily discernible: the “independent variable” takes
on “values” and from those the “values” of the variable quantity are “to be deduced”.
The whole is called a “function”.
Not one mathematician before defined “function” in this way. This is definitely
new: such detailed talk of the “values” of the variables. (We remember Bolzano’s
definition, which was given half a generation later: p. 111.)
Besides, a new entity deserves a new notation! Since Leibniz (and in the tradition
of an idea from Descartes), the “variables” in mathematics are usually written as
small and italicized letters, mostly “x”, “y”, “z”. Now Cauchy introduces new and
important entities. Consequently, it is utterly worthwhile that he introduces a new
notation for these entities (“values”). (He clearly asks for this in his Introduction!)
Thus Cauchy consistently denotes the “values” of a “variable” in one of the
following two ways: either he adds a lower index (e.g. “x0” and“x1” denote “values”
of the “variable” x) or he uses the same letter written in capitals and upright: “X”,
too, denotes a “value” of x for Cauchy. To sum up,
In each case, Cauchy demands for an “independent” variable the precise specification which
“values” it is allowed to take.

In general, Cauchy writes this: “Suppose x between x0 and X.” This means that
the “variable” x takes all the “values” between the “values” x0 and X. Today we
write this as “x0  x  X” or in the language of sets “x ∈ [x0 , X]”.
We will still have to think about the formulation “we are able to deduce” in his
definition of function, but we postpone this to the section after the next. Let’s ask
first: is there anything else new with Cauchy’s concept of function?

Cauchy’s Concept of Function Is as Conservative as Possible for


a Revolutionary

Apart from introducing the “values” of the variables (and consequently insisting
on having a restricted domain for the “independent” variable), Cauchy’s concept of
function does not differ from that of Euler!
This we can guess from a comparison of the wordings of both definitions.
124 10 Cauchy: The Bourgeois Revolutionary as Activist of the Restoration

This can be seen in Cauchy’s analysis: all his functions are given as formulae—
just as with Euler!
And it can be understood by comparing Cauchy’s definition with that given by
Bolzano (some ten years later, p. 111). Bolzano operates as generally as possible—
Cauchy stays as closely as possible to Euler’s analysis.
In fact, Cauchy confines himself to transforming the “Calculus of Expressions”
(“Algebraic Analysis”) of his forerunners as closely as possible into his “Calculus
of Values” (or “Analysis of Values”). He did not try to go beyond the intellectual
horizon marked by the “Calculus of Expressions”.
Considering the undisputed revolutionary nature of Cauchy’s departure from
the “Calculus of Expressions”, one may judge this adherence to tradition to be
appropriate or maybe even a clever diplomatic move. However, compared with
Bolzano’s way of thinking about analysis and especially about the concept of a
function, Cauchy’s thought is ultra-conservative—just as his personal convictions
and his ideology.

Cauchy’s Concept of the Value of a Function

Let us now return to the formulation in regard to the concept of a function which we
have mentioned earlier: the “we are able to deduce”. What could be meant by: we
“deduce” the value of a function?
The problem is clear-cut. A “function” according to Cauchy (as well as to Euler)
is e.g. x1 . Which values are there to be deduced? Clearly, all those which result
from the instruction “Divide 1 by . . . ”. This is simple—only x = 0 is excluded, for
division by 0 is not possible. So what?
In this case, Cauchy is forced to define what is to be done—what the “value” of
the function in such a case of exclusion ought to be.
And Cauchy does, of course! He is completely explicit and declares:
If a particular case arises in which the given function cannot immediately give the value
of the function under consideration, we seek the limit or limits towards which this
function converges as the variables approach indefinitely the particular values assigned
to them. If there exist one or more limits of this kind, they are regarded as the values of
the function under the given hypotheses; however many they may be. We call singular
values of the proposed function those values determined as we have just described.

Cauchy’s Concept “value of a function”: A First Example

We consider the function x1 . If x = 0, we do not immediately obtain a value.


Therefore we will look for all the “limits” which we can deduce from lim x = 0 :
lim x1 = ?
x→0
“Function” and “value of a function” in Cauchy 125

This is not difficult: the smaller the value of x becomes, the larger the value of
1
x will be. And the sign of the value of x1 is the same as the sign of x; both signs
are possible. Cauchy writes an arbitrary large number (in other words, an infinite
number) in the same way as Euler: “ ∞”.
Besides the numbers, ∞ is a “value”, too.

The result is lim 1


= ±∞. —Alright?
x→0 x

A Surprise: Cauchy’s “limit” Is Not Unambiguous!

This example (it is given by Cauchy) shows that Cauchy’s “limit” is not in the least
uniquely defined! It is possible to get different “limits” by different selections of
values of the independent variable(s). In other words,
Cauchy’s concept of limit generalizes the concept of convergence.

Only in the case when the “limit” is uniquely defined, do the notions of limit and
convergence coincide (see p. 110). And because “convergence” for the discrete is
analogous to “continuity” for the continuum, we have:
The unique “limit” is a new formulation of “continuity”.

The notion of “limit” highlights another aspect of “convergence”. (It denotes the
aim rather than the way to that aim.)
I have to confess: there has been a little bit of cheating in my argument! It
presupposed that ∞ was a “limit”. This is commonplace, also in textbooks of
analysis, for it is very convenient.
However, the assumption is false! The “value” ∞ has different properties from
the “numbers”! At no time does a finite “value”, an ordinary number, approach
the “value” ∞ as close as one wishes—which is indeed the very definition of the
concept “limit”! Quite to the contrary, the distance of some number a from ∞ is
always infinitely large: ∞ − a = ∞.
Nevertheless, it is common practice to take lim x1 = ∞, for some positive x, to
x→0
be really true. In other words, when one considers the strict meaning of the concept
of “limit”, this equation is not really true. Instead, it is a—very convenient—
agreement. Let us also stick to it.
126 10 Cauchy: The Bourgeois Revolutionary as Activist of the Restoration

A Second Example Relevant to Cauchy’s Concept “value of a


function”

For a better understanding of Cauchy’s concept “value of a function”, let us consider


the sine function as a second example. We may visualize it by the following picture,
a “curve” in the coordinate system:

The sine function


The “value of the function” sin x takes on every number between −1 and 1 for all
the “values” of x taken from any interval of length 2π . The “value of the function”
Y = sin X can be determined by a computer program (today’s style) or looked up
in a table (old-fashioned)—or computed with the help of an infinite series (Leibniz-
style).
According to our agreement, ∞ is a “value” too. Therefore, we may ask now:
what is the “value of the function” sin x for the “value” x = ∞?
What is sin ∞?
The answer is not too hard, or is it? It is each value from −1 to +1.
A short proof: let Y be any value in the interval [−1, 1].

The value of the sine function for X = ∞


We are sure that there exists a value X between − 12 π and + 12 π with sin X = Y.
(For this is a property of the sine function.) If Y = 0, we also have X = 0,
for we have sin 0 = 0. If Y = +1, we have X = 12 π , for we have sin 12 π = +1,
etc. And we also have sin(X + 2π ) = Y, for 2π is the period of the sine function.
Consequently, we have sin(X + 4π ) = Y, too, and so forth.

sin(X + 2k · π ) = sin X = Y , for all natural numbers k.


Some Very Surprising Consequences from Cauchy’s Concept of “value of a. . . 127

Now we let k become infinitely large: k → ∞, and consequently 2k → ∞ as well


as 2k · π → ∞. This amounts to
 
sin lim (X + 2k · π ) = sin(X + ∞) = sin ∞ = Y .
k→∞

However, Y was any arbitrary “value” between −1 and +1. Hence, sin ∞ may
be any of these values from −1 to +1 too. If we write all values from −1 to +1 as
the interval “[−1, +1]”, we get our result as a simple formula:

sin ∞ = [−1, +1]

And this was to be proved! (Cauchy invents another symbol for this, see p. 147, but
it means the same.)

Some Very Surprising Consequences from Cauchy’s Concept


of “value of a function”

The Methodical Significance of Cauchy’s Definition of This


Concept

A crystal-clear definition of the concept “value of a function” is an absolute must


for Cauchy. After all, he did announce: “in determining these conditions and these
values, I will make all uncertainty disappear”. Since Euler, “function” has been the
central concept of analysis, the most important entity whatsoever. Therefore, Cauchy
has to declare unmistakably which “values” the “function” ought to take without ifs
and buts.
Cauchy has met this requirement entirely. We have already examined his
definition of the concept “value of a function” (pp. 124f).

The Historical Significance of Cauchy’s Definition of This


Concept

We emphasize:
Cauchy is the first mathematician to define the concept of “value of a function”.

From today’s perspective, that seems to be surprising. However, if one is a


little bit familiar with the kind of analysis before Cauchy, if one knows that the
“Calculus of Expressions” set the academic standards before this time, the surprise
will disappear. Those who know the “Calculus of Expressions” only a little, will
128 10 Cauchy: The Bourgeois Revolutionary as Activist of the Restoration

be fully aware that the concept of value was of negligible relevance, playing only a
minor role.
Cauchy is the very first mathematician who explicitly introduces the concept of
“value” into the canon of fundamental concepts of analysis. This is a legacy of his
long practice as an engineer during which Cauchy noticed: without numbers (as
“values”), the algebraic expressions are of no use! This experience forced him to
include the notion of value, which had been neglected until then, and place it at the
centre of his theory.

The Political Significance of Cauchy’s Definition of This


Concept

Cauchy’s manner of determining the concept of “value of a function” is of the


greatest interest.
What does he say? This: the “value of a function” is everything that is demanded
by the existing conditions of the situation. Cauchy says: if the situation is not as
clear as daylight by itself (which means if the value cannot easily be calculated
directly), we must take all imaginable and FITTING values. This is contrary to
Bolzano’s arbitrary requirement for all the values to be somehow conceivable—
instead of only choosing those which fit to the given (existing) conditions. These are
the “limits”, all of them. Cauchy says: if we are not able to calculate the required
“value” directly, then we must exhaust all possibilities of the EXISTING conditions
completely—neither more nor less. This is still the generality of Euler’s approach.
Do not question the existing limits of thought—this could threaten the order of
the world! A greater contrast to Bolzano can hardly be conceived.

The Technical Significance of Cauchy’s Definition of This


Concept

Cauchy’s definition of the concept of “value of a function” is also in a purely


technical sense of greatest interest, for it is of mathematical substance. In other
words, it allows for a very simple but powerful proposition, namely the following
theorem:

Fundamental Theorem of Functions. If the function f (x) with the independent variable
x has a unique finite value for the finite value x = X, then it is continuous at this
value—and conversely.

We learnt from Bolzano what continuity in the “Calculus of Values” meant


(p. 108). Cauchy has precisely the same notion.
Some Very Surprising Consequences from Cauchy’s Concept of “value of a. . . 129

On p. 125, we already saw that the unique “limit” is the same as “continuity”,
and this is expressed exactly by the content of this theorem.—
A function is “continuous” for each value that is directly deduced from a
calculatory expression. Today, the technical proof of this elementary fact is shown
in every textbook of analysis. Its main idea is based on the fact that all values can
be determined from a calculatory expression—as the operations are continuous.
Let me add two remarks.
1. The name of this theorem is new. It seems to be suitable.
2. Until today, nobody else has formulated this theorem, not even Cauchy.
What is the reason for that? Why did nobody formulate (and prove) this theorem
which is true in Cauchy’s analysis until today? Let me name three aspects.
(a) Within Cauchy’s analysis, the theorem is trivial, a mere truism. One only has to
recall the strict analogy of the concepts “convergence”—resp. unique “limit”—
and “continuity”. However, something which is trivial does not constitute a
“theorem” for Cauchy!—Nevertheless, there is a formulation by Cauchy in a
letter which is exactly the proposition of the theorem (letter to Coriolis from
13. February 1837).
(b) As all mathematicians might have noticed already,
In today’s analysis this theorem is wrong!
The reason is that in today’s analysis, we use a concept of “value of the
function” which coincides with that of Bolzano. As Bolzano’s concept of “value
of the function” allows for a completely arbitrary determination of the value of
a function, this value cannot have any regular properties or nature. Therefore
the “Fundamental Theorem” cannot hold within the conceptual framework of
modern day analysis. It is just the conservative character of Cauchy’s analysis
which makes this theorem valid—more precisely, it is the stipulation of Cauchy
that the “values of the function” have to be deduced and that they are not allowed
to be chosen arbitrarily. It was only due to the revolutionary Bolzano who
mandated this, and we follow this conceptual framework until today, with all
the consequences.

(c) Taking both points into account we can conclude: not necessarily Cauchy,
but all those mathematicians, who deal with Cauchy’s analysis today, have
every reason for stating this theorem. It shows explicitly that and in which way
analysis has changed from Cauchy until today. The mathematicians who deal
with Cauchy’s analysis today are the historians of mathematics. It is they who
have to supply the information why they still refrain from pointing out this
fine theoretical change. Since the 1990s (when I published my interpretation of
this theorem), they have not risen to the challenge. More precisely, they ignore
these facts. (This in turn raises the question of the current state of the history
of mathematics as a science: what is its character today? Are the historians of
mathematics afraid of mathematics?—We shall return to this.)
130 10 Cauchy: The Bourgeois Revolutionary as Activist of the Restoration

Excursus: Preview of a Failed Revolution of Analysis in the


Years of 1958 and 1961

In the years 1958/61, analysis had the chance to revolutionize itself, to undergo a
conceptual upheaval comparable with the one caused by Cauchy. It was proposed
to found the new analysis on a different concept of number.
These ideas were published by two German authors and, quite independently,
by an American author. The latter was born in Germany to Jewish parents who
went into American exile in order to escape German fascism. He later became a
mathematics lecturer in the USA.
Cauchy’s example shows clearly that the detailed formulation of any analysis
also depends on number and, consequently, on the accepted concept of number. The
consequence is this:
The introduction of another concept of number—i.e. of numbers with other properties
than those used by Cauchy and ever since Cauchy—would change the configuration of
analysis. The form of the “Analysis of Values” would have to change accordingly.

This was exactly the issue when in 1958 the article An Enlargement of Calculus
by Curt Schmieden (1905–91) and Detlef Laugwitz (1932–2000) and in 1961 the
paper Non-Standard Analysis by Abraham Robinson (1918–74) were published.
Whereas Schmieden/Laugwitz spoke of a “proper enlargement” of calculus, Robin-
son declared a “proper enlargement of classical calculus” and at once came up with
the name for his new theory: Nonstandard-analysis.

History Does Not Recur, Not Even in Mathematics

But 1958 was not 1821. One and a half centuries after Cauchy’s textbook, it was
plainly impossible to reorganize mathematics as completely as Cauchy had done
(at least in large parts, with the exception of one aspect which will be mentioned
on pp. 135f). Analysis was established far too deeply, in thousands of heads and
hundreds of textbooks.
In 1821, the formation of analysis was the duty of a few prominent authorities.
In those days, active mathematicians who dealt with analysis knew of each other.
Nearly all of them lived in Paris or went there temporarily to do their years of
apprenticeship. But by 1958, analysis was a worldwide established subject, at
all scientific colleges and universities. Courses and syllabuses existed as well as
textbooks and teachers—who were all trained to think in the same analytical way.
Should all of them change their minds? Why? Who could direct or initiate this? Who
should execute or control this? Such an expenditure, and for what purpose?—The
traditional analysis was not wrong or impracticable. At worst, it was constructed
in a somewhat complicated way (see Chap. 14). This is one of the main reasons
used by the proponents of the nonstandard-analysis to persuade others to accept
Excursus: Preview of a Failed Revolution of Analysis in the Years of 1958 and 1961 131

their theory. But only to change the way of thinking, maybe to simplify it, and
this without producing new results? Of course, this is not enough to overthrow a
worldwide church—moreover in a field where the mindset is canonized in all detail
in an unparalleled manner. It would have required at least a new truth in order to
give such a revolution at least a tiny chance of success. However, nonstandard-
analysis did not offer any new truth—but only the old in a new guise. The hope
of some pioneering rebels to persuade by pointing to the existence of some new
mathematical subjects (e.g. to delta-functions, even rational ones!—see p. 240)
remained unfulfilled. Delta-functions are a fascination only for specialists.
Under these circumstances a new revolution of analysis in the last third of the
twentieth century did not occur. The tanker Analysis had grown far too sluggish and
big and could not be reversed by individuals or by small groups.
As soon as a system of thought is broadly established as well as institutionalized in a
society—and in this way becomes universally valid—, it is impossible for a group of
persons to take over. Majority is power when there is no absolute rule.

To put it ideologically, this is not a question of mathematical truth whatsoever.


Nonstandard-analysis is neither more nor less “true” than standard-analysis (as the
commonly used form of analysis is called now, for the sake of precision). Just as
Cauchy’s “Calculus of Values” is neither more nor less “true” than the “Calculus of
Expressions” from Euler and Lagrange. The “Calculus of Values” is more useful in
respect to the actual needs of the time (technical usage) than the former construction.
However, it is not “more true” in any possible sense. The proofs of its theorems are
by no means more precise than the proofs of the earlier theorems.
Unfortunately, historians of mathematics very often state the opposite of this,
today more than ever before. (Exceptional authors like Henk Bos or Kirsti Andersen
only confirm the rule.)
Of course, this is nonsense! Why should we, today, be qualified to think more
precisely than our ancestors? There is no obvious reason for this. Majority is
not law (or truth). Those who judge in this presumptuous manner are either too
lazy to familiarize or not capable of familiarizing themselves with another way of
thinking—as we are attempting in this book.

A Rebellion of Nonstandard-Analysis?

Many articles dealing with Cauchy’s analysis suddenly appeared in the 1970s and
1980s. The reasons for this can only be guessed. One hypothesis is that this hype
was a last attempt by the proponents of nonstandard-analysis to justify their theory
by trying to place it within a historical tradition. And this in place of offering some
new truth: history instead of mathematics?
The idea was: if it could be Nonstandard-analysis is the “true”, the “right”
understanding of analysis, then it would show its superiority—at least from a moral
point of view.
132 10 Cauchy: The Bourgeois Revolutionary as Activist of the Restoration

This line of argumentation drew on history. The master plan was to prove that
some first-rate mathematicians were ESSENTIALLY doing nonstandard-analysis!—
an idea, no doubt.
The first step was very easy. All proponents of analysis living before the twentieth
century liked to use the attribute “infinitely small”. However, this was exactly
the novelty of Nonstandard-analysis, its characteristic distinction from standard-
analysis:
In Nonstandard-analysis there do exist “infinitely small” (and similarly also “infinitely
large”) numbers—which do not exist in standard-analysis.

(We will return to this, see pp. 226f, especially p. 227.)


That is why at first sight it was easy for the nonstandard-analysts to live off the
reputation of their famous forerunners. In respect to Leibniz, the fine details of his
concepts had still not been published, and his phraseology “infinitely small” could
therefore be interpreted very freely: namely in a modern nonstandard-analytical
way. With Euler and Johann Bernoulli this could be done effortlessly, for they really
talked about such “infinite” numbers—we have read this! Consequently, they could
easily be co-opted as being essentially proponents of Nonstandard-analysis. It did
not matter here that neither Euler nor Johann Bernoulli made any attempt to explain
how such a concept of number could be explicated: as really great mathematicians
must think in the right way—there should be no problem!
Obviously, this practice deviates completely from doing history of science,
as it is undertaken here. (By the way, even today there exist fundamentalists of
history of mathematics who are unable to understand this way of thinking. A
rational discussion with them is completely impossible. —Do you imagine such
a controversy in mathematics?)
It was of great help when philosophers started to join the debate. Thus, the
philosopher of science Imre Lakatos (1922–74) subscribed to the opinion that
Cauchy’s style of thought was a nonstandard-analytical one. Lakatos’ in some way
risqué arguments could be turned into merely stringent ones (which helped me
to some renown after having written a book on it). As a result, the youngest of
the triumvirate which started Nonstandard-analysis in the years 1958/61, Detlef
Laugwitz, took the plunge in construing Cauchy as a true nonstandard-analyst,
who actually thought within these concepts. Laugwitz succeeded in giving many
beautiful mathematical arguments, but he failed in his intent, for:
The interpretation of Cauchy’s “Calculus of Values” as an (early) state of Nonstandard-
analysis can be proven to be false.

The reason of this will be explained shortly.


After Leibniz, Euler and Johann Bernoulli—as just stated—, there still remained
Cauchy. Needless to say, Cauchy also uses the attribute “infinitely small”. Yet, there
is something else!
Excursus: Preview of a Failed Revolution of Analysis in the Years of 1958 and 1961 133

A Digression Within the Excursus: Looking Back at a Criticism


of Cauchy

Five years after Cauchy’s textbook of analysis appeared, a claim of the young
mathematician Niels Henrik Abel (1802–29) was printed which says that Cauchy’s
book contains a false theorem. (This will be spelled out in some detail on pp. 140f.)
However, there was no direct reaction from Cauchy to this reproach.
Consequently, the analysis of the year 1826 was faced with a factual problem: is
this theorem true or false? Cauchy had stated and proved it—Abel had pronounced
it wrong. Two nowadays still very much appreciated mathematicians contradicted
each other in their judgement.
24 years later, a treatise from Philipp Ludwig Seidel (1821–96) was published
which also criticized that theorem of Cauchy. It also stated and proved an alternative
theorem. Again, there was no reaction from Cauchy.
However, three years later (in 1853), Cauchy published a treatise which
contained—in some hidden way, which we will examine later—his self-defence
in this matter. Because Cauchy did not explicitly label this statement to be a
“self-defence”, it remained unnoticed. Only much later, when mathematicians
finally realized the significance of Cauchy’s treatise of 1853, they started to discuss
it (p. 140).

Back to the Upheaval of Nonstandard-Analysis

With the birth of Nonstandard-analysis 1958/61, a quarrel had started which finally
reached a climax in the 1970s and 1980s. Today this controversy is completely over
and this to such an extent that it could not offer any contribution to a new idea
concerning Cauchy’s theorem. The controversy is:
Which of the two alternatives is true?
1. Cauchy stated and proved in his textbook a (very elementary) false theorem.
2. Cauchy’s theorem is true—because by “convergence” Cauchy meant not the
same as we do today, but instead he meant what we call nowadays “uniform
convergence”.

The first alternative is not plausible at all. Inevitably, every mathematician makes
mistakes. But a mistake such as this? A mistake regarding a theorem studied in the
first course of analysis—which did not lead to a proper defence from its author, and
this even after the publication of a criticism? Who should believe in this unlikely
narrative?
Subsequently, the nonstandard-analyst s felt newly empowered: they simply
changed the interpretation of the decisive concept of this theorem from Cauchy’s
(for him it is “convergence”) to mean now “uniform convergence”—and that the
theorem in this new meaning is correct is disputed.
134 10 Cauchy: The Bourgeois Revolutionary as Activist of the Restoration

However, the nonstandard-analysts clearly remained a clear minority. Therefore,


the alternative (1) remained the dominant opinion and Cauchy became the most
popular half-wit in the textbooks of analysis.

Walking on Very Thin Ice

The question for the nonstandard-analysts was: how could they justify changing
the interpretation of Cauchy’s concept? Their idea showed great promise. They
argued: Cauchy did not understand by “number” what everybody else (for at
least one hundred years) thought of the concept but, quite contrarily, what we, the
nonstandard-analyst s mean by it TODAY!
From the mathematical point of view, they had succeeded. The reason being:
the range of numbers within Nonstandard-analysis is considered to be definitely
greater than the range of real numbers, say: the decimals (however, see chapter 14).
And in the theorem under consideration Cauchy makes a certain supposition: he
demands that a “series” is “converging” for all values (and he meant for all real
numbers). However, if by the “for all numbers” he meant more than we thought
until now (namely besides the “real” numbers also the “hyper-real” numbers), then
this surplus in the assumption allows for a surplus in the conclusion, and that is,
indeed, the claim of the theorem.
This mathematical argument is indubitably correct. Robinson had published
this reasoning already in 1963, and Lakatos, this ingenious propagandist of his
own ideas, exaggerated its importance, especially in philosophy of science. (To be
honest, once I joined in this enterprise.)
Yet, the decoding and interpretation of former mathematical concepts is neither
the duty nor a self-evident competence of a pure mathematician! Instead, it is a main
task for the historians of mathematics.
However, a detailed historical investigation of the concepts used by Cauchy
within his analysis was not forthcoming. Nobody undertook an inquiry of Cauchy’s
concept of number to check whether he really considered such exotic entities like
“infinitely small” numbers. And, of course, Cauchy did not do that!
In any case, Nonstandard-analysis needs to perform some acts of conceptual
acrobatics in order to construct “hyper-real” numbers in a mathematically accept-
able way. Here Robinson’s construction stands out, but without studying (at least)
one semester of modern logic, one is not able to follow his construction. It is,
however, unlikely that Cauchy should have anticipated such a stilted concept in
1821 in any possible sense. For this, we need a further supporting argument! But
nobody has supplied one: instead there were plenty of rhetorical fireworks!
In short, this attempt by some nonstandard-analysts to justify their theory by
recourse to the development of analysis remained without historical substantiation.
(This is true, although at the time of my dissertation, around 1980/81, and some
years later, I was convinced of the opposite—and so were the referees of my thesis,
who were mathematicians.)
Cauchy’s Concept of Convergence: A Big Misunderstanding 135

Not earlier than 1990, I undertook a thorough examination of those concepts


which Cauchy really used. My result can be summarized as follows: if you have
to choose between two possibilities, just take the third! Or, to put it somewhat
more technically, the translation of Cauchy’s concept of “convergence” in modern
analysis is neither “convergence” nor “uniform convergence”—but a third notion.
This will be explained in the next section.

Cauchy’s Concept of Convergence: A Big Misunderstanding

Strictly speaking, the discussion of Cauchy’s concept of convergence belongs to the


section which is called “some very surprising consequences from Cauchy’s concept
of ‘value of a function’”, but because of its huge importance it will be considered
separately.

A Mystery of History: Cauchy’s Concept of Convergence

There is no other concept of analysis which caused such long, comprehensive,


and vehement controversies and which also went beyond the field of history of
mathematics, than Cauchy’s concept of “convergence”. Why is this?
The decisive point is: although Cauchy’s analysis is shaped very revolutionarily,
it remains firmly tied to established thinking—which is thoroughly alien to us today.
Let us return to Cauchy’s concept of function (p. 123). Which basic notion does
he use? It is “quantity”. We should bear in mind:
Cauchy’s first and unique foundational concept of analysis is “quantity”. (From that he
also deduced the concept of “number”.)

As a consequence, Cauchy determined “convergence” for those objects, exclu-


sively for quantities.
Originally Cauchy’s concept of “convergence” applies to “quantities”.

(To overlook this has been the decisive mistake in the mathematical–historical
debate following the creation of Nonstandard-analysis and its efforts to co-opt
Cauchy since the early 1960s. However, today no mathematician is able to give an
account of this concept: “quantity”. Cauchy’s definition of “quantity” can be read
on p. 118.)
Here is Cauchy’s definition of “convergence”:
We call a series an indefinite sequence of quantities,

u0 , u1 , u2 , u3 , ... ,
136 10 Cauchy: The Bourgeois Revolutionary as Activist of the Restoration

which follow from one to another according to a determined law. These quantities
themselves are the various terms of the series under consideration. Let

sn = u0 + u1 + u2 + . . . + un−1

be the sum of the first n terms, where n denotes any integer number. If, for ever increasing
values of n, the sum sn approaches a certain limit s, the series is said to be convergent, and
the limit in question is called the sum of the series.

The uk are “quantities”, e.g. “functions” fk (x) (Cauchy still uses the comma in
its traditional meaning of the plus sign):

f0 (x) + f1 (x) + f2 (x) + f3 (x) + . . .

We realize that, in the case of series, Cauchy deviates from his usual way of
signification; in this case, the index does not indicate a “value” but instead belongs
for practical reasons to the name of the “quantity”. (Here the name “f ”—in Cauchy:
“u ”—, supplied with an index, does not denote a variable! It is merely a “variable”
name.) In regard to Euler and Lagrange we have already seen how difficult it is to
deal with series without the index notation: pp. 75 and 95.
Starting from a “sequence”, Cauchy now constructs finite sums sn (x):

sn (x) = f0 (x) + f1 (x) + f2 (x) + . . . + fn−1 (x) .

If the sn (x) approach a “limit”, this is called “convergence”. Written in Cauchy’s


way, we have

lim sn (x) = s(x) .

Before we proceed with the technicalities, we need to discuss an emerging


problem: Cauchy speaks of a “limit”. However, a “limit” is a value. Yet this s(x)
is a quantity (as can be seen by the “x”), no value. Notwithstanding, Cauchy speaks
of the “limit”! The only possible explanation is that Cauchy means “the value of
lim sn (x) for a x = X ”!
Consequently, Cauchy’s definition of “convergence” is this:
A series of functions

f0 (x) + f1 (x) + f2 (x) + f3 (x) + . . .

is called “convergent” for the value x = X, if for increasing n the sum

sn (x) = f0 (x) + f1 (x) + f2 (x) + . . . + fn−1 (x)

has for x = X a unique “limit”.


Cauchy’s Concept of Convergence: A Big Misunderstanding 137

It is possible to reformulate this definition. Instead of “lim sn (x) = s(x)”, we


may write “lim(sn (x) − s(x)) = 0”, couldn’t we? Using Cauchy’s notation for this
difference, which is still in use today, “rn (x)” (“reste” = “rest”), we get:
A series of functions

f0 (x) + f1 (x) + f2 (x) + f3 (x) + . . .

is called “convergent” for the value x = X, if for that value the special sum (the “rest”)

rn (x) = fn (x) + fn+1 (x) + fn+2 (x) + . . .

we have

lim rn (x) = 0 .

The Solution of the Mystery

Until 1990, there has been a consensus among all Cauchy specialists about this
formulation. But since 1990, there has existed a new idea, arising from two open
questions:
1. What IS “ rn (x)” for Cauchy?
The only possible answer is a “variable”, depending on the TWO “independent”
variables n AND x.
2. What IS “ lim rn (x)” for Cauchy?
Again the answer is simple: lim rn (x) as a mathematical notion in Cauchy’s
analysis is a “value of a function” and, consequently, “the limit or the limits”—
to Cauchy: all of them!—of the variable rn (x) (for the value x = X).
And that is the point! By “lim rn (x) for the value x = X”, Cauchy not only means

lim rn (X) ,
n→∞

but moreover, there are also all “limits” to be included:

lim rN (xk ) , ( )
N→∞
xk →X

for rn (x) is a function of the two variables n and x.


Cauchy never writes subscripts when he uses “lim”. Only later did it become
customary. The example above shows why it was superfluous for Cauchy: he always
means all possible limits. Our example also shows that this notation—which today
is a must—is generally somewhat laborious. Thus, as Cauchy usually means all
limits, he can use “lim” unindexed.
138 10 Cauchy: The Bourgeois Revolutionary as Activist of the Restoration

The Mathematical Significance of This Solution

We remember Cauchy’s theorem and that it has been considered false (except by
nonstandard-analysts) until today. It reads literally (only the sign “ X ” is added two
times):
Theorem I. When the various terms of the series

u0 , u1 , u2 , ... , un , un+1 , ...

are functions of the same variable x, continuous with respect to this variable in the
neighbourhood of a particular value X for which the series converges, the sum s of the
series is also a continuous function of x in the neighbourhood of this particular value X.

In the special literature, sometimes the name “Cauchy’s Sum Theorem” is used.
The decisive assumption of this theorem is that the series of functions “con-
verges” and this also “in the neighbourhood” of the particular value X.
Today we can offer this assumption in the terms of all three rivalling interpreta-
tions:
1. The standard-analysts understand Cauchy using the concepts which are custom-
ary today. Which means they say “ lim rn (X) = 0”.
n→∞
2. The nonstandard-analysts understand Cauchy using their concepts and say
“ lim rn (X ) = 0 for all values of X in the neighbourhood of the value X
n→∞
in question.”
3. According to the new interpretation, obtained from an investigation of Cauchy’s
concepts, we reach the conclusion that Cauchy means what is formulated in
line . (By “neighbourhood”, Cauchy means “X + α” instead of other “values”
X = X.)
Stated in the terminology of modern analysis, this amounts to the following:
1. In his theorem, Cauchy presupposes the “convergence” of the series and
consequently the theorem is wrong.
2. In his theorem, Cauchy assumes the “uniform convergence” of the series and
consequently the theorem is true.
3. In his theorem, Cauchy demands the “continuous convergence” of the series
and consequently the theorem is true.
The exact definition of “continuous convergence” can be found in some text-
books of the theory of functions, published since 1921. For the understanding of the
historical development, it is not necessary to understand this definition; it is enough
to accept the following box.
In case (1), Cauchy looks bad. In case (2), the standard-analysts look bad. In
case (3), after having investigated Cauchy’s concepts, the historian of mathematics
reaches the judgement: Cauchy did not make a mistake here, but he proves
something other than the standard- as well as the nonstandard-analysts think.
Cauchy’s Concept of Convergence: A Big Misunderstanding 139

In the terminology of modern analysis, we state the following (“⇒” means “leads
logically to”, and not one of the following inversions is true):
continuous convergence ⇒ uniform convergence ⇒ convergence

The weakest presupposition (on the right side) is not strong enough to prove
the theorem. The presupposition in the middle is stronger and suffices—however,
there is no clear evidence in Cauchy’s text that he might have meant this. The
strongest presupposition (on the left) is Cauchy’s true condition—and it is all the
more sufficient to prove his assertion.

Cauchy’s Proof of His Theorem

We finish our argument by replicating Cauchy’s proof of his theorem. The proof is
simple.
The claim is that the sum s(x) is continuous at a value x = X. Continuity means
s(x + α) − s(x) decreases (for the value x = X) together with α and becomes
arbitrarily small. This has to be proven:
 
lim s(x + α) − s(x) = 0 for lim α = 0 and for the value x = X .

This is a consequence which follows directly from the usual representation of s as


s = sn + r n :

s(x + α) − s(x) = sn (x + α) − sn (x) + rn (x + α) − rn (x) . (∗∗)

Since the sn are finite sums of continuous functions, sn (x + α) − sn (x) is a finite


sum of infinitely small quantities and, consequently, infinitely small. Only the two
other summands remain rn (x + α) and rn (x)—not to forget, in each case for the
value x = X.
However, we have lim rn (x) = 0 (for the value x = X): this is exactly the
presupposed “convergence” of the series of functions for the value x = X.
Similarly, we know lim rn (x + α) = 0 for the value x = X, because of Cauchy’s
concept of “value of a function” —here for the value X! For we know “f (X)” are
all limits lim f (X ) for X → X, and this can be written as lim f (X + α) for α → 0,
or just for our function rn (x): lim rn (X + α) for α → 0. This is required to be = 0
because of the assumed “convergence”—in Cauchy’s sense!—FOR THE VALUE X.
End of the proof.
The very last step has always been controversial:

1. The standard-analysts read “x + α” as VALUES different from X and say Cauchy


cheats by assuming “convergence” for values different from the value X—but
this is not permitted. However, for Cauchy “x + α” is always a “variable” and
never a “value”!
140 10 Cauchy: The Bourgeois Revolutionary as Activist of the Restoration

2. The nonstandard-analysts argue no problem, α is an “infinitely small” num-


ber—and these are included in Cauchy’s presupposition of “convergence”
even in the neighbourhood. And this fits!—however, not to Cauchy’s way of
thinking.

It is impossible clearly to demonstrate the concept of “infinitely small” number


within Cauchy’s texts. But Cauchy offers a concept of “value of a function” (in
detail, all the limits!) nobody had ever thought of. Today this concept is unknown,
and before 1990 nobody read Cauchy’s textbook very carefully.1 And because of
Cauchy’s concept of “value of a function”, “lim rn (x +α) = 0 for the value x = X”.
This is guaranteed by the presupposition of “convergence (at the value x = X)”.
—Bingo!

Cauchy’s Self-Defence

As stated earlier, Cauchy later published a treatise in which he answered his critics.
He did so in a very polite manner and did not name them (nor their criticism)
explicitly. Niels Henrik Abel had criticized Cauchy’s theorem in all detail in 1826
(see p. 133). Abel declared the theorem to be wrong. However, he did not prove this,
but merely claimed that a particular function is a counter-example (see p. 156).
In a treatise from 1853 Cauchy offers a calculation which proves Abel’s claim to
be false.
In place of Cauchy’s somewhat demanding calculation,2 we transcribe it and
Abel’s counter-example into a simpler version.
Abel says something like the following. The series

x 0 + (x 1 − x 0 ) + (x 2 − x 1 ) + (x 3 − x 2 ) + . . .

converges for each value of x between 0 and 1 inclusively. Its single terms x n+1 −x n
are continuous. However, the whole sum is not, and therefore it is a counter-example
to Cauchy’s Sum Theorem of p. 138!
Let us examine this argument. Abel is correct in saying that the terms x n+1 − x n
of this series are continuous, but that the sum is not. Why the latter? If we let 0 
x < 1, this sum has the value 0, for we have

sn (x) = x 0 + (x 1 − x 0 ) + (x 2 − x 1 ) + . . . + (x n − x n−1 ) = x n ,

and consequently (for x = X and since 0  X < 1)

1 Not even the translators Robert E. Bradley and C. Edward Sandifer in 2009 did!
2 See Die Analysis im Wandel und im Widerstreit, p. 352.
Cauchy’s Concept of Convergence: A Big Misunderstanding 141

s(X) = lim sn (X) = lim Xn = 0 . (††)


n→∞

However, for x = 1, the series is plainly one:

s(1) = 1 + (1 − 1) + (1 − 1) + (1 − 1) + . . . = 1 .

Subsequently, this function s(x) is discontinuous for the value x = 1, because for
X < 1 we have from ††:

lim s(X) = lim 0 = 0 ,


X→1

whereas we have s(1) = 1 = 0. In Abel’s mindset, this disproves Cauchy’s theorem.


Nevertheless, this does not follow from Cauchy’s way of thinking! Quite
contrarily: Cauchy is able to prove that this series of functions does not “converge”
for the value x = 1—namely in the special sense in which Cauchy had established
the concept of “convergence”! But if this series does not meet the presuppositions
of his theorem, his theorem cannot state anything about it.
To conclude, we will show that our example of a series for the value x = 1 does
not “converge” in Cauchy’s understanding of that notion. This is quite easy. We only
have to study lim rn (x) for the value x = 1. We have

rn (x) = (x n+1 − x n ) + (x n+2 − x n+1 ) + (x n+3 − x n+2 ) + . . . = −x n .

Is lim rn (x) = −x n for the value x = 1 also = 0?—No, for the value x = 1, we get
lim rn (1) = lim −1n = −1 = 0. Quite simple!
Abel’s counter-example (which is technically somewhat more complicated)
cannot be refuted that easily. An additional consideration is needed. This additional
consideration can also be explained by using our simple example (the original
argument was given by Cauchy in 1853). It runs like this. Instead of taking x = 1,
we investigate the lim rn (x) for x = 1 − α and lim α = 0. All the values 1 − 1/n are
< 1 and we have lim (1 − 1/n) = 1. Therefore lim rn (1 − 1/n) is also a “value of
n→∞ n→∞
the function” rn (x) for x = 1. However, it is (compare with the formula on p. 77,
for k = 1, we get a = e, and also, choose x = −1 resp. i = − ω1 = N):
 N
lim rN (1 − 1/N) = lim − 1 − 1
N = −(e−1 ) = 0 .
N→∞ N→∞

A completely flawless calculation—which constitutes in Cauchy’s analysis the


proof that this chosen series, in his understanding does not “converge” for the value
x = 1! (Cauchy only skipped the part “lim rn (x) =” in his calculation—this step
he assumed to be clear to his learned readers.) Clearly, the meaning of Cauchy’s
calculation can only be understood if one relies on his concept of “convergence” as
well as—and at least equally importantly!—his concept of “value of a function”.
142 10 Cauchy: The Bourgeois Revolutionary as Activist of the Restoration

However, his critics have refused to do this: in the nineteenth, in the twentieth, and
(until now) in the twenty-first century.
By the way, this example of a series (just like that which Cauchy had provided)
is not “uniformly convergent”. That is why it (as well as Cauchy’s series), sadly, is
not suited to refute the claim of the nonstandard-analysts, that Cauchy had thought
of his “convergence” in the same way as them, mathematically. That is a pity,
because a technical argument would be considered by many as mathematically more
conclusive than a merely philosophical one.

What Are the Reasons for the Prevailing Misunderstanding of


Cauchy’s Notion of Convergence?

Why do modern mathematicians not understand Cauchy’s concept of “conver-


gence”? The brief answer is: because they are no historians of mathematics. A more
detailed answer consisting of two parts is given below:
1. Cauchy has defined this concept for “quantities”—whereas we (in analysis)
define it exclusively for “numbers”.
2. Cauchy’s concept of “value of a function” is quite different from that which we
use today (p. 124), a fact that has stayed unnoticed until now.
We have reached the result: if we take Cauchy at his word, the quarrel about his
theorem which has now lasted for more than 96 years has finally been resolved.
Strangely enough, today nobody is interested: each publication of this explication
remained without professional resonance. Will this change now?
Today it is even possible to accuse Cauchy—by ignoring his conceptual limits—
of new “errors” in his analysis and to publish this in a scientific journal viewed to
be first rate, like Historia Mathematica. And even worse, the named journal refused
to publish a criticism of such an “erroneous” article. (The editor then decided,
relying on the—clearly contra-factual—argument, this journal would not publish
“Letters to the Editor”. However, I did write an article, not a letter, but it was not
even reviewed.) From this, we may conclude that history of mathematics is not a
separate subject, but only the servant of ideological warfare. It cannot be part of an
independent scientific discourse.

Cauchy’s Concept of Derivative—Again a Misunderstanding

The notion “derivative” was introduced by Lagrange. Lagrange defined it as a


“function” f  (x) which can be determined from the series expansion of the function
f (x) by changing the argument, i.e. from the series expansion of f (x+a) (formula §
on p. 96).
Cauchy’s Concept of Derivative—Again a Misunderstanding 143

Instead of working with “derivatives”, Euler operated with “differential quo-


tients”, which he took to be true quotients made from two “infinitely small”
quantities. This is a complicated notion and was therefore avoided by Lagrange.
As we have read in the preface of his textbook on analysis, Cauchy made a point
of “making all uncertainty” of the formulae “disappear” by the “determination of the
values”. Cauchy follows this principle also in case of the derivative. He introduces
a formulation which is still, today, presented to the beginners—even though, again,
in another meaning.
Cauchy refers to the fact that for a “continuous” function f (x), an “infinitely
small” increase α of the variable x (i.e. he takes x + α and demands lim α = 0)
produces an “infinitely small” increase of the function. Cauchy likes to denote these
“increases” by capital delta (
), i.e. α =
x as well as f (x +
x) − f (x) =
y.
We read Cauchy (by “the two terms”, he always means numerator and denominator
of the fractions):
By consequence, the two terms of the ratio of differences


y f (x +
x) − f (x)
=

x
x
will be infinitely small quantities. But, while these two terms indefinitely and
simultaneously will approach the limit zero, the ratio itself may be able to converge to
another limit, either positive or negative. This limit, when it exists, has a determined value
for each particular value of x; but, it varies along with x.

y
Cauchy defines this “value” of lim
x for lim
x = 0 to be the “value” of a new
“function”:
The form of the new function of the variable x, which will serve as the limit of the ratio

y f (x+
x)−f (x)

x =
x , will depend on the form of the proposed function y = f (x). To
indicate this dependence, we give to the new function the name of derived function, and we
represent it, with the help of an accent mark, by the notation

y or f  (x) .

As already stated above, even today the “derivative” (as we call the “derived
function” of Lagrange and Cauchy today) is introduced to beginners in this way.
However, as Cauchy’s thinking basically differs from that of ours today, we must
add two things.
1. To say only the least, in Cauchy’s way of thinking, this manner of defining
a “function” is problematical because Cauchy establishes a “function” to be
a “variable” which has certain properties (see p. 123). However, the “derived
function” is not defined to be a certain variable, but it is constructed as a value-to-
value relation. (At most one may, in hindsight, call that value-to-value relation,
defined by Cauchy, a “variable”.)
144 10 Cauchy: The Bourgeois Revolutionary as Activist of the Restoration

The manner of defining a “function”,

X −→ f  (X) ,

is very modern (to be precise, this is the method we use today), but it definitely does
not fit to the concept of “function” given by Cauchy!

In other words, by giving this notion of the “derived” function, Cauchy violates
his own conceptual framework. Strictly speaking, in Cauchy’s thinking, his
“derived function” is no “function” at all!
2. What is Cauchy’s precise definition of “derivation”? It is that he determines
which is the (uniquely defined!) “value” f  (X) for the “value” x = X—namely
the value which is the “limit of f (x+
x)−f

x
(x)
”.
f (x+
x)−f (x)
The value x = X produces the value lim
x for lim
x = 0.

However, we must also take into account that Cauchy takes each value X (also)
as a limit lim xn = X! In other words, Cauchy does mean the following:
f (xn )−f (x̄n )
The value x = X produces the (unique!) value lim xn −x̄n for xn = x̄n ,
lim (xn − x̄n ) = 0 (with lim xn = X = lim x̄n ).

Nonetheless, in today’s language, this is the definition of “continuous derivabil-


ity”! This entails that Cauchy’s definition of derivability demands that the derived
function f  (x) is continuous. (And that even though his formula is the same as
ours!—However, we already know that Cauchy’s concept of “value of a function”
differs from that of today—and this must have mathematical consequences. Math-
ematics involves not just formulae but also the interpretation of the formulae. In
Cauchy’s words: we must make the “uncertainty” of the formulae “disappear”.)
After having realized that Cauchy’s concept of convergence is equivalent to
today’s concept of continuous convergence, we shall find it evident that Cauchy uses
also the modern concept of “continuous derivability” instead of our “derivability”.
Let me add two remarks, a historical one and a mathematical one.

(a) The historical one: Laugwitz argued already in 1987 that Cauchy’s “derivability” is
our modern “continuous derivability”. His argument was different and I was unable to
accept it. Unfortunately, even in Die Analysis im Wandel und im Widerstreit, pp. 333–
335, I had not understood Cauchy correctly.
(b) The mathematical one: Strangely, in today’s lectures for beginners, it is not mentioned
that the formula
f (xn ) − f (x̄n )
lim
xn →x0 xn − x̄n
x̄n →x0
xn =x̄n

determines the “continuous derivability” of the function f at the value x0 . In 1965,


Paul Lorenzen (1915–94) called this concept, which was introduced by Guiseppe Peano
(1858–1932) in 1892, “free derivability”.
Cauchy’s Concept of the Integral 145

Cauchy’s Concept of the Integral

In the “Algebraic Analysis”, the integral was introduced as the inverse operation of
differentiation, i.e. as the “indefinite integral”—and, consequently, as a “function”.
Cauchy ends this tradition by introducing the “definite integral”—which is to say
the integral as a “value”.
We have already studied Leibniz’ method of calculation of the area when one side
is curved (pp. 29f) and seen that this method of Leibniz has only been known since
the late twentieth century. Subsequently, if Cauchy proceeds in a similar manner, he
could not have been influenced by Leibniz (or one of his followers).
This similarity is due to the visualization of the geometrical construction, not to
the concepts.
• Where Leibniz takes a “curved line”, Cauchy deals with a “function” f (x) of
an independent variable x.
• Where Leibniz introduces “points of division” Bk on the horizontal below,
Cauchy sets forth “values of division” xk in the assumed interval between x0
and X.
• Where Leibniz uses “straight lines” Dk Bk , Cauchy relies on “values of a
function” f (xk ).
• Both take different quantities for their approximation:

– Leibniz takes an “area” which nearly fits, namely the rectangles Bk Pk .


– Cauchy decides to choose the suitable approximation (xk+1 − xk ) · f (xk ),
which is a “value”.

• And of course, where Leibniz proves that his approximate area differs from the
true area by a quantity which can be made smaller than any given quantity,
Cauchy defines the analytical concept of the “definite integral” as a “limit”:

X
n−1
f (x) dx = lim (xk+1 − xk ) · f (xk ) .
n→∞
x0 k=0

Thereby Cauchy assumes f (x) to be a “continuous” function. (The subscript


“n → ∞” is of course missing in Cauchy.)
To define something is an easy exercise. The real difficulty is to prove that the
definition makes sense. In our case, “making sense” is to show that the defined
“limit” is well-defined and unique—in other words, that this limit (a value) does
not depend on the method of calculation, that there are many ways to arrive at that
limit or, to be precise, that each division x0 , x1 , x2 , . . . xn = X of the interval from
x = x0 to x = X produces the same value.
In other words, there are two different approaches with their own intricacies.
Whereas Leibniz had to deal with the estimation of the greatest possible error,
146 10 Cauchy: The Bourgeois Revolutionary as Activist of the Restoration

Cauchy is concerned with comparing those limits that result from different divisions
of the interval and subsequently has to prove that they are all equal.
We may skip the details here, for they can be looked up in any modern textbook
of analysis. We shall only examine the structure of Cauchy’s argumentation.3

Cauchy’s Basic Idea in His Proof of the Existence of the


Definite Integral

The decisive step of Cauchy’s proof is the validity of this equation:


n−1
(xk+1 − xk ) · f (xk ) = (X − x0 ) · f (x0 + θ · (X − x0 )) .
k=0

• On the left, we have the sum of products (xk+1 − xk ) · f (xk ). Geometrically,


these are the “steps” below the graph of the function, in case of “increasing”
functions as those by Leibniz.
• On the right, we have just one product, consisting of the length of the whole
interval and the value of the function at the value x = x0 + θ · (X − x0 ), where
θ is an unknown value between 0 and 1. For such a θ , x = x0 + θ · (X − x0 )
is any value inside the interval from x0 to X. —That this is possible, i.e. that a
value with this property really exists, is guaranteed by the so-called (and very
famous) “Intermediate Value Theorem”. Bolzano formulated this theorem very
precisely and proved it in 1817. We did not touch on this part of Bolzano’s
treatise. The decisive point is that the Intermediate Value Theorem is only valid
for “continuous” functions. That is why Cauchy demands the continuity of f (x)
X
in his definition of the definite integral f (x) dx .
x0

Let us summarize. A (finite) sum of products (xk+1 − xk ) · f (xk ) may be replaced


by a single product which is made from the length of the interval and some value of
the function inside of the interval.
The problem with this proof is the comparison of different divisions of the
interval such as x0 , x1 , x2 , . . . xn = X and x0 , x1 , x2 , . . . xm
 = X. The idea is

to “refine” each of these divisions in such a way that both refinements coincide.
This is not hard but a little bit technical.
For each part of such a common “refinement”—say x0 , x1 , x2 , . . . xl = X—we
have the above equation and in this way is the sum of products with xi -intervals
transformed to a single product of length (xk+1 − xk ).
The rest goes without saying.

3 It is shown in Die Analysis im Wandel und im Widerstreit, pp. 339–343.


Literature 147

One cannot deny that Cauchy’s proof of that useful equation which transforms a
sum of products into a single product (with the length of the interval as one factor)
is remarkable. To do so, Cauchy relies on a very general theorem regarding the
estimation of quantities which he had already stated in his first textbook of analysis.
(His textbook on the calculus of integrals was printed only eight years later, in 1829.)
When adapted to a form which is required for the proof of the existence of the
definite integral, this theorem reads:
Theorem. Let b, b , b . . . denote n quantities of the same sign and
a, a  , a  , . . . be the same number of arbitrary quantities, then we have

αb + α  b + α  b + . . . = (α + α  + α  + . . .) · M(b, b , b , . . . ) ,

where M(b, b , b , . . . ) denotes any quantity in between the greatest and the
smallest of the b, b , b , . . ..

It is very clear that “quantity” is a forceful basic concept in Cauchy ’s analysis.

What Is Cauchy’s “x ”?

A final summary:
Cauchy splits the formerly unique “x” of the old “Calculus of Expressions” for his
new “Calculus of Values” into two:
1. a small “x” without index to denote the “independent” variable,
2. a small and indexed “x” or a capital (and upright) “X” to denote a “value” of
the independent variable.
In “determining precisely” these conditions he makes “all uncertainty disappear”.

Literature

Cauchy, A.-L. (1821). Cours d’Analyse. In Œuvres Complètes, vol. 3 of series II. Paris: Gauthier-
Villars, 1897. http://gallica.bnf.fr/ark:/12148/bpt6k90195m.r=cauchy+oeuvres.langFR. Ger-
man: Itzigsohn 1885.
Cauchy, A.-L. (1823). Resumé des Leçons données à l’École Royale Polytechnique sur le Calcul
Infinitésimal. In Œuvres Complètes, vol. 4 of series II, pp. 5–261. Paris: Gauthier-Villars, 1899.
http://gallica.bnf.fr/ark:/12148/bpt6k90196z.r=cauchy+oeuvres.langFR.
Cauchy, A.-L. (1829). Leçons sur le Calcul Différentiel. In Œuvres Complètes, vol. 4 of series
II, pp. 265–572. Paris: Gauthier-Villars, 1899. http://gallica.bnf.fr/ark:/12148/bpt6k90196z.r=
cauchy+oeuvres.langFR.
Cauchy, A.-L. (1833). Résumés analytiques (Turin). In Œuvres Complètes, vol. 10 of series
II, pp. 7–184. Paris: Gauthier-Villars, 1895. http://gallica.bnf.fr/ark:/12148/bpt6k902022.r=
cauchy+oeuvres.langFR.
Cauchy, A.-L. (1837). Extrait d’une lettre à M. Coriolis. In Œuvres Complètes, vol. 4 of series
I, pp. 38–42. Paris: Gauthier-Villars, 1884. http://gallica.bnf.fr/ark:/12148/bpt6k90184z.r=
cauchy+oeuvres.langFR.
148 10 Cauchy: The Bourgeois Revolutionary as Activist of the Restoration

Cauchy, A.-L. (1849). Sur quelques définitions généralement adoptées en Arithmétique et en


Algèbre. In Œuvres Complètes, vol. 14 of series II, pp. 215–226. Paris: Gauthier-Villars, 1938.
http://gallica.bnf.fr/ark:/12148/bpt6k90206f.r=cauchy+oeuvres.langFR.
Itzigsohn, C. (1885). Algebraische Analysis von Augustin Louis Cauchy. Berlin: Springer. http://
gdz.sub.uni-goettingen.de/dms/load/img/?PPN=PPN379794896.
Lakatos, I. (1966). Cauchy and the Continuum: The Significance of Non-standard Analysis for the
History and Philosophy of Mathematics. Cited from Lakatos 1980, vol. 2, pp. 43–60.
Lakatos, I. (1980). Philosophcal Papers (2 vols) vol. 1: Mathematics, science and epistemology.
Cambridge University Press.
Laugwitz, D. (1987). Infinitely small quantities in Cauchy’s textbooks. Historia Mathematica, 14,
258–274.
Lorenzen, P. (1965). Differential und Integral. Eine konstruktive Einführung in die klassische
Analysis. Frankfurt am Main: Akademische Verlagsgesellschaft.
Mayer, J. T. (1818). Vollständiger Lehrbegriff der höhern Analysis. Göttingen: Vandenhoek und
Ruprecht. urn:nbn:de:bvb:12-bsb10082386-8.
Peano, G. (1892). Sur la définition de la dérivée. Mathesis, 2(2), 12–14.
Riede, H. (1994). Die Einführung des Ableitungsbegriffs. Mannheim, Leipzig, Wien, München: BI
Wissenschaftsverlag.
Robinson, A. (1963). Introduction to model theory and to the metamathematics of algebra.
Amsterdam: North-Holland Publishing Company.
Spalt, D. D. (1981). Vom Mythos der Mathematischen Vernunft. Darmstadt: Wissenschaftliche
Buchgesellschaft, 2 1987.
Spalt, D. D. (1996). Die Vernunft im Cauchy-Mythos. Thun und Frankfurt am Main: Harri Deutsch.
Spalt, D. D. (2002). Cauchys Kontinuum: Eine historiografische Annäherung an Cauchys Sum-
mensatz. Archive for history of exact sciences, 56, 285–338.
Spalt, D. D. (2015). Die Analysis im Wandel und im Widerstreit. Freiburg: Karl Alber.
Chapter 11
The Interregnum: Analysis on Swampy
Ground

On the Utility of History of Mathematics

The Special Quality of Our Perspective

The last chapter explained Cauchy’s basic view of analysis. This explication claims
that it would meet Cauchy’s approval in case he were to read it as our contemporary
(whatever this may mean).
Moreover, this explication is nonetheless one by Cauchy himself—or one which
he would have been able to give. This exposition of Cauchy’s basic view of analysis
is an exposition from today, from our actual viewpoint—what does this mean?
The given exegesis of Cauchy’s basic view of analysis is not a translation of
his words into modern language. (For otherwise, we would not have spoken of
“quantities”, as this notion does not exist at all in modern analysis.) Instead, our
aim is that:
1. The focus of the previous chapter was Cauchy’s basic view of analysis within
the context of his own concepts. (Other chapters present the concepts of other
mathematicians.)
2. This was opposed to other basic views of analysis: (a) the “Algebraic Analysis”
of Euler and Lagrange, (b) the actual (standard) analysis, and (c) the actual
Nonstandard-analysis.
3. The last two views were beyond Cauchy’s grasp, for they did not exist in his
times. Consequently, these views are precisely the addition which history of
mathematics can utilize in order to assess the potential of Cauchy’s analysis
retrospectively.
4. These opposing perspectives supplement the theory created by Cauchy. They
show aspects that were not available to Cauchy (as well as to his contempo-
raries). Yet, they do not change Cauchy’s mathematical theory.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 149
D. D. Spalt, A Brief History of Analysis,
https://doi.org/10.1007/978-3-031-00650-0_11
150 11 The Interregnum: Analysis on Swampy Ground

5. These are eminently mathematical aspects. They allow for definite answers to
questions of the following kind: (a) Is the “characteristic function” of the rational
numbers (i.e. f (x) = 1 if x is rational and f (x) = 0 otherwise) in Cauchy’s
view a “function”? (b) What does Cauchy understand by “limit”: a uniquely
determined value or several values? (c) Does Cauchy deal (in modern language)
with usual “convergence”, with “uniform” convergence, or with “continuous”
convergence or with something quite different? (d) Is Cauchy’s notion of
“differentiability” our modern “continuous differentiability”? (e) Is his “Sum
Theorem” true or false? etc.
6. This is the very technique of obtaining a historical as well as philosophical
assessment of the mindset of the mathematician under discussion. This allows for
the building of bridges to general history as well as to the history of philosophy.

What Additional Knowledge Have We Gained That Was Not


Available to Cauchy and to His Contemporaries?

Surely, Cauchy was very aware of the dissenting character of his lectures from
the “Algebraic Analysis” of Euler and Lagrange. He explicitly declared this in
the preface to his first textbook (p. 116). Retrospectively we can say that such a
fundamental transformation remains unique within the history of analysis.
However, from today’s point of view, we can give a more detailed account:
Cauchy presented his very own view of analysis. It differs mathematically very
clearly from all current versions of analysis, especially from standard-analysis and
Nonstandard-analysis.

Crucially, we have to add:


Such an understanding was principally impossible for Cauchy and for his
contemporaries—for the conceptual framework of these versions did not exist at their
time.

In the year of publication of Cauchy’s first textbook, 1821, it was ABSOLUTELY


NOT IMAGINABLE that there may exist DIFFERENT (and consequently, competing)
mathematical theories about ONE topic. The most prolific mathematicians of that
time were just engaged in creating a new geometry, a fact that was going to shock
the scientific community. Thus, it still took many decades until “non-Euclidean”
geometry was accepted as a decent mathematical theory (besides the earlier one
which only then was called “Euclidean”)—as a different but equally exact and
mathematically legitimate theory.
However, geometry is no exception and the same can be said about analysis.
Only the twentieth century brought this awareness of the essential ways in which
mathematics develops and changes, and even today this knowledge is still very far
from being widespread.
The Teaching of Analysis Without a Curriculum 151

The Teaching of Analysis Without a Curriculum

The higher schools had already started to teach analysis in the eighteenth century.
In Germany, the successful textbooks of Abraham Gotthelf Kästner (1719–1800),
written also for private studies, illustrate this. In the nineteenth century, this tendency
increased enormously. The beginning of the era of industrialization required ever
more mathematical skills for ever more people.
Which kind of analysis was taught at the universities?
Seen from today’s point of view, one could ask more precisely: did they teach
“Algebraic Analysis” or the “Analysis of Values”?
However, this detailed question does not make sense—at least if one were
expecting a considered, meaningful answer from the point of view of those past
teachers. As we have just stated: save Cauchy, surely no mathematician of the first
half of the nineteenth century was aware that different theories of analysis were
possible. Of course, Cauchy knew only one rival theory: the “Algebraic Analysis”
of Euler and Lagrange. (This is, strictly speaking, still true today.)
Instead, analysis then was simply analysis (and as a rule, so it remains today): a
theory thought (or supposed) to be undivided. It was this doctrine which was taught
as “calculus” to the students.
For example, in my experience and to my knowledge, it had not been realized
until 1991 that the very meaning of the concept of “convergence” depends on the
precise meaning of the concept of the “value of a function” (see pp. 135f).
For a very long time, the concepts of analysis—like all mathematical notions—
were taken to be clear-cut and unchangeable. As objects of a “Platonic world”, they
seemed to be eternal.
The insight that this is false—that the non-existence of an eternal code of
mathematics, for mathematics is created by a specific human culture, like all the
other sciences—is a new one. This is far from common knowledge. (Although, even
today nobody is really able to conjecture where this super code of mathematics could
be written down.)
In case you know only one version of analysis, there is no basic problem with
the curriculum. Then the only maxim is teach the analysis, particularly those parts
which are important for your students!
This is what was usually done. Each university or college developed a curriculum
of its own, depending on the knowledge of the local staff. There was no awareness
of the necessity of synchronizing teaching in different places (in order to compare
students’ ability or, what amounts to the same, to keep the unity of mathematics).
In the early nineteenth century, nobody realized that, as a consequence of this missing
coordination of teaching, the unity of analysis (to be understood as a system without
contradictions) was at stake.

Actually, not one of the scientists of that time was able to realize this—this is
an objective statement. The standard of knowledge had not yet reached the level to
152 11 The Interregnum: Analysis on Swampy Ground

engender such a possibility. Only today have we arrived at this insight. Only today
we know:
If the basic concepts of a mathematical theory are not stated with absolute clarity, the
subsequent mathematical propositions become somewhat mathematically imprecise.
Here, mathematically imprecise entails nothing other than that the relevant propositions
might produce a contradiction, be it in respect to each other or be it in respect of the
propositions of other authors.
The “contradictions” under consideration are those which could, in principle, be detected
by the mathematicians of that time. They are not “contradictions” which could only be
detected with the benefit of hindsight.

We have already mentioned two such contradictions, namely controversial


theorems: the “Fundamental Theorem of Functions” (p. 128) and “Cauchy’s Sum
Theorem” (p. 138).

Analysis as Freestyle Wrestling

Indeed, analysis developed in the nineteenth century in the manner of freestyle


wrestling or in today’s world, like mixed martial arts. This will be shown in the
current chapter with reference to the following three points:
A. Missing conceptual precision
B. Political rather than rational argumentation
C. Parallel methods which differ in precision

A. Missing Conceptual Precision—or—Riemann Invented the


Modern Concept of Function

Already in the preface to Cauchy’s first textbook of analysis—as we propose to call


it: “Analysis of Values”—we can read: “In determining the values of the quantities
I will make all uncertainty disappear.” (p. 116).
As “function” has been the central concept of analysis since its invention by
Euler, the very first consequence of Cauchy’s dictum is this: determine the concept
“value of the function” precisely!
Cauchy did so—we discussed this at length (see pp. 124f). (Some may have
thought: much too long!)
And what about Cauchy’s colleagues and successors?
The astonishing answer is: for a whole generation, or fully 30 years, nothing
happened. This nothing has to be taken seriously:
Neither was Cauchy’s definition of “value of the function” accepted by his colleagues or
successors, nor was it opposed by a competing one.
Analysis as Freestyle Wrestling 153

The different positions are briefly given as examples below:


1. Dirichlet in 1837 Carl Gustav Jacob Jacobi (1804–51) sings Dirichlet’s praises.
In a letter to Alexander Humboldt, Jacobi writes gushingly:
Only he, not me, not Cauchy, not Gauß, knows what a completely strict mathematical proof
is. If Gauß says he has proved something, I take it as likely; if Cauchy says so, there are
equally many pros and cons to bet; if Dirichlet says it, it is sure.

Let’s see!
Dirichlet is generally taken to be the inventor of the modern concept of a function.
This is incorrect. I do not know one single place where Dirichlet defined the concept
of “value of the function”.
Even worse, according to Dirichlet, the “value of the function” is not necessarily
uniquely determined—as is demanded in all actual textbooks of analysis. On the
contrary, Dirichlet invented a notation, which is still used today, for functions that
have two different limits on either side, depending on whether the left-hand side or
the right-hand side is taken (for Dirichlet, “−0” or “+0”). He writes the following:
If the function f (x) undergoes a sudden change for singular values of x without becoming
infinite there, then this curve consists of several interrupted parts. For every such abscissa,
we detect two ordinates, actually, one of them belonging to the ending and the other one to
the starting piece of the curve.
In the following, it will be useful to distinguish those two values of the function and
denote them by f (x − 0) and f (x + 0).

That is to say, if the function has a jump for the value x, Dirichlet distinguishes
explicitly the TWO “values of the function” to be f (x + 0) and f (x − 0).
Not only this! Some pages later, Dirichlet treats an incidence of just the same
kind—but his calculations lead to a third “value of the function”:
When a disruption of the continuity occurs and, consequently, the function f (x) has two
values, the [trigonometric] series [of this function], which naturally has only one value for
each x, gives half of the sum of these values.

That is, the function has at this point two “values” (in some sense determined
and thereby following Cauchy) f (x + 0) and f (x − 0)—as well as a third “value”
which is algebraically determined from the trigonometric series to be the mean of
the other two.
A completely strict mathematician is expected at this point to declare which one
(or more) of these “values” is (are) to be taken as the “value(s) of the function”.
However, Dirichlet keeps completely silent on this. This is not strict at all!
Interestingly, there is one formulation by Dirichlet from which we can understand
that he, as well as Cauchy before him, took the “Fundamental Theorem of
Functions” to be self-evidently true.
We have just read that Dirichlet did not accept Cauchy’s concept of “value
of a function” which is (not strictly logical, but based on the facts) a necessary
assumption of Cauchy’s Fundamental Theorem of Functions. In other words, if
the following interpretation were true, we would have to conclude that Dirichlet’s
thinking regarding the basic notions of analysis was self-contradictory.
154 11 The Interregnum: Analysis on Swampy Ground

By virtue of lecture notes from the year 1854, we know that Dirichlet defined
“continuity” as follows:
y = f (x) is called a continuous and definite or one-valued function of x, if each value of
x only belongs to one value of y and if each gradual change of x corresponds to a gradual
change of y, i.e. if for some fixed x, the difference

f (x + h) − f (x)

for steadily decreasing h converges on the limit zero.

I confine myself to the following simple interpretation, which goes as follows:


(i) In his first sentence, Dirichlet gives one single definition. (ii) In this definition, Diri-
chlet uses two names: “continuous” and “definite or one-valued”. (iii) Consequently, Di-
richlet took both names to mean the same. (iv) The meaning of “definite or one-valued”
is self-explanatory. (v) To sum up, in this definition, Dirichlet supposes the validity of the
“Fundamental Theorem of Functions”—end of the simple interpretation.

The rules of higher rhetoric allow for a contradictory interpretation of Dirichlet’s


passage. We shall leave the details to the specialists of rhetoric and plainly state its
result: it claims that Dirichlet is giving two definitions in his first single sentence.
However, the question arises: is it legitimate that Dirichlet opens his lecture
by giving such an artificial and pretentious formulation, wherein he defines two
different concepts in one single sentence? In one sentence, which allows a simple
understanding with a fundamentally different meaning of the two given notions? For
indeed, this sentence is the very first of Dirichlet’s lecture notes.
I believe this to be unrealistic and therefore keep to my straightforward interpre-
tation.
This simple interpretation fits well to the following formulation from (p. 131) of
the lecture notes (emphases original): this function is
thoroughly continuous, i.e. everywhere definite and specified, finite and changes gradually.

To Dirichlet as to Cauchy, the “and changes gradually” is a consequence of the


“definite and specified”, whereby “definite” implies at least “finite”.
2. Riemann in 1851 Dirichlet’s pupil Bernhard Riemann conceived things differ-
ently. The first half of the first sentence of his doctoral thesis, published in 1851,
goes as follows:
If one thinks z to be a changing quantity, which gradually can take all possible real values
and if to each of its values there corresponds one singular value of an indefinite quantity w,
then w is called a function of x . . .

Herein Riemann states point-blank: all you need in order to have a “(real) func-
tion” is, that to each “value” exists a “corresponding” single “value”—neither more
nor less. How this “correspondence” is accomplished (be it through a “calculatory
expression” as with Euler or a “dependence” as with Cauchy or otherwise) is left
completely open by Riemann. He says no more than “(definite) correspondence”.
That sounds as if Riemann simply drew the consequences from Cauchy’s defini-
tion of the “derivative”—which, ironically, does not fit at all Cauchy’s own concept
Analysis as Freestyle Wrestling 155

of function (see p. 144). Riemann was going to change all this, because by now
the “Calculus [of differentials and integrals]” required the function “derivative”.
Consequently, the concept of function needed urgently to be adjusted to fit new
developments.
Riemann coins the concept of “function” which is still valid today: to each “value”,
x uniquely corresponds a “value of the function” f (x)—nothing more, plainly this.
Nothing is assumed about the manner or the realization of this correspondence.

Thus the matter is stated. Obviously, Riemann’s style differs from ours. It is
his style to talk about “quantity” when we, today, mostly speak about “sets” and
“elements”, maybe even about “ordered sets”. However, these are differences of
expression, not of mathematics. The mathematical objects “function” and “value of
the function” of Riemann and of ours are the same.
Clearly, concerning philosophical aspects, this is put somewhat roughly: style
and content cannot be separated completely. However, this is not the place to
differentiate that precisely; our formulation shall suffice here.
As one can easily imagine, in spite of Riemann’s exceptional mathematical
genius, the publication of his doctoral thesis was not an instant success. It did not
lead to the immediate acceptance of Riemann’s new view in 1851. Initially, many
mathematicians did not follow Riemann, among them his illustrious antagonist
Weierstraß—at least almost until the very end of his career. Certainly, we shall return
to this.
In the long run, it turned out that Riemann was on the right track, especially when
he opted for the most general concepts possible for both a function and a value of the
function. Bolzano had already anticipated those (p. 111), but as this revolutionary
thinker had been relegated from university by the conservative government and
banned from publishing, his idea remained unpublished in the nineteenth century.
Riemann can therefore take all the credit for having invented the two most central
concepts of actual mathematics and for their introduction into the mathematical
discourse.
3. Björling in 1864 Anybody who did not live in Paris in the midst of the nineteenth
century and who did not have an excellent teacher but really wanted to study analysis
sooner or later realized that he or she had to study Cauchy’s textbooks. All those who
only had Cauchy’s books at their disposal might have been well motivated to read
them very carefully.
Such a studious student was Emanuel G. Björling (1808–72) from Sweden. The
relevant lexicons do not know him. A footnote of one of his treatises, published in
1854, reads as follows:
We should record that it is said for a purpose and not because of some laxity that a function
of x which changes continuously in between its two limits is only allowed to have a single,
unique, finite value for each limit of x.

A careful inspection shows that this is exactly what the “Fundamental Theorem
of Functions” says! (For Björling demands in no uncertain terms: if the value
156 11 The Interregnum: Analysis on Swampy Ground

of the function for some “limit”—or a “value”—is unique, then the function is
“continuous” at this value.)
I have been unable to detect any other explicit statement by authors other than
Björling and Cauchy, which clearly embraces the content of the “Fundamental
Theorem of Functions” (and which therefore presumably assumes Cauchy’s concept
of the value of a function). The possible exception of Dirichlet has already been
discussed.

B. Political Instead of Rational Reasoning

As already mentioned on pp. 133f, Cauchy’s (only today so-called) “Sum Theorem”,
stated in his textbook on analysis, was repeatedly criticized. The most popular
criticisms are those by Niels Henrik Abel from 1826 and by Philipp Ludwig Seidel
from 1850. How do these criticisms run?
The answers to this question are as sobering as they are interesting:
Neither Abel nor Seidel criticize Cauchy’s proof.
Instead, they just state their different opinion and try to justify it.
Both critics do not criticize but proclaim yet another point of view on the
subject.
Clearly, this is a political way of reasoning, not a mathematical one.
Just as in our everyday experience, political debates run as follows: the specific
views are presented and supported by some arguments, and all opposing views are
condemned—just because they contradict the speaker’s own views, but not because
they are inconsistent in themselves or unconvincing (or even illogical). This seems
to be the ubiquitous practice of political debates today.
Abel and Seidel conceived analysis just in contrast to Cauchy. That is not
objectionable. At the same time, this does not justify the judgement that Cauchy’s
theorem is wrong. The only thing Abel and Seidel are justified in saying is: “We see
things differently from Cauchy!” This would be totally legitimate, and if they were
to produce their arguments, it could even be constructive.
1. Abel did not give any argument. He simply claimed Cauchy’s theorem to
“suffer from exceptions” (a polite formulation towards the authority of Cauchy).
He even produced an example of such an alleged “exception”. But he proved neither
(a) that and why Cauchy’s argumentation should be wrong, nor (b) that and why his
example should contradict Cauchy’s theorem. Abel argued by brute force: “I differ
from Cauchy and in this way!” No argument whatsoever, neither against Cauchy
nor in favour of the correctness of his own claim.
The mathematical value of this Abelian claim is obviously nil.
Thus, it is very strange that the widespread assessment of Abel’s criticism of
Cauchy was completely different. Practically all mathematicians (and historians of
Analysis as Freestyle Wrestling 157

mathematics) of today praise Abel to be a perspicacious critic of Cauchy. On what


grounds?
A possible answer might be this: the proponents of this assessment did not
familiarize themselves with Cauchy’s way of thinking or with that of Abel (which
in this case is less important). (What would have to be done in this case is explained
on pp. 140f.) They do not rate the mathematical stringency of the given arguments
by Cauchy and Abel. Instead, they simply ask “who is right?” Just as if there were
only one single CORRECT construction of analysis. As stated before, this is a purely
political (and one-sided) judgement.
However, this stage of mathematical development has been overcome for at least
50 years. (If we include the intuitionists, even for a century.) Nevertheless, the
majority of mathematicians (and worse, of the historians of mathematics, too) are
not willing to take notice of that. Why not?
2. Seidel, a generation later, was much more constructive than Abel. Indeed,
Seidel, too, did not try to detect Cauchy’s reasoning error. (As we know, he would
not have succeeded as such an error does not exist.) Still, Seidel at least justified his
view—thereby showing his dissenting conception from Cauchy’s analysis in strict
and clear mathematical terms.
This does not mean that Seidel specified this dissent—not at all! Instead, Seidel
produced clear arguments for his own view, and this we are going to examine here.
(It is the job of the historian to compare this way of argumentation with Cauchy’s.) It
was Seidel’s achievement to coin the concept which is called “uniform convergence”
today. (To be precise, Seidel came up with the opposite of the concept of “uniform
convergence” and called it “arbitrarily slow convergence”. But clearly, the definition
of a concept also defines its opposite.)
Seidel offered a precise definition of his new concept and with this notion he
formulated a theorem that logically contradicts Cauchy’s Sum Theorem—if Cau-
chy’s concepts (and especially his concept of “convergence” and, consequently, his
theorem) are taken in a way which was not intended by Cauchy, that is to say: as
understood and taken by Seidel
For those who want to know more: Seidel formulated and proved the following
theorem:
Theorem. If you have a converging series which represents a discontinuous function and if
the single terms of the series are continuous functions, than it must be possible to specify
values in the immediate neighbourhood of the place at which the function jumps, for which
the series converges arbitrarily slowly.

Seidel’s theorem says: if a “converging” series of continuous functions represents


a function which is discontinuous at the value x = x0 , then the convergence of the
series is not the usual one but one which is only “arbitrarily slow”.
Let us consider it the other way round: if we only demand the usual convergence
of the series at the value x = x0 (as was done by Cauchy, in Seidel’s understanding,
which as we already know is erroneous), then the continuity of the function at the
value x = x0 may not be guaranteed.
158 11 The Interregnum: Analysis on Swampy Ground

Seidel has a complicated way of constructing his analysis. The same holds for
today’s standard-analysis.
One more final word about the invention of the concept of “uniform” convergence
for this notion is somewhat important in modern analysis.
As stated above, Seidel invented this concept in 1850 (more precisely, its
opposite). However, he was not the only one to do this. Abel published a treatise
where he used the name “steadily convergent” (“beständig konvergent”) already in
1829, and in a letter published in 1830, Abel explains his understanding of this
notion. Christoph Gudermann (1798–1852) presumably took this concept and later
taught it to his gifted student Karl Weierstraß. It was Weierstraß who then made
this concept well-known (under the name “uniform” convergence, “gleichmäßige”
Konvergenz)—we shall return to this later.

Parallel Methods with Different Levels of Precision

It is central for Cauchy to use the concept of “limit”, written “lim”. This notion
is the technical realization of the programme presented in the preface to his first
textbook: the precise determination of the relevant “values” of the “variable”. The
“limit” mediates between the “variable” and its “values”.

“For x = 0, we have lim rn (x) = 0”

means the “variable” rn (x) (with the two variables n and x) has for the “value”
x = 0 and n → ∞ the unique “value” 0.
Unlike us today, Cauchy uses the operator “lim” always without subscripts:
“ lim ”. This can be explained by the fact that Cauchy always means all possible
...→...
limits—at least in the context of his accompanying text. Thus, Cauchy has all limits
to make up the “value(s) of the function” for the “value” x = X, and therefore any
specification of the limits is completely superfluous. With the above,

“ lim rn (x) = 0 for x = X” Cauchy means our lim rn (xk ) = 0 ” .


“ n→∞
xk →X

All the limits are taken, n → ∞ as well as xk → X and all their inter-dependencies
whatsoever.
Of course, if one uses another concept of “lim”—as, e.g., Riemann does—one
must specify exactly which limits are meant. In this case, one must use subscripts
with “lim”. We today have learnt analysis in another way from that developed
by Cauchy. That is why we automatically tend to read Cauchy’s formula in the
following way:

“ lim rn (x) = 0 for x = X” means “ lim rn (X) = 0” .


n→∞
Analysis as Freestyle Wrestling 159

We take only one limit: n → ∞; the other one: xk → X is omitted, and also the
possible dependencies between them. All this because today we replace Cauchy’s
concept of “value of the function” by that of Riemann. Cauchy understood “ lim”
differently from us today.
This leads to the modern misunderstanding of Cauchy, and this misunderstanding
transforms (in modern language) Cauchy’s concept of “continuous convergence” to
the simple “convergence”.
Not every mathematician since Cauchy has used the concept of “limit”. For
instance, Seidel’s treatise from 1850, mentioned above, contains many considera-
tions of limits, but the operator “ lim” is never used. For example, his comments on
the formula ∗∗ on p 139 go as follows:

sn (x + α) − sx (x) < ε1
rn (x) < ε2
rn (x + α) < ε3

where ε1 , ε2 , ε3 designate arbitrarily small absolute quantities and all the inequalities are
taken independently of the positive or negative sign.

ε1 , ε2 , ε3 are meant to designate (small) numbers. (Actually, Seidel used other


Greek letters, but this does not matter mathematically. The use of epsilons facilitates
our understanding as we are used to them.) By the way, Seidel understands by “x”
a “value”, not a “variable”.
Today the method used by Seidel is called “epsilontics”, for today the Greek
letter epsilon usually denotes certain very small tolerances. If it is possible to
show—let us stick to the above example—that we have

s(x + α) − s(x) < ε1 + ε2 + ε3

as long as we have α < δ, it is proved that we actually have

lim rn (x) = 0 .

(Of course, the problem is always somehow to specify the security margin δ.)
Then it is proved: rn (x) can be made smaller than any arbitrarily small given number
ε1 + ε2 + ε3 (independent of the sign). Or the other way round: it is impossible that
lim |rn (x)| = h with a number h > 0—consequently, only lim rn (x) = 0 is left.
The usage of the operator lim is often called “limes calculation”. The name is
somewhat inappropriate, for we do not calculate with the “ lim”, but we calculate
the limit. However, we cannot change the common terminology.
Using both names, we can summarize:
Analysis is possible with epsilontics as well as with the help of limes calculation.
160 11 The Interregnum: Analysis on Swampy Ground

The difference between the two methods is that:


The limes calculation needs a variable “quantity”, epsilontics does not.

That is why epsilontics is more general than limes calculation.


The general implementation of limes calculation is due to Cauchy. He used the
operator lim as a central tool in realizing his programme, for this operator mediates
between the “variable” (which is a “quantity”) and its “values”.
In contrast to Cauchy’s analysis, modern theory lacks the notion of “quantity”
and certainly does not have it as a foundational concept. Therefore, if modern
analysis operates with limits, another understanding of the operator lim than that of
Cauchy is required. One expression of this other understanding is today’s demand
clearly to specify the limes in the subscript “ lim ”.
...→...

The Foundational Uncertainty of Limes Calculation:


A Prominent Misunderstanding by Prominent Scholars

Today’s mathematicians are well versed in epsilontics. This may lead to a disregard
of earlier limes calculations. For example, the distinguished mathematician of the
later twentieth century, Detlef Laugwitz, claimed, in all seriousness, that Riemann’s
and Cauchy’s formulations of the basic definition of convergence were wrong:
Even more surprisingly, one finds [in Riemann] obviously wrong formulations as in the case
of the convergence criterion of a sequence sn (x) [. . . ] Almost literally, the same misleading
formulation can repeatedly be found in Cauchy since his Cours d’analyse of 1821.

This judgement, which is printed in a specialized textbook, is far from modest.


However, is it well-founded?
A closer reading shows: no, this judgement is not substantiated. It is based on
a misreading. This will briefly be explained, for it shows the fundamental problem
with modern limes calculation.
At first, one easily realizes that this formulation which Laugwitz judges to be
“obviously wrong” is to be found not only in Cauchy and Riemann but also in
Georg Cantor and even in Karl Weierstraß, who is commonly judged to be extremely
meticulous. This must create some doubt. So let us take a closer look.
Laugwitz criticizes some notes, taken from Riemann’s lectures by one of Rie-
mann’s students. They state (emphases are original):
An infinite series, the terms u0 , u1 , . . . of which follow a law of formation, is convergent if
the sum


m=n
sn = um
m=1

for increasing n approaches a fixed limit which is called the value of the series. Therefore,
the general condition for convergence is that

sn+k − sn ,
The Foundational Uncertainty of Limes Calculation: A Prominent. . . 161

for arbitrary k and increasing n, decreases without end, or in other words that for every k


n+k
lim um = 0 .
n=∞
n+1

Weierstraß writes:
In case of such an unconditionally summable series, one easily proves that

lim |sn+k − sn | = 0
n=∞

for each k, i.e. the change of sn becomes arbitrarily small for increasing n.

The problematic formulation is the condition for the segment length k. Riemann
says “for arbitrary k”, while Weierstraß writes “for each k”. So what constitutes the
problem?
The problem Laugwitz touched upon is best understood through the example of
the “harmonic series”:

1+ 1
2 + 1
3 + 1
4 + 1
5 + ...

On p. 85, we have already studied it in detail and have shown:


However far we proceed in the series, there always exists a segment of
length k, where the sum of its terms exceeds 12 :

1
n+1 + 1
n+2 + ... + 1
2n >n· 1
2n = 1
2 .

Therefore, it does not converge. However, if you specify the length k of the
segment already before you decide on the length N of the section, you get for the
“harmonic series” the following:
For each tolerance ε and for any length k for a segment, there exists a length N such
that for any section of length greater N each partial sum of a segment of a length which
is not larger than k is less than ε. Expressed algebraically,
1
n+1 + 1
n+2 + ... + 1
n+k < ε,

if only n > N .

That is to say, at first one decides on the value of k (and that of ε) and only
afterwards does one determine the value of N. It is N = N(k) (or, more exactly,
N = N (k, ε)).
One can easily be convinced of it. The condition sounds like a generalization
of the necessary condition for the (single) terms of a converging series to become
162 11 The Interregnum: Analysis on Swampy Ground

arbitrarily small. In epsilontic formulation, for each tolerance ε, there exists a


section of length N such that

1
n <ε if n > N .

That is true, because one only has to choose N > 1ε (which is always possible; why
actually?); for then we have N1 < ε, and from n > N, we get n1 < N1 —all is done:
1 1
n < N < ε.
Now, one term is just a sum with only a single summand. Let us now take a sum
with two summands:

1
n + 1
n+1 .

With the help of the larger of them (which is the first one), we easily get an upper
estimate:

1
n + 1
n+1 < 2
n .

It is the same with a sum consisting of k summands:

1
n + 1
n+1 + 1
n+2 + ... + 1
n+(k−1) < k
n .

From n + k > n, it follows 1


n+k < 1
n and consequently

1
n+1 + 1
n+2 + 1
n+3 + ... + 1
n+k < k
n .

And now the trick used in the first step is: for a given tolerance ε, we choose N > kε
and consequently N1 < kε or Nk < ε. Then from n > N, it follows n1 < N1 as well as
k k
n < N and finally

1
n+1 + 1
n+2 + 1
n+3 + ... + 1
n+k < k
n < k
N < ε.

That was to be proved. The decisive point is that at first one chooses the length k of
the segment and the tolerance ε—only afterwards the section length N .
We have learnt two things: (i) both propositions in the boxes are valid for the
“harmonic series”. Against our initial impression, they do not contradict each other.
(ii) The proposition in the second box does not guarantee the convergence of the
series!
After finishing this excursion into epsilontics, we return to Riemann, Weierstraß,
and the others and go back to limes calculation!
The last two boxes show: for the characterization of convergence the segment
length k is of importance. For we have just shown that the “harmonic series”
complies with the condition of the last box—nevertheless it does not converge, as
The Foundational Uncertainty of Limes Calculation: A Prominent. . . 163

the box before it demonstrates. The condition in the last box is too weak and does
not guarantee convergence.
Now the difficult part, the interpretation of the texts. A possible condition for
convergence given by limits is:

lim (sn+k − sn ) = 0 for each k.

The segment of the series under examination in case of the “harmonic series” is

sn+k − sn = 1
n+1 + 1
n+2 + ... + 1
n+k .

Now, what about k?


Riemann says (p. 160) that the equation lim (sn+k − sn ) = 0 is valid “for
every k”.
Laugwitz claims: in saying this, Riemann fails to notice the weakness of the
condition in the last box. In more detail, Laugwitz says: (i) in his definition of
convergence, Riemann demands only the condition of the last box. (ii) However, the
“harmonic series” fulfils just this condition (this we just proved). (iii) Nevertheless,
the “harmonic series” does not converge (this is shown the box before that).
(iv) Consequently, Riemann’s definition of “convergence” is wrong.
Anyhow, this is only an assertion of Laugwitz. It is indeed possible to understand
Riemann this way: he demands that after having fixed the section length N, the
partial sums of every segment length k stay less than the tolerance ε. That is,
For each tolerance ε, there exists a section length N such that from this section length
N onwards, every partial sum of segment length k is less than ε. In formulae,
1
n+1 + 1
n+2 + ... + 1
n+k < ε,

if n > N AND FOR EVERY k.

This is different from what the last box states, because N is now determined
independently from k! N is not an N(k).
If the clause “for every k” is taken in this way, i.e. if the tolerance ε is de
facto prescribed independently of the length k of a segment, then the penultimate
box shows that the “harmonic series” violates this condition. Consequently, the
“harmonic series” is not a counter-example for the condition of convergence when
understood in this way.
To sum up: if Riemann is read and understood in the way which is explained
above, then his definition of “convergence” is correct. And the very same holds for
Weierstraß’ formulation.
The problem with the limes calculation is this: a formulation like “it is demanded
that lim (|sn+k − sn | for all k) = 0 ” is impossible for purely syntactic reasons:
n→∞
an equation does not allow for text. Subsequently to deduce from this that “Rie-
mann and Cauchy are unable to define the concept of ‘convergence’ correctly!” is
inadmissible because it is not properly justified.
164 11 The Interregnum: Analysis on Swampy Ground

Let me conclude in the following way:


1. In the first half of the nineteenth century, both methods, limes calculation and
epsilontics, were practised side by side.
2. With some difficult issues, it is hardly possible to offer an unequivocal formu-
lation in limes notation. For example, Laugwitz misunderstood Cauchy’s and
Riemann’s characterizations of convergence. (With this late historiographical
text, Laugwitz did not live up to his excellent mathematical renown.)
3. This problem simply disappears if the limit is taken as a matter of principle for
all variables—as it was practised by Cauchy.

What is “x ” After Cauchy?

Before Riemann’s clear definition of the value of a function to be uniquely determined,


there was an anarchy of significations. “x” could designate everything: a “variable” as
well as a “value”.

We observed this already in Bolzano: even he used “x” to designate a “variable”


as well as any of its “values” (p. 108).
Only Cauchy made things clear and distinguished even in notation scrupulously
between “variables” and their “values” (p. 147). Remarkably, this Cauchyan manner
of notation was not taken up by anyone, as far as I know.
Even Riemann’s decision to demand the value of the function to be uniquely
determined did not end this anarchy of significations.
Even after Riemann had demanded that the “value of the function” must be uniquely
determined, the notation “x” could designate a “variable” as well as a precise “value”
of it.
In just the same way, “f (x)” could be chosen as the name of a function (i.e. a
“variable”) as well as for a “value of the function” (i.e. the “value” of the function
at the “value” x).

Cauchy’s code of notation would have blessed analysis with greater clarity, even
after Riemann.

Literature

Arendt, G. (1904). Gustav Lejeune Dirichlets Vorlesungen über die Lehre von den einfachen und
mehrfachen bestimmten Integralen. Braunschweig: Vieweg. reprint VDM-Verlag Dr. Müller,
Saarbrücken, 2006.
Björling, E. G. (1854). Om det Cauchy’ska kriteriet på de fall, då functioner af en variabel låta
utveckla sig i serie, fortgående efter de stigande digniteterna af variabeln. Kongl. Vetenskaps-
Akademiens Handlingar, För År 1852, 165–228.
Literature 165

Dirichlet, G. L. (1837). Ueber die Darstellung ganz willkührlicher Funktionen durch Sinus- und
Cosinusreihen. Repertorium der Physik, I., 152–174. see also: Kronecker & Fuchs (Eds.) 1889,
1897, vol. 1, pp. 133–160.
Kronecker, L. (vol. 1) & Fuchs, L. (vol. 2) (Eds.) (1889, 1897). G. Lejeune Dirichlet’s Werke.
Berlin: Georg Reimer.
Laugwitz, D. (1996). Bernhard Riemann 1826–1866. Wendepunkte in der Auffassung der Mathe-
matik. Basel: Birkhäuser Verlag.
Detlef Laugwitz 2008. Bernhard Riemann 1826–1866. Turning points in the conception of
mathematics. Boston: Birkhäuser Verlag. English: Abe Shenitzer.
N. N. (1886). Ausgewählte Kapitel aus der Funktionentheorie, Vorlesung von Prof. Dr. Karl Weier-
straß, Sommersemester 1886. In Siegmund-Schultze, R. (Ed.) Teubner-Archiv zur Mathematik
(Vol. 9) Leipzig: Teubner, (1988).
Neuenschwander, E. (1987). Riemanns Vorlesungen zur Funktionentheorie, allgemeiner Teil.
Mathematisches Preprint No 1086, Technische Hochschule Darmstadt.
Neuenschwander, E. (1996). Riemanns Einführung in die Funktionentheorie.. In Abhandlungen
der Wissenschaften in Göttingen, Mathematisch-physikalische Klasse (Vol. 44) third series.
Göttingen: Vandenhoeck & Ruprecht.
Pieper, H. (1987). Briefwechsel zwischen Alexander von Humboldt und C. G. Jacob Jacobi. Berlin
(east): Akademie-Verlag.
Riemann, B. (1851). Grundlagen für eine allgemeine Theorie der Functionen einer veränderlichen
complexen Grösse. in: Heinrich Weber unter Mitwirkung von Richard Dedekind (Ed.).
Bernhard Riemann, Gesammelte mathematische Werke. reprint of the second edition from 1892
(pp. 3–48). New York: Dover Publications (1953).
Seidel, P. L. (1850). Note über eine Eigenschaft der Reihen, welche discontinuirliche Func-
tionen darstellen. In Abhandlungen der mathematisch-physikalischen Classe der königlichen
bayerischen Akademie der Wissenschaften, Jahre 1847–49 (Vol. 5, pp. 379–393). München:
Weiß’sche Druckerei.
Chapter 12
Weierstraß: The Last Effort Towards a
Substancial Analysis

A Famous Man

Karl Theodor Wilhelm Weierstraß (1815–97) invented analysis as a precise mathe-


matical doctrine. So, at least, goes the story told to every German student of analysis
nowadays. According to that myth it was Weierstraß who was the first to establish
clear concepts in analysis and who first carried out strict proofs with the help of
“epsilontics” developed by him. Only in this way could the vague ideas of his
predecessors be expressed clearly.
As with any commonplace, this is nonsense. To be sure, Weierstraß left a lasting
legacy in analysis, but he did not achieve the marvels ascribed to him. As the
foregoing chapters have shown, he could not have achieved this. All basic notions
of analysis—function, differential, integral, continuity, convergence—had already
been known for a long time while little Weierstraß learned to add. All, apart from a
single notion: the concept of number. Here, indeed, Weierstraß produced a marvel:
he invented a concept which none of his students was able to grasp and, in addition,
not one mathematician after him has been able to reinvent, since about 1870!—this
gradually became clear after a lecture note from winter semester 1880/81 emerged
unexpectedly in the mathematics library of Frankfurt university in summer 2016
(see pp. 172f).
Having set out to study the development of the basic concepts of analysis, we
shall refrain from celebrating Weierstraß’ successes; instead, we shall concentrate
on his contributions to the basic concepts of analysis. However, regarding the
concept of number, both happened to coincide.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 167
D. D. Spalt, A Brief History of Analysis,
https://doi.org/10.1007/978-3-031-00650-0_12
168 12 Weierstraß: The Last Effort Towards a Substancial Analysis

The Core of His Fame

It stands to reason that we shall not withhold a hint on the facts which constitute
Weierstraß’ renown as the “founder of rigorous analysis”. It is based on Weier-
straß’ success in smashing the image of analysis of being—besides geometry—
the second mathematical field which draws, by using clear concepts, convincing
proofs from simple images. It was Weierstraß who (besides Bolzano, but this was
unknown in the nineteenth century) designed analysis in such a way that it became a
conceptual minefield, where each combination of notions produces surprises which
render further investigations ever more difficult.
Earlier, analysts had already created grotesque structures. However, they were
fenced in as extraordinary phenomena which were of no importance or could be
controlled. In contrast to that, Weierstraß reversed things. He made these oddities
the touchstone for the quality and the worth of the concepts instead of marginalizing
them as meaningless anomalies.

Dirichlet in 1829: Analysis Has Frontiers!

In 1829, Dirichlet published a crazy idea: a “function” f (x) which for each rational
value of x has the “value” c and for each irrational value of x the “value” d, which
is different from c.
There was not one concept of “function”, including Dirichlet’s own (p. 153),
which was suitable for this creation. Although it has only two different “values”
(in Dirichlet’s thinking: equivalent to “ordinates”), namely c and d, this monster
cannot be drawn. It cannot be drawn, because the rational numbers and the irrational
numbers are intertwined with each other in such a way that they cannot be kept apart
in a drawing, but only as concepts.
Each irrational number can be written as a decimal number with infinitely many
digits which do not recur. If you stop this decimal number, say X, at any digit, you
get a rational number X . However, the place of the digit may be pushed as far out
as one likes. The difference between X and X is of magnitude 10−k , if the number
is cut off at the k-th decimal place. In other words: the difference between X and X
can be made as small as one wishes. In each case we have:
An irrational number has no nearest rational number.—The reverse is also true.

In the case of Dirichlet’s so-called “function” this means: near each “value of the
function” c there is a “value of the function” d—without the possibility of detecting
the length of this “near”. Consequently, it is futile to ask for the “jumps” of this
“function” between its values c and d. Indeed, there is a distance between c and
d (as they are assumed to be different), but the smallest difference of such values
X and X̄, which belong to c and d respectively, cannot be given. Or is it 0? So no
difference?
The Core of His Fame 169

It is true: Dirichlet published this so-called “function”, but he did not really care
about it. He only stated it. Why? Dirichlet wanted to say: there are cases in which
you cannot “integrate”—analysis has limits. Without integral, no calculus. Dirichlet
gave this “function” to demonstrate the limits of the theory of analysis. This strange
object was not worth examining for him. And this is no wonder, because actually it
is—according to his own concepts—no “function” at all.

Riemann 1854: These Limits of Analysis Can Be Shifted!

Dirichlet’s student Bernhard Riemann was less humble. His idea was to fence in
such extraordinary phenomena as proposed by his teacher. Riemann was not willing
to delay the development of analysis. He tried to show that further progress was
indeed possible and that it would not cause insurmountable problems.
In his habilitation thesis, written in 1854, Riemann constructed a function (and
according to his concepts, see p. 155, it actually is a “function”!) which was not
as unruly as that of his teacher, but which, nevertheless, was a hard nut to crack.
Riemann presented a function which is “discontinuous” for incredibly many of its
values, but nevertheless harmless—meaning that it can be integrated, and therefore
qualifies as a legitimate object of calculus.
If Riemann’s construction is reversed, starting from his integral, one gets an
actual oddity, namely a derivative—a function!—which is “discontinuous” for a
great plenitude of values, in Dirichlet’s view: with a great plenitude of jumps. “A
great plenitude” means: as many as there are fractions, and also ordered like them.
The discontinuities of this derivative have the following properties:
1. They are infinitely many.
2. In between any two of them, there is another.
Such an enormous amount of discontinuities, that’s a bit much! Nevertheless,
this function can be integrated. With the help of a trick, Riemann guaranteed for the
discontinuities: the closer they get, the more harmless they become. No doubt, this
is a smart idea!

Weierstraß’ Shocking Function

On 18 July 1872, Weierstraß presented a mathematical subject to the Royal


Academy of Sciences in Berlin, which was suitable radically to transform the
existing picture of analysis.
Two years after this event Leo Koenigsberger (1837–1921) still writes the fol-
lowing in his textbook with the title “Lectures on the Theory of Elliptic Functions”:
One of the main doctrines of analysis is the fundamental theorem that to each function of a
real variable there really belongs a differential quotient, i.e. that the ratio of the increments
of the function and of the variable can only be zero or infinite or is discontinuous because
170 12 Weierstraß: The Last Effort Towards a Substancial Analysis

of finite jumps in singular points of a finite length. However, for the rest it has finite values
which are independent of the infinite small growth of the variable . . .

According to Koenigsberger, basically each function can be differentiated (“to


it belongs a differential quotient”)—apart from possibly a few singular values (at
these “singular values” the derivative may not be defined).
This “main doctrine” or this “fundamental theorem of analysis” is totally wrong.
This was shown by Weierstraß in his momentous lecture on 18 July 1872, where he
proved the following:
Theorem. There exist continuous functions which cannot be differentiated at any single
value.

That Bolzano had already given such a function 40 years earlier (p. 113),
remained unknown for another 40 years.
This theorem states the contrary of what Koenigsberger still says in his textbook
2 years later and which he calls one of the “doctrines” of calculus.
A. Koenigsberger claimed in 1874 that a continuous1 function can generally—that
means: for nearly all values—be differentiated.
B. Weierstraß proved in 1872: there exist continuous functions which cannot be
differentiated for any single value.
A greater contrast is hardly imaginable.

The Aftermath of Weierstraß’ Construction

Weierstraß had presented a shocking function; soon functions like that were called
“capricious”. In retrospect, we may say: analysis was at a crossroads. (i) Should
one say: this we did not mean, such crazy “functions” we do not want, this counter-
intuitive stuff we bar from analysis? (ii) Or should one say: okay, up to now we
had a wrong idea of analysis—we must change our attitude and think much more
carefully about the relations of our analytical concepts?
We remember that the exotic function which Dirichlet presented, rather casually
in 1829, did not cause any reaction. Neither Dirichlet nor any of his contemporaries
(as far as is known) responded in some way or another. Dirichlet’s capricious
function (as it was called later) was simply ignored in 1829.
43 years later the world had changed. Not a single leading analyst of that time
thought of putting Weierstraß’ oddity aside and to return to do business as usual.
Quite the reverse, immediately the analysts of that time set out to create other
exceptional functions. (Even a name was coined: “capricious”.)
This started the search for a version of a “Value Analysis” which was as general
as possible. Surely it is to Weierstraß’ great merit that he triggered this development.

1 We can assume that his omission of “continuous” was quite unintentional.


What is Weierstraß’ Understanding of a “Function”? 171

What is Weierstraß’ Understanding of a “Function”?

In 1854, Riemann proposed at the very beginning of his doctoral thesis that a
“function” is nothing else than this: to each “value” of an independent variable
there corresponds exactly one “value” of the function (which is “the value of the
function”—see p. 155). Today this opinion of “function” is generally accepted.
However, the victory of Riemann’s concept of function was by no means
guaranteed. It was Weierstraß who fought nearly all his life against this, according
to his own opinion, wrong understanding of the subject. Weierstraß justified his
objection in clear words. In his lecture on the foundations of the theory of analytic
functions, in 1874, he stated:
This general definition of the function is the same as the following geometrical one: a curve
is a line which is not straight in any of its parts. From such a definition there cannot be
deduced any positive property of the defined.

The last sentence explains what Weierstraß demanded from a correct definition
of a mathematical concept: it must be possible to develop the relevant properties
from the definition of an object. In short: the definition has to designate the essence
of that object.
This coincided with the theory of definitions which Western philosophy had
taught since Antiquity, i.e. those started by Aristotle or by Leibniz. (And, very
interestingly, even by Georg Cantor, in 1883.) For clarity, I am going to coin a
word for this basic view of mathematical thinking and will speak of a “substancial”
definition in this case. Here, “substancial”, one of the archaic forms of “substantial”,
draws attention to the fact that the given definition has theoretical substance.
Clearly, such a “substancial” definition does not drop from the sky. It has to be
acquired through mathematical working. This was explained by Weierstraß quite
explicitly in his lecture of 1874:
It is impossible to define the concept of an analytical function in a few words—instead, this
has to be developed little by little and this is the task of this lecture.

Obviously, the result of this Weierstraßian lecture cannot be given in a few


words. Instead, we shall at least name the most important concepts which Weierstraß
needed in order to define his concept of function. They are: (1) “polynomial”,
(2) “infinite series”, (3) “uniform convergence” as well as “domain of convergence”
and (4) “development of a power series centred at a value”.—All of them being far
from elementary concepts.

A Sudden Change

For nearly all of his scientific life, Weierstraß was thoroughly convinced that
Riemann’s concept of function was of no worth—precisely because it is devoid
172 12 Weierstraß: The Last Effort Towards a Substancial Analysis

of substance and it is impossible to derive any important property of a “function”


therefrom.
However, Weierstraß recasted his position in the very last lecture of his life on
this topic, in 1886. At the age of 75 this mathematical Nestor changed his mind
and advocated the opposite of his previous position. “Continuity” was the only
property which he sustained, but besides this he subscribed to the supposition “that
the function is a so-called arbitrary: it is lawless”.
How come? A late revelation?
Not at all! Even later, in 1886, Weierstraß was still a proponent of substancial
thinking. Instead, his change of mind regarding the proper definition of a “function”
was changed by a basic mathematical result he had reached in the preceding year.
In 1885 Weierstraß had formulated and proved a profound theorem of analysis
that still bears his name: the “Weierstraß Approximation Theorem”.
Theorem. Each function which is continuous within a closed interval can be uniformly
approximated by polynomials.

In other words: Weierstraß had succeeded in drawing from the following,


seemingly simple, two assumptions
(a) the domain is a closed interval, and
(b) the “function” (following Riemann) is continuous in this domain,
the essence of a “function”. According to his view, the essence of a “function” is
constituted by a power series, at least nearly. (This nearly is specified with the help
of the technical term “uniformly approximated”.)
Weierstraß was now prepared to accept Riemann’s “empty” concept of
function—for meanwhile he had succeeded in deducing from two seemingly simple
additional phrases (finite and closed domain, continuity therein) an important
mathematical content (uniform approximate representation of a function through a
polynomial sequence). On Friday, 25th of June 1886, he announced the following
definition:
If to any system of values (u1 , u2 , . . . , un ) there corresponds one and only one value x of
the function, we call this a unique[ly determined] function of the variables u1 , u2 , . . . , un .

To sum up: although Weierstraß revised his mathematical view, he retained his
basic philosophical conviction. He continued to insist on substancial foundational
concepts.

Weierstraß’ Concept of Number

|(
We do not know any publication by Weierstraß on his concept of number.
Nevertheless, his thinking about this concept is documented in some detail. This
is due to two facts: first, he always stated his ideas on the concept of number at
Weierstraß’ Concept of Number 173

the beginning of his lecture cycle called “Preliminaries to the Theory of Analytic
Functions” and secondly, there do exist several and sometimes very detailed notes
taken from these lectures by students.
Nevertheless, it seemed impossible to figure out a clear-cut concept of (real and
complex) numbers from these notes. That’s why I proposed, in 1990, the view that
Weierstraß worked on this concept all his life.
However, I have changed my mind since. Weierstraß does have a clear-cut
concept of (real and complex) numbers. The real problem was: his students were
unable to understand him and therefore they could not correctly report on it!
This is proved, for instance, by the lecture notes taken by Emil Strauß (1859–92)
in winter 1880/81. They were discovered only recently and that by great chance: at
the mathematical library at Frankfurt/Main university, in August 2016.

The Fundamental Hindrance That Obstructed the


Understanding of Weierstraß’ Concept of Number

Only after having completed the German original of this book in winter 2016/17,
was I able finally to solve the riddle. The clue is this: Weierstraß relied on a concept
not one historian of mathematics dared to think of: the concept of a (pure) set.
As we have noticed by now, Weierstraß was actually a conservative thinker, and
set-theory was invented by his much younger pupil Georg Cantor (1845–1918) and
by Richard Dedekind (1831–1916), from the second half of the 1870s onwards.
Weierstraß never related to this theory.
Nevertheless, as Strauß’ lecture notes clearly document, Weierstraß relied on the
concept of set!
However, only on the concept—that means: Weierstraß did not make “set” the
topic of mathematical reasoning. He did not invent set-theory. Weierstraß only used
the concept of “set”. Again and again he proved the associative and the commutative
law for unions, as well as for products of sets in special cases—and that with the
help of analytical arguments!
Nevertheless, Weierstraß’ construction of the real as well as of the complex
numbers is a true set-theoretical one. Which is to say: only after the invention of set-
theory (in the twentieth century) Weierstraß’ construction of real numbers developed
into being the most elementary we know today.
Yet Weierstraß’ students had no chance of understanding him. For, not surpris-
ingly, Weierstraß never placed any emphasis on the concept of set. He did not
even coin a name for it. Why should he? In his thinking it was a mere auxiliary
concept and, indeed, a very trivial one. Why emphasize it? (Again, an excellent
mathematician is not necessarily an excellent teacher.)
Consequently, this new concept escaped the understanding of his students—and
what is not understood will not be adequately recorded. Normally.
174 12 Weierstraß: The Last Effort Towards a Substancial Analysis

The Peculiarity of the Student Emil Strauß

The student Emil Strauß was somewhat peculiar. In his lecture notes he did not
(only) write down what he really understood (and valued as important)—but plainly
everything he thought he might grasp. In one, very curious, case his writing actually
amounts to complete mathematical vacuity—nevertheless this passage is highly
valuable for the historian because it shows a sharp attack by Weierstraß on his
colleagues; whereas parallel lecture notes, taken in the same lecture by other
students (among them e.g. Adolf Kneser (1862–1930)) do not even show a hint
of this attack.
Thus Emil Strauß’ notes contain some sentences which clearly prove that
Weierstraß operated with (pure) sets. You only have to take these single sentences
literally—which is not an easy task! It took me a considerable amount of time to
understand this. However, if the result finally adds up to a coherent mathematical
statement, one can be fairly certain that one is embracing Weierstraß’ idea.
Weierstraß’ idea is truly simple: he takes the utmost possible generalization of
decimal numbers!

Preliminary

This generalization is carried out in three easy steps, which all follow this
assumption:2
Preliminary: We take the natural numbers as well as the fractions as given.

This Preliminary comprises Weierstraß’ utmost precise construction of the


natural numbers as well as the fractions from scratch. This foundational part is
a philosophical construction, and as our topic is analysis—not foundations—, we
have to omit this here. However, I cannot resist the temptation of disclosing my
judgement: Weierstraß’ philosophical construction of the natural numbers is, in my
opinion, the most precise which is known today, including that of Edmund Husserl
(1859–1938), who attended Weierstraß’ lecture in winter 1880/81.

The Construction

We proceed with the construction of Weierstraß’ real numbers:

2 As already announced as well as justified in the introduction (see p. xxii), in the following I do

not stick to my historiographical method, but instead I present a former idea completely in modern
language.
Weierstraß’ Concept of Number 175

1. Disregard the sign and start with the interpretation


 of a decimal  number d0 .d1 d2 d3 . . . to
d0 d1 d2
be a (sometimes infinite) set of fractions: 10 0 , 101 , 102 , . . .

Now we go beyond this and allow that


2. the nominators and
3. the denominators can be any natural number.

That’s all! These three steps give us the utmost general concept of “irrational”
number, together with the two direct operations “addition” und “multiplication”.
We now have the following
Definitions (addition, multiplication).
1. The “addition” of two such numbers is their set-theoretical union— followed, of course,
by the addition of each two elements with the same denominator.
2. The “multiplication” of two such numbers is, what we may call the “Weierstraß
Product”: multiply each element of one number with each element of the other number
and finally add in the resulting set all fractions with equal denominator.

These two operations are well-defined. They are associative as well as commu-
tative and they even obey the distributive law—because these laws hold for sets and
the natural numbers and their operations.
This was truly easy, wasn’t it? Unfortunately, we are not yet through!
What is missing? Well, we have the set of all imaginable (real) numbers (without
sign) you may dream of. We are able to add and to multiply them—but do not know
how to compare them yet! But of course, we need equality as well as comparisons
like, e.g.:
     
1
3 = 3
, 3 , 3 ,
101 102 103
. . . < 12 !

Further Preliminaries for Tackling a Real Difficulty

How shall we deal with this plenitude of numbers? Where should we begin?
Since Weierstraß’ numbers are finite as well as infinite sets, we shall start with
the finite ones. Weierstraß himself gave no one name to his numbers, but we shall
do. As the reader may expect:
Definition (fraction). Each number which is a finite set, will be called a “(general) fraction”.

If we were pedantic we would choose the name “general fraction”, for we already
have constructed the “fractions” in our initial Preliminary. However, things seem
fairly clear and therefore we might be allowed to dispense with this “general”. Thus,
let us stick for one more moment with the attribute “general”, for we have also to
state the following:
176 12 Weierstraß: The Last Effort Towards a Substancial Analysis

Definitions (equality, transformation).


1. Two (general) fractions are “equal”, if one of them can be transformed into the other.
2. A “transformation” of a (general) fraction is any finite series of changes of one or more
of its elements (a) either by cancelling or its opposite, or (b) by splitting it into two or
more fractions or vice versa.

   
Therefore, we have e.g. 1 1
2, 4 = 5 1
7 , 28 , since

1
2 + 1
4 = 3
4 = 21
28 = 5
7 + 1
28

from the Preliminary. Similarly, we state:


Definition (less than). A (general) fraction a is “less than” (<) another number b if b can
be transformed such that a ⊂ b.

Caution: you need to transform b, for we need to have

1
3 < 1
2 = 2+1
6 = 1
3 + 1
6 ,

and this you cannot get by transforming a. Please, remember: this definition is
not an arithmetical, but a set-theoretical one!—It is true: Weierstraß went for the
arithmetical version, not for our set-theoretical one.
We opted in favour of the set-theoretical definition for “less than”, because it is
Weierstraß’ solution in the case of the infinite sets. The real difficulty of this number
concept is to define equality and to make comparisons for infinite sets—because the
condition to permit only finitely many transformations is here not enough. We need
to have
       
1 = 2, 4, 8, . . .
1 1 1 1 1 1 1 1 1 1
as well as 6 , 12 , 24 , . . . < 2 , 4 , 8 , . . . ,

but one cannot carry out infinitely many transformations.

Solution of the (Perhaps) Real Difficulty

The concepts of equality and comparison in cases where numbers which are infinite
sets are involved, seem to be the real difficulty of this number concept. These
definitions do not come easily.
Really? Don’t we have the fractions (the finite sets)? Doesn’t it suggest that we
should take them to be our standard?
To make his students aware of the difficult notion of equality in the case of infinite
objects (sets), Weierstraß presented the following “fallacy” (as he called it): starting
with
Weierstraß’ Concept of Number 177

x= 1
2 + 1
4 + 1
8 + ... ,

one gets

x+1=1+ 1
2 + 1
4 + 1
8 + . . . = 2x ,

and this implies

x = 1.

Weierstraß comments: “This conclusion uses a ‘law of equality’, which has not
yet been stated.” (Interestingly, only the lecture notes of Emil Strauß document this
passage—and they show the (obvious) mistake “1 + x” instead of “2x”—, whereas
e.g. the notes of the listener Adolf Kneser do not.)
Be that as it may, Weierstraß already had an idea how to solve this problem, and
according to the sources, he had it already in, or (presumably) even before, 1868.
That was at least 4 years earlier than Dedekind—who published the same idea, albeit
with quite a different interpretation. The idea is this:
Definitions (equal, less than). A (general) fraction f is “less than” a number q, if there
exists a finite subset g ⊂ q with f < g. Two numbers are “equal”, if each fraction less than
one is also less than the other; if only one part of this condition is fulfilled, the relation “less
than” (<) resp. “greater than” (>) is defined.

In 1872, this was Dedekind’s idea of defining the (irrational) numbers (see
p. 202). However, Weierstraß had used this idea to compare the (irrational) numbers
already in 1868 or even earlier. We see the difference in their ways of thought in
those approaches: Weierstraß used the idea of comparing his numbers, which he had
previously defined (Weierstraß is a “substancial” thinker!), while Dedekind used
this relation to define his numbers (therefore, I propose to call him a “relational”
thinker).
The arithmetical laws for these relations (=, <) are easily proved.
Weierstraß abhorred infinite numbers (he sided with Leibniz and Cauchy against
Johann Bernoulli and Euler) and proposed the
Definition
  (finite).
 A number a is called “finite,” if there exists a natural number
n := n1 with a < n.

(Weierstraß did not even dream of infinite natural numbers—at most in night-
mares. However, what would Johann Bernoulli have said, see p. 52?)
And then Weierstraß banished the infinite numbers by calling them all together
“∞”.—Consequently, ∞ is not a true number, as it violates the arithmetical laws.
178 12 Weierstraß: The Last Effort Towards a Substancial Analysis

The Benefit of Structural Thinking


 
If we add the number 01 , these finite Weierstraßian numbers constitute what we
call nowadays a monoid. Actually, we even have two monoids: one for each of the
two operations, addition and multiplication.
Today, at least in our second semester at university, we are taught how to enlarge
a monoid (M, ◦) to a group (G, •, ): we just take G to be M × M and operate this
way:

(a, b) • (c, d) := (a ◦ c, b ◦ d) ,
(a, b) (c, d) := (a ◦ d, c ◦ b) ,
(a, b) = (c, d) :⇐⇒ a ◦ d = c ◦ b ,

where “ ” is the “inverse” operation of “•”.


Actually, this is just what Weierstraß did with his finite numbers, taking them to make up our
set M! Exactly in this way he constructed his “real” numbers to be pairs of his “irrational”
numbers. And since his “irrational” numbers M constitute two monoids, he actually got a
ring G = M × M with the unit 1.

Heavy stuff, indeed! However, for those familiar with elementary modern algebra
(after Bourbaki), It’s mere child’s play. Nonetheless, it’s mere really astonishing that
Weierstraß managed to come up with this construction, without any algebra to hand!
Clearly, not one of his students, not even the most talented, was able to grasp this
construction. Besides, the concept of “ordered pair” had not yet been created (this
only happened in the twentieth century). The same holds for the use of arbitrary
symbols for operations. Inevitably, Weierstraß’ students were unable to distinguish
between our “◦” and “•” from above, and, even more dramatically, they confused
the formal “ ” with the arithmetical (“true”) “−”—because Weierstraß always used
“+” as well as “−”! Weierstraß was not a protagonist of mathematical revolutions,
but he was an enormously rigorous thinker.3
That is why we could not access this idea before Emil Strauß’ lecture notes where
found (in August 2016) and my subsequent interpretation. It was very lucky that
Weierstraß disclosed his view of the numbers in (nearly) all its elementary details
in winter 1880/81—and this in the presence of this ambitious student. Of course,
Weierstraß did not go into all these details in each of his foundational lectures!

3 In case he knew that a = ā + c as well as b = b̄ + c, he easily concluded (a, b) = (ā, b̄) for

all irrational numbers a, b. The decisive point is, that he already knew the first two equations—for
otherwise he would have used general subtraction of his irrational numbers, which is impossible.
Weierstraß’ Concept of Number 179

The Benefit and the Disadvantage of Structural Thinking

Well, even a Weierstraß was not perfect and could make an error. Weierstraß taught
procedures to divide his irrational as well as his real numbers—in fact, in different
lectures he proposed two different methods.
However, our structural analysis of his construction genuinely proves: they are
both wrong. The reason is that this idea of construction only produces a ring, not
a field. And it is true: both of the methods lectured by Weierstraß are erroneous.
They both rely on the possibility of subtracting “irrational” numbers—but this is
impossible. (In hindsight: if it were possible, it would not have been necessary to
construct the operation of subtraction in a pure formal way—as it is done by taking
M × M which Weierstraß actually did.)
This line of thought inevitably leads to the question: Do Weierstraß’ real numbers
form a ring (with 1 as unity) or a field? We may put it this way: can Weierstraß’
real numbers be divided, yes or no?
This question is very productive. It can be answered in two ways:
1. The real numbers are objects—it does not matter how they are constructed.
These objects can be taken as subjects of a structural analysis. (You may
construct the integers by adding the “signs” + and − to the natural numbers
and zero—nevertheless, we obtain the structure of a group from this monoid.)
2. Whether some objects form a certain structure depends on the definition of
“operation” taken.
This is a book neither on algebra nor on philosophy of mathematics, but solely
one on the history of analysis. Therefore, we have to limit ourselves to elementary
aspects of this topic and merely state two things:
• There is a fundamental difference between direct operations and their reverse,
the inverse operations. The latter are methods of trial and error, they cannot be
carried out directly.
This fundamental difference between these two kinds of operation is smashed
by taking the algebraic point of view. It makes a great difference whether we
say: “An operation is a method of generating a new object from two given
ones.” or if we state: “An operation is defined whenever there exists an object
which fulfils certain conditions.” The first standpoint is Weierstraß’, the second
is Hilbert’s.
• We all know how to subtract and how to divide decimal numbers (we just take
suitable approximations). We have learnt this in school. Consequently, we are
all able to subtract and to divide Weierstraß’ real numbers—we only need to
take them in their decimal (or binary or some equivalent) form.
We may put this in higher algebraic language and say elaborately: (i) Equality is
an equivalence relation for Weierstraßian real numbers. (ii) In each equivalence
class there exists a decimal (or . . . ) number—which in Weierstraß’ sense may
be called the “value”. (iii) The decimal (or . . . ) numbers are complete. (This
proof is due to Hilbert.)
180 12 Weierstraß: The Last Effort Towards a Substancial Analysis

Therefore, we answer the above question in a personified way:


In Weierstraß’ view, his real numbers do not constitute a field, whereas according to
Hilbert’s view, they do.

(That Weierstraß himself, erroneously, held his real numbers to be a field, is a


historical fact.)

Postscript

It was as late as 1994 when John Horton Conway (1937–2020) characterized the
easiest way of constructing the real numbers:
Proceed from the natural numbers to the non-negative rationals (or the strictly positive ones
if you prefer), then construct the non-negative (or positive) reals from these, so having no
sign-problem, and then construct signed reals from these in the way that we constructed
the signed integers from the natural numbers. I think that this is in fact the simplest way to
construct the real numbers along traditional lines.

Nevertheless, neither Conway nor any other mathematician ever since came up
with this construction of the real numbers—although, as we have seen, it is really a
straightforward generalization of the decimal numbers.
The inevitable conclusion is, yet again, the well-known truism: Mathematics is
REALLY difficult!
|)

Infinite Series

With his construction of the general concept of “quantity”—which is our “real


number”—Weierstraß was able to do analysis. In particular, he built the concept
of “infinite series” or simply “series”.
Weierstraß repeated the method used for his numbers step-by-step. He started
with the study of “infinite series” of “fractions”, followed by those of “irrational
numbers”. That is, he considered the series

a1 + a2 + a3 + . . . ,

where ak are irrational numbers. Each of it has the form


(k) (k) (k)
ak = α0 · 1 + α1 · 1
n1 + α2 · 1
n2 + ...

(k)
wherein the order does not matter. Taking σi to be the sum of all multiples αi of
 (k)
the fractions n1i for all the k terms of the series (i.e. σi = αi ) one gets
k
Infinite Series 181

 
s= ak = σi · 1
ni .
k i

(Of course, Weierstraß did not use the symbols of set-theory—this is our modern
reading of his original concepts.) Therefore, to assure the finiteness of the series
Weierstraß demands:
1. All irrational numbers ak of the series must be finite. (Clearly, this demands
 (k) (k)
that in all summands ak = αi · n1i the numerators αi of the fractions with
i
denominator ni have to be finite; however, this condition is only necessary but
not sufficient.)
2. Each “element” of s (which is to say, each σi · n1i ) has a finite multiplicity σi ,
or: all σi are finite.

Subsequently, this “sum” s = ak is made up as follows: in each “element” the
k
numerator of the fraction with denominator ni has to be σi , which was built from all
the irrational numbers ak of the series.
The well-known example 1 · 1 + 1 · 12 + 1 · 13 + . . . shows, that the given conditions
are not sufficient. Therefore, the relevant question is:

In which cases is the series s = ak of the irrational numbers ak finite?
k

To answer this question, we need to know Weierstraß’ concept of the “summability”


of a series.

Summability

Since Cauchy we are used to defining the concept of “convergence” of an infinite


series

u0 + u1 + u2 + . . . = sn + rn

to be the condition

lim rn = 0 .

Weierstraß gave another concept, and, to make things more clear, he also chose
another name: “summability” (“Summierbarkeit”). His notion is this:
Definition (summability). An infinite series

a1 + a2 + a3 + . . .

is called “summable” if its sum is finite (see p. 177).


 
As we already know: the sum s of the series ak is σi · 1
ni .
k i
182 12 Weierstraß: The Last Effort Towards a Substancial Analysis

Unfortunately, Weierstraß did not state this definition. We have to deduce it from
the text or, more precisely: from the name he used for the concept.

Summability for Series of Irrational Numbers

Weierstraß gives a criterion for the sum of an infinite series to be finite. In the case
of a series of irrational numbers, he says (his style is moderately updated):

Theorem. The infinite series ak of irrational numbers ak is finite if and only if it is possible
k
to determine a number g, which is larger than each sum of finitely many ak .

Weierstraß proves this theorem by demonstrating itto be “correct”. That means:


if and only if the stated criterion is fulfilled, then s = ak is finite.
k
Clearly, if the sum of the infinite series is finite, the given criterion is valid. But
what about the opposite direction?
If the specified criterion is fulfilled, we
 shall get the sum s of the series to be
s = lim g, i.e. the least upper bound of ai . However, Weierstraß did not say a
i
single word about how this sum can be defined. That is why his proof of the above
theorem cannot be understood that easily. It plainly lacks the following definition:
Definition (sum of an infinite series, summability). The sum of an infinite series is a (finite)
number which is the least upper bound of any finite sum of its terms.

Nowadays we are to take this definition to be the specification of Weierstraß’


definition of “summability”.
Weierstraß even proves the following:
Theorem. If the infinitely many irrational numbers ak are arbitrarily grouped, and the sums
of these groups are built, the result will always be the same.

In the course of the long proof, some important theorems on irrational numbers
are revealed.

Theorems on Irrational Numbers

Theorem. Each finite irrational number a, with infinitely many elements αi · 1


ni can be
represented in this way:

a = a0 + a1 ,

where (i) a0 has finitely many elements; (ii) a1 can be made arbitrarily small.

This statement may be expressed this way: each finite irrational number a can be
approximated by a fraction (a0 ) to any degree of accuracy (a1 ).
Infinite Series 183

This leads Weierstraß to his theorem for the “equality” of two finite irrational
numbers:
Theorem. Two finite irrational numbers a and b are equal if and only if it is possible to split
each of them into two: a = a0 + a1 and b = b0 + b1 (a0 and b0 being fractions) such that
|a0 − b0 | < ε as well as a1 < ε and b1 < ε are valid for any previously given ε.

(It is easy to see that the subtraction of a (general) fraction is possible.) Stated
without formulae:
Two finite irrational numbers are equal if and only if it is possible to split each of the
quantities into two, an approximating fraction and an (irrational) excess, such that the
difference of these approximating fractions as well as each of the two remainders in itself is
less than any previously given tolerance ε.

Formulated this way, the theorem seems to be clear. The technical proof in the
lecture notes comprises one large page. Relying on this theorem, Weierstraß “easily”
proves the following
Theorem. In an infinite sum, equal irrational numbers can replace each other.

Summability for Series of Real Numbers

The addition of real numbers is different from that of irrational numbers. (Conway
did not ponder on that!) Therefore, in case of series of real numbers another criterion
for summability is required. Weierstraß presents it as follows:
Main Theorem. If the series x1 , x2 , x3 , . . . has the property that there exists a finite number
g such that the absolute sum of arbitrarily many and arbitrarily chosen terms is less than g,
then the series is summable.—The reverse also holds.

To prove this, Weierstraß argues thus:


(a) Take the positive of the chosen terms xk (i.e. the pairs xk = (ak , bk ) of irrational
numbers ak , bk with ak > bk ). One can change each of them in such a manner
((ak , bk ) = (ak , bk )) that its “negative” part (bk ) is as small as one wishes.
(b) The same can be done with the negative terms xk (i.e. the pairs xk = (ck , dk ) with
ck < dk ) such that xk = (ck , dk ) where ck is as small as one wishes.
(c) Consequently, it is possible to change the positive of the chosen terms in such a manner
that their “negative” parts (bk ) are less than the terms fk of a converging series fk =
f.
(d) The same can be done with the negative of the chosen terms: their “positive” parts (ck )
can be made less than the terms hk of a converging series hk = h.
(e) Consequently, the series can be replaced by (i) a sum of positive (ak ) (ii) as well as
“negative” fractions (dk ) plus (iii) a sum of the “negative” irrational numbers (bk ),
which if taken absolutely (bk ), have a finite value (< f ), as well as (iv) a sum of
irrational numbers (ck ), which is also finite (< h). (This is true even if the series is not
summable!)
(f) However, the criterion demands that there exists a finite number g such that the absolute
sum of any finite set of terms xk is less than g. Consequently, each of these sums is less
than (g + f ) + (g + h).
184 12 Weierstraß: The Last Effort Towards a Substancial Analysis

(g) Therefore, each absolute sum (of finitely many) terms of the series is guaranteed to be
less than 2g + f + h.
(h) From this, Weierstraß concludes that the series is summable—which proves the Main
Theorem.

As Weierstraß operates with the absolute value of the partial sums, his last
conclusion relies again on the (unsaid) concept of the least upper bound, as it has
been in the case of series of irrational numbers.

The Concept of “Convergence”

Weierstraß uses the notion of “convergence” only towards the end of his lecture
(on p. 155 of Strauß’ notes). There he even identifies it with his concept of
“summability”.
When speaking of “convergence”, Weierstraß relies of course on the order of
the terms in the series. After having introduced the “complex” numbers, Weierstraß
proves the following:
Theorem. If the (complex) series is summable and has the sum s, it is possible to determine
to any given ε 4 a natural number n such that

|s − sn | < ε .

And he adds: “Because of this theorem one calls a summable series also convergent
and says that the sum sn converges to the value s.”
Weierstraß continues: if the concept of “limit” of a series is defined, this notion is
taken to be equivalent to the concept of its “sum”. However, the value of this “sum”
may depend on the order of the terms. Series with this property Weierstraß suspends
“for the time being”, closing with the following
Definition ((un)conditional convergence). If the order of the terms influences the value of
the sum, the series is called “conditionally convergent”; otherwise it is called “uncondi-
tionally convergent” (or “absolutely convergent”—in case of real numbers both notions are
equivalent).

Upshot of Weierstraß’ Concept of Number

Weierstraß takes the concept of “series” to be a generalization of the concept of


“sum”. That is why he dispenses with the idea of order when dealing with series.
Therefore, his notion of the finiteness of the sum of a series, namely “summa-
bility”, is narrower then our concept of “convergence” (which demands ordered
series). Consequently, Weierstraß’ “summability” coincides with our concept of
“unconditional convergence”.

4 Weierstraß, indeed, uses “ δ” . . . !


Weierstraß in Retrospect 185

Results:

1. Weierstraß succeeds in founding analysis on a well-established concept of


(irrational, real and complex) number.
2. Additionally, Weierstraß succeeds in formulating a substancial concept of
“equality” for “irrational” numbers.
3. In his untimely anticipation of an algebraic construction Weierstraß completes
the monoid of his irrational numbers to a group, thereby giving a unique
elementary construction of the real numbers, which has not been repeated since.
4. Because Weierstraß took the concept of “series” to be a generalization of
the concept of “sum”, his notion of finiteness for the sum of a series (called
“summability”) is narrower than the usual concept of “convergence” used
nowadays: “summability” is “unconditional” convergence.

Weierstraß in Retrospect

Weierstraß was not the first to define clear concepts of analysis, nor did he invent
epsilontics. He learnt all of that.
Nevertheless, he pioneered other foundational landmarks in analysis:

1. In his “Approximation Theorem” Weierstraß succeeded in developing cal-


culatory expressions (“polynoms”) as a rather good approximation for any
“continuous” function, which has a “compact” domain. This paved the way for
Riemann’s concept of a function in analysis.
2. Weierstraß actually developed his concept of the real numbers constructively.
Albeit, this simple and elegant construction came far too early and therefore
could not be expressed in a suitable language. This is also why his students,
even the most gifted, were unable to grasp it. Mysteriously, this concept was
not re-invented in fourteen decades.
3. Weierstraß’ idea of turning “capricious” functions into legitimate objects of
analysis paved the way for “Value Analysis” finally to become a very general
theory of “Functional Analysis”.

Weierstraß’ thinking was clearly traditional—I proposed the name “substancial”.


He was not keen on manipulating formulae: quite contrarily, he substantiated every
step within his construction of the numbers.
And it is a real paradox of history that Weierstraß was (to my knowledge) the
very first mathematician to give a strictly formal construction of a group by forming
the Cartesian product of a monoid with itself—even decades before the concept of
“ordered pair” was created, not to speak of the concept of a “monoid”.
186 12 Weierstraß: The Last Effort Towards a Substancial Analysis

Today’s judgement is inevitably this: Weierstraß has given a precise concept


of real number; (i) it is totally systematic, (ii) it does not rely on the concept of
“negative” number and (iii) it is the most elementary known to this day.
However, nothing originates from nothing. Weierstraß has to pay a price for his
concepts to be of such an elementary character. This price is the lack of inverse
operations: Weierstraß’ irrational numbers cannot be subtracted nor divided (see
p. 179). To implement the inverse operations for Weierstraß’ irrational numbers one
has to change from his most elementary concept to a more ambitious one, to a
“positional system” of numbers. But the way of subtracting and dividing decimal
numbers we are all taught in school.

Literature

Cantor, G. (1883). Über unendliche lineare Punktmannigfaltigkeiten (No. 5: Grundlagen einer


allgemeinen Mannigfaltigkeitslehre). Mathematische Annalen, 21, 545–586. cited following
Zermelo 1932, pp. 165–209.
Conway, J. H. (1994). The surreals and reals. In Ehrlich, P. (Ed.). Real numbers, generalizations
of the reals, and theorems of continua. Synthese library, Studies in epistemology, logic,
methodology, and philosophy of science (Vol. 242, pp. 93–103). Dordrecht, Boston, London:
Kluwer Academic Publishers.
Dirichlet, G. L. (1829). Sur la convergence des séries trigonométriques qui servent à représenter
une fonction arbitraire entre des limites données. Journal für die reine und angewandte
Mathematik, 4, 157–169. See also Kronecker & Fuchs (Eds.) 1889, 1897, vol. 1, pp. 117–132.
Hettner, G. (1874). Einleitung in die Theorie der analytischen Funktionen von Prof. Dr. Weier-
strass, nach den Vorlesungen im S[ommer]s[emster] 1874. mimeographed by the Mathematical
Institute of the Göttingen university (1988).
Kneser, A. 1880/81. Weierstrass, Einleitung in die Theorie der analytischen Functionen. Staats-
und Universitätsbibliothek Göttingen, manuscript, signature Cod Ms A Kneser B 3.
Koenigsberger, L. (1874). Vorlesungen über die Theorie der elliptischen Functionen. Leipzig:
Teubner.
Kronecker, L. (vol. 1) & Fuchs, L. (vol. 2) (Eds.) (1889, 1897). G. Lejeune Dirichlet’s Werke.
Berlin: Georg Reimer.
N. N. 1886. Ausgewählte Kapitel aus der Funktionentheorie, Vorlesung von Prof. Dr. Karl
Weierstraß, Sommersemester 1886. In: Reinhard Siegmund-Schultze (Ed.), vol. 9 of the series
Teubner-Archiv zur Mathematik. Leipzig: Teubner, 1988.
Strauß, E. (1880/81). Weierstrass, Einleitung in die Theorie der Analytischen Functionen. Univer-
sity of Frankfurt am Main, archive, file number 2.11.01; 170348.
Zermelo, E. (Ed.) (1992). Georg Cantor. Gesammelte Abhandlungen mathematischen und
philosophischen Inhalts. Hildesheim: Georg Olms Verlagsbuchhandlung, reprint 1966.
Chapter 13
Analysis’ Detachment from Reality—and
the Introduction of the Actual Infinite
into the Foundations of Mathematics

A Pessimistic Mood

The Initial Situation in 1817

Bernard Bolzano was the first to introduce the concept of value into analysis in 1817
(pp. 101f). “Values” are, first of all, numbers as well as (usually) ±∞ (± infinity).
However, what is a number, precisely?
Basically, numbers arise from counting: 1, 2, 3, . . . Anyhow, that is not enough
for analysis. From the beginning, analysis was there to measure: areas, volumes,
gradients, curvatures, etc. But exact measurement needs more than just the natural
numbers 1, 2, 3, . . . ; additionally, it requires at least the fractions: 12 , 13 , 14 , . . .
However, the surveyors of classical Greece already knew:
Even fractions do not suffice for exact mathematics.

For example, it is impossible to express the exact length of the diagonal of a


square with side 1 in fractions. (This proof is very instructive and not complicated,
but it does not belong to analysis and so it is omitted here.)
In 1821, Cauchy repeated the age-old saying of the mathematicians:
“Number” is the ratio of two quantities.

This only transfers the problem to the question: what are “quantities”? but does
not solve it.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 187
D. D. Spalt, A Brief History of Analysis,
https://doi.org/10.1007/978-3-031-00650-0_13
188 13 Analysis’ Detachment from Reality—and the Introduction of the Actual Infinite. . .

Failure

For a long time, there was no progress regarding this issue. Since 1817, Value
Analysis progressed rapidly in many fields—but not in that of the concept of
number. No new idea stirred up the minds of the leading mathematicians. (Bolzano
tried a proposal; however, judged as a radical political thinker, he had been
dismissed from the academic world, and, consequently, his ideas, which were hidden
in unpublished manuscripts, remained unknown in the nineteenth century.)
Obviously, it was a problem. In the foregoing chapter, we have seen that Weier-
straß worked in order to solve it. Unfortunately, his solution was not understood and
consequently remained unknown.
In 1867, half a century after Bolzano’s important treatise, the gifted student of
Riemann, Hermann Hankel (1839–73), stated in a book the following pessimistic
judgement:
Every attempt to treat the irrational numbers in a formal manner and excluding the notion of
[extensive] quantity itself is bound to lead to the most abstruse and laborious artificialities
which—even if they could be carried out with complete rigour and we have every good
reason to doubt this will be possible—will not be of any favourable scientific value.

(The word “extensive” is added to elucidate Hankel’s intention.)


In other words: in 1867 Hankel was convinced that for any “foundation” of a
suitable analytical number concept, the recourse to intuition is inevitable—even if it
were to be possible to give a pure notion of number (that is one without any help of
intuition), this might lead to such complications (“abstruse”, “artificialities”) that it
would be of no help to science.

1872: Two at Once

Only 5 years later, Hankel’s conviction had been overcome and proven to be wrong.
Within an interval of only a few weeks, two (completely different) constructions of
the real numbers were published. Both constructions avoided any recurse to intuition
but instead were purely conceptual. The most elementary one, Weierstraß’ idea
(see pp. 174f), remained unknown until today. That is why these two constructions
are the ones still taught to students today. Usually just one of them, because one
definition of a concept is quite enough. Besides, on closer examination, it turns out
that both concepts of number amount to the same.
One generation later, in 1899, David Hilbert found a way of characterizing
this concept of number. Hilbert set up a system of axioms that lists all essential
properties of the real numbers. Now logicians were able to prove that any two
systems of numbers that fulfil those axioms are interchangeable. In other words:
strictly speaking, there exists only one single system of numbers that has exactly
these properties. We already hit upon this (pp. 179f), and we will return to it yet
again (pp. 205f).
Georg Cantor’s Construction of the Real Numbers 189

The inventors of these constructions of the “real” numbers are Georg Cantor
(who went on to invent set-theory) as well as the triumvirate Traugott Müller (in
1838), Joseph Bertrand (in 1849), and Richard Dedekind (in 1872, he also invented
set-theory). In the following, we will outline these two constructions.

Georg Cantor’s Construction of the Real Numbers

The Situation

First, we have to mention briefly Heinrich Eduard Heine (1821–81). The older
Heine was working together with the younger Georg Cantor at Halle University,
and as a consequence, each of them published a treatise on this topic at the same
time. Cantor’s paper was actually on another subject, and thus it contained the great
news only roughly. In contrast to this, Heine tackled the issue fundamentally and
pedantically went into all the details.
Aiming at a comprehensive presentation, Heine outlined in his introduction the
general situation of analysis in 1872. According to him, it was like this (all emphases
are his):
The progression of the theory of functions is essentially hindered by the fact that some
elementary theorems, although they are proved by an astute scientist, are still doubted. From
this it follows that research results, which are based on those theorems, are not generally
held to be valid.
A possible, and very likely, explanation is the fact that the principal teachings of
Mr Weierstraß disseminate in wide sections, directly via his lectures and through oral
informations as well as indirectly through copies from lecture notes which originate from
his courses. But they have not yet been published by himself in an authentic version.
Consequently, there exists no place where one can find the theorems gradually and
systematically developed.
However, the truth of these theorems relies on the not completely established definition
of the irrational numbers, which is frequently confusedly and negatively affected by
geometrical ideas, particularly by the generation of a line through motion.
These theorems are valid if the definition of irrational numbers given below is assumed
to be their foundation: where numbers are called equal if they do not differ by any given
number, being as small as one wishes, and where, in addition, the irrational number is due
to actual existence, such that a function has for each single value of the variable a definite
value, too, be it rational or irrational.
However, if one took another point of view, there could be very good reasons to make
objections to the truth of the theorems.

This extraordinarily lucid passage aptly formulates (like Cauchy’s preface to


his textbook of 1821 at that time) the situation of analysis in 1872, as well as
the possible changes caused by this new construction. “Possible”, because Heine
explicitly left open, whether the community of mathematicians will accept this new
idea of his fellow mathematician Georg Cantor or not. Obviously, Heine did not take
the mathematical/political success of this construction for granted—although he felt
capable of demonstrating its mathematical force (and actually did so).
190 13 Analysis’ Detachment from Reality—and the Introduction of the Actual Infinite. . .

Cantor’s Construction Arises Out of Weierstraß’

Georg Cantor (1845–1918) heard Weierstraß’ lectures. However, Cantor was not
Weierstraß and therefore he thought differently. This other way of thinking produced
an idea, which even today, one and a half centuries later, belongs to the fundamentals
of analysis. What is this idea?—A surprising one!
According to Weierstraß (we did not go into all the details), a real number is a
pair of two irrational numbers (cf p. 178):

k0 ∗ 1 + k1 ∗ n 1 + k2 ∗ n 2 + . . .
1 1

l0 ∗ 1 + l1 ∗ m1 + l2 ∗ m2 + . . .
1 1

What is written here as “1” or “ n1i ”, “ m1i ” was thought by Weierstraß to be a “unity”
resp. its “exact parts”, whereas the “ki ” and “li ” are their “multiplicities”. Each
of the above lines is a distinct part of the “real number”, the first being called
“positive”, the second “negative”—yet with a very special meaning (see p. 183).
And, last but not least: for Weierstraß, the + was a priori commutative, even in
case of infinitely many elements: Weierstraß thought an irrational number to be a
SET —a concept completely unknown to his contemporaries in the early 1870s.
Cantor introduced a fundamental change into this construction. In my opinion:
Cantor plainly misunderstood Weierstraß.
1. He completely dismissed the concept of “unity” (as being a useless philosophi-
cal notion). In other words, Weierstraß’ “unities” “ n1i ” were taken by Cantor to
be mere “fractions”.
2. Consequently, Cantor took Weierstraß “elements” to be the familiar “fractions
ki lj
ni resp. mj ”.
3. As fractions may have either a positive or a negative sign, Cantor was able to
remove the (ordered) pair! Because our “(a, b)” had not yet been invented, Wei-
erstraß wrote “A+(−B)” instead—and Cantor mistook Weierstraß’ formalism.
4. However, the real historical paradox is this: Cantor did not grasp that Weierstraß
used the concept of “set” to construct his “irrational” numbers! Instead, Cantor
relied on the then usual concept of ordered set and defined a “real” number to
be a sequence of positive or negative fractions.
5. In a complete misunderstanding of Weierstraß, Cantor finally demanded the
sequences to be “convergent”—in regard to the strict deductive procedural
character of Weierstraß’ teaching, it is absolutely clear that Weierstraß was
unable to introduce the deeply analytical notion of “convergence” into the
groundworks of his theory. But Cantor did.
All told, Cantor’s new concept of “real number” is this:

A “quantity” is a convergent SEQUENCE of rational numbers:


a1 , a2 , a3 , a4 , . . .
Georg Cantor’s Construction of the Real Numbers 191

Cantor’s definition reads this way:

This fact I express in the following words: “The sequence1 has a definite limit b”.

Cantor calls “this fact” the “limit”. “This fact” is a “converging sequence” (of
rational numbers). In accordance with Cantor, the converging sequence (of rational
numbers) is the new “real” number, the “limit”—although Cantor writes “has”
instead of “is”. He does not choose his words very carefully.
To emphasize again:
• Where Weierstraß thought “elements” to be multiplicities of “units” ki · n1i ,
Cantor simply assumes “rational numbers” ak = αnkk .
• Weierstraß’ “elements” of “aggregates”, deprived of order, are turned into well-
arranged “terms” of a “sequence”.
• Weierstraß’ formal ordering of two irrational numbers is skipped by Cantor;
instead, Cantor takes recourse to signed numbers (rational numbers).
• Most importantly: whereas Weierstraß thought a number to be a “sum”
(an elementary notion of arithmetic), Cantor implemented the deep idea of
“convergence”—which is an analytical concept. Consequently, Cantor’s notion
cannot (legitimately) be introduced as a preliminary to a deductive construction
of analysis.
Synopsis:
Cantor made the following changes to the notion of “quantity”:
1. The “elements” are turned from fractions without order into ordered “rational numbers”.
2. The purely formal “order” (of two irrational numbers to make up one real number) is
replaced by the introduction of real “signs”—in other words: the negative numbers are
taken as an actual formative ingredient.
3. The arithmetical “sum” is upgraded to an analytical “sequence”, together with its “limit”.

Each of these changes replaces elementary ideas by more complicated ones.


Taken together, they override all of Weierstraß’ elementary properties of “quantity”
and deprive his concept of “real” number of the building blocks for the foundations
of analysis.
In 1872, Cantor invented a clear concept of number, detached from all intuitive
ingredients—and technically broad enough to base wide areas of ordinary analysis
on firm ground. (The “conditionally”—or “not absolutely”—convergent series were
included.) Hankel was refuted—all was just great!
However: no progress without a price!

1 Cantor actually used the notion “series” (“Reihe”)!


192 13 Analysis’ Detachment from Reality—and the Introduction of the Actual Infinite. . .

The Philosophical Price of This Progress

Every major change is a matter of give and take. No progress (e.g. a technical
one) comes without drawbacks (e.g. social or ecological ones). Therefore, we ask:
what are the losses in regard to mathematics because of this conceptual change
proclaimed by Cantor (and Heine)?
The answer to this simple question is equally simple: its grounding in real life.
This is important, and we emphasize it:
The acceptance of Cantor’s concept of number separated analysis from reality.

Our “Brief History” has no room for such philosophical details, but nevertheless
we may say: the fundamental aim of the substancial thinker Weierstraß is to prove
that the basic concepts of his theory (which is analysis) are abstractions taken from
the real world. He demonstrates this for his concepts of number.
As a consequence of this, his basic notions are of utmost generality: “set”, “unit”,
“exact parts of the unit”. Additionally, an impressive testimony to this view is the
fact that the negative numbers are mere conceptual constructions, not abstractions
from reality. In plain language: for Weierstraß, negative numbers do not exist.
Cantor does the opposite. Right from the start, he implements very abstract math-
ematical notions: “order”, “sign” and finally the tricky concept of “convergence”!
While Weierstraß toils with philosophical abstractions based on reality, Cantor
performs conceptual acrobatics right from the start.
Cantor thinks of his analysis (a) as without any relation to reality, (b) in mere technical
mathematical concepts.

Cantor conceives analysis purely formally—more precisely: analysis as a whole.


This arises from the fact that his basic concept (converging sequence) is a formal
one.
It is not Cantor, but Heine, who goes one step further and provides a purely
formal ontology for his concepts. Heine, not Cantor, takes the numbers to be nothing
more than signs [Zeichen]. Heine writes:
Requirement: To add to each number series a sign.
One introduces the series itself as a sign, put in brackets, such that e.g. the sign belonging
to the series a, b, c, d, etc. is [a, b, c, d, etc.].

In addition, in § 2 of his article, he puts:


1. Definition. General number or number-sign [Zahlzeichen] is the name of the sign
belonging to the series of numbers.
2. Definition. Number-signs are called equal or are interchangeable, if they belong to equal
series of numbers, and not equal or not interchangeable, if they belong to different series of
numbers.

Heine declares in crystal-clear words “numbers” to be “signs”.


Georg Cantor’s Construction of the Real Numbers 193

Cantor’s contemporary Gottlob Frege (1848–1925), who is a reliable philosoph-


ical nit-picker, sharply contradicts him in 1893:
We see that the number-signs have an entirely different importance from that which was
attributed to them before the rise of formal theories. They are no longer mere external
auxiliaries like blackboards and chalk; rather, they comprise the essential components of
the theory itself.

On the other hand, Cantor argues against this plain formalism Heine was devoted
to (see on p. 191).
Cantor defended himself against philosophical criticism like Frege’s, in 1889, as
follows:
I never claimed that the signs b, b , b , . . . were, strictly speaking, concrete quantities. As
abstract objects of thought they are merely quantities in a figurative sense. The decisive
feature is that with the help of these abstract quantities b, b , b , . . . it is possible to
determine real, concrete quantities exactly.

To this, Frege had a clear answer (which is also a criticism of Dedekind, see
p. 202):
It demands a strong faith indeed, to take signs that are written with chalk on a blackboard
or with ink on a paper, that can be seen with the bodily eye, as abstract objects of thought.
It is the kind of faith which moves mountains and creates the irrational numbers.

The philosophical mind Frege insists on the basic question to the new formal
mathematics of Cantor, Heine, Dedekind (he will soon have his say), their disciples,
and followers:

Why have arithmetical equations a practical application? Solely, because they express
thoughts. How could an equation expressing nothing, which was no more than a group of
figures that could be transformed, according to certain rules, into another group of figures,
be applied? Now, it is applicability alone, which elevates arithmetic from being a mere game
to the rank of a science. Thus, applicability necessarily belongs to it. Is it advisable then, to
exclude from arithmetic what it requires to become a science?

At the same time, Frege, with his notion of “applicability” offered the mathe-
maticians a comfortable excuse: they could confine themselves to “pure” theory—
applications were the job of the practical men, not theirs.
This was not a smart move by Frege. It is not the excuse of a division of labour
between “pure” and “practical” mathematics that is at stake. Mathematics is a whole
(which, of course, has many fields—this is one of the viewpoints of this book). At
stake is the reference to reality of mathematics, those components of reality that
allow the specialists to model our world thereon, as Weierstraß had been still aware
of—we shall continue this line of thought at the end of this section, p. 198.
194 13 Analysis’ Detachment from Reality—and the Introduction of the Actual Infinite. . .

The Mathematical Price of This Progress

The concept of number that Cantor invented has two important mathematical
consequences.
1. It is generous to such a degree that it allows for the validity of all conventional
arithmetical methods.

We remember: Weierstraß judged some of the generally accepted ways of


calculating to be “fallacies” (p. 177).
Oh, well: Cantor does not really allow all the calculations, namely not those
from Johann Bernoulli and Euler, which embrace infinitely large or infinitely small
“numbers” (see the example on p. 76). Because:
2. The infinitely small “numbers” were excluded from analysis (and therefore also the
infinitely large ones).

The latter is the result of the following reasoning: Cantor learnt from his
teacher Weierstraß that the definition of a new notion of number also requires
the definition of the basic arithmetical concepts “equal”, “addition”, “subtraction”,
“multiplication” etc. in regard to the new objects.
The operations are simple. In the case of two converging sequences of rational
numbers

a1 , a2 , a3 , a4 , ... ,
a1 , a2 , a3 , a4 , ... ,

one operates as follows to get

a1  a1 , a2  a2 , a3  a3 , a4  a4 , ...

( is one of the four basic arithmetical operations; of course, division by zero is


forbidden) a converging sequence of rational numbers. For the “definite limits” b
and b , as well as for the result of the calculation—which is the “limit” b of the last
sequence—this means:

b  b = b .

So far, so good.
But what about the last “=”? Under which circumstances shall we take two
converging sequences of rational numbers to be “equal”?
Cantor argues as follows:
The two sequences a1 , a2 , a3 , . . . and a1 , a2 , a3 , . . . always relate to each other in one
of the following three ways, each of which excludes the others: either 1. an − an becomes
infinitely small for increasing n or 2., starting from a certain n, an − an remains always
larger than some positive (rational) quantity ε, or 3., starting from a certain n, an − an
remains always smaller than some negative (rational) quantity −ε.
Georg Cantor’s Construction of the Real Numbers 195

In the first case, I establish

b = b ,

in case of the second b > b , in case of the third b < b .

In other words, Cantor identifies, for example, the two new numbers

1, 12 , 14 , 18 , . . . and 0, 0, 0, 0, . . .

His argument: both numbers 1, 12 , 14 , 18 , . . . and 0, 0, 0, 0, . . . have the same


limit—and “limit” is just the name Cantor has given to his new numbers b, b , etc.
Let us postpone the discussion as to whether this is really clear for a moment.
Instead, we stick to Cantor’s way of doing things. We realize:

Cantor identifies the classical “infinitely small” variable 1, 12 , 14 , 18 , . . . with the number 0.

This and only this makes the “infinitely small” quantities in the late nineteenth and the
twentieth century disappear: because of a pure definition—and not because of Weierstraß
and his intricate epsilontics!

As we have shown: Cantor did this. However: is he actually allowed to do this?


Is such a definition of “equality” really permitted? It seems, after all, that the two
numbers taken to be “equal” are different!

Notoriously, Mathematicians Call Different Things “Equal”

The logician Frege opposed this kind of sloppiness—seen in Cantor’s reasoning—


quite firmly in his later years. Frege writes:
With respect to the equality sign, we will do well to abide by our stipulation, according to
which equality is complete coincidence, identity.
To be sure, bodies of equal volume are not identical; however, they have the same
volume. In such a case, the signs on each side of the equality sign are not be taken as
signs for bodies but for their volumes, or alternatively for the measuring numbers resulting
by measurement in terms of the same unit of volume. We will not be speaking of vectors
as equal but of a particular specification—let us call it “directed length”—for these vectors,
which may be the same for different vectors.
According to this conception the progress of science will not require an extension of the
formula a = b; instead there are only new determinations (modi) of the objects which are
to be considered.

Two decades earlier Frege’s view had been this:


Whether we say, like Leibniz, “the same” or “equal” does not matter. “The same” looks like
a complete coinciding, whereas “equal” only expresses a coinciding in this or that respect;
however, one can introduce such a manner of speech that this difference vanishes. For
example, one says “the lengths of these straight lines are equal” instead of “these straight
lines are equal in length” [. . . ] In this manner we used the word in the examples above.
Indeed, all laws of equality are covered by the general replaceability.
196 13 Analysis’ Detachment from Reality—and the Introduction of the Actual Infinite. . .

(Weierstraß, too, took replaceability to be the essence of equality.)


In consequence, following Frege’s first opinion by calling only identical things
“equal”, no mathematics will result, but only logic. Mathematics usually uses
“equal” in the sense given directly above. However, it is this “equal (in a certain
respect)” that is often not understood by non-mathematicians, for the mathemati-
cians usually do not EXACTLY say what they mean.
1. x 0 = 2 means: “The value of the indefinite x0 is 2”. So the “indefinite” is going to be
“determined”!?—Well, is this permitted? Is it, actually, permitted to “determine” an
“indefinite”? Why is this permitted—as it is then no longer “indefinite”?—Besides, it
is not clear whether this is an assumption or a solution (to a question).
2. x 0 = X means: “x0 and X are the same value (without saying, which value it is)”.
Here two different “indefinite” values are determined “in the same sense” —without
specifying, what this “determination” is going to be.
3. f (x) = x 2 means: “f is the quadratic function”. This is an assumption or a solution.
“f ” is the name of a function.
4. lim x1 = ±∞ means: “It can be proved that the function x1 has the two ‘limits’ +∞
x→0
and −∞, if x → 0”.
5. etc.

In mathematics, equality has very different meanings. It is only this unspoken


knowledge that enables mathematics.
Mathematics is a speaking, a writing.
The procedure of speaking or writing is inseparably connected to mathematics—
which are both part of making mathematics a creative act. However, normally, this
is not spoken of.
Weierstraß proved the following:
Theorem. If a and b are fractions, a = b implies b = a.

This “equality” results from a “transformation”. However, this is not always the
case. Another “equality” may result from an “identification” (not taken in its logical,
but in its mathematical sense: identical in a certain respect).

An Unexplored Mathematical Potential (Of Cantor and His


Contemporaries)

Let us return to the question asked on p. 195:


Is Cantor allowed to identify any two of his new numbers when they differ by an infinitely
small quantity (to use the terminology of that time)?

Weierstraß would have demanded a justification for this definition. (This might
be the reason for Weierstraß to abstain from commenting on Cantor’s way of
constructing the concept of the real numbers.) By contrast, Cantor did not see any
necessity for this. Instead, he claims: “The essence of mathematics is its freedom”.
Georg Cantor’s Construction of the Real Numbers 197

There is no harm in mathematicians reflecting on their doing. Let us make up for


this and add at least some thoughts.
Of course, the mathematician is permitted to fix a definition—whenever it makes
sense, it is free from contradictions and is productive. It is beyond all doubt
that Cantor’s definitions fulfil these requirements. Nevertheless, we may add our
reflections. The real question, of course, is: must Cantor give these definitions?
To pose this question is almost a denial of it. As nearly always in life, alternatives
do exist. However, Cantor discredited them a priori by choosing the name “limit”;
remaining more open, he could have spoken of a “(new) number” or something
similar.
In taking Heine’s formal attitude seriously, one may arrive at the following idea:
Alternative definition. Two converging sequences of rational numbers

a1 , a2 , a3 , a4 , ... ,
b1 , b2 , b3 , b4 , ...

(with the same limit) are called “equal” only if they are identical. The equality of two
corresponding terms an and bn is established by the usual arithmetical equality.

That means: not only 1, 12 , 14 , 18 , . . . and 0, 0, 0, 0, . . . are “different”, but


both are also different from 1, 0, 14 , 18 , . . . —Nevertheless, 1, 12 , 14 , 18 , . . . and
2 2 2 2
2 , 4 , 8 , 16 , . . . are “equal”.
It is even possible to go further, as Detlef Laugwitz showed in the 1970s. One
can say: in analysis the limits do not depend on a finite number of different pairs
of corresponding terms. As we are talking about the convergence of sequences of
rational numbers, the above definition can be made weaker:
Alternative definition II. Two converging sequences of rational numbers

a1 , a2 , a3 , a4 , ... ,
b1 , b2 , b3 , b4 , ...

(with the same limit) are called “equal” only if they are, with the exception of only a
finite number of terms, identical. The comparison of two corresponding terms an and bn
is established by the usual arithmetical equality.

Obviously, these two changes in Cantor’s definition of equality have important


consequences:

1. Now there exist “infinitely small” numbers of a new kind. Take, e.g.
1, 12 , 14 , 18 , . . . It differs from 0, 0, 0, 0, . . . (and is even greater than
it).
2. It is not always possible to compare two new numbers, e.g. 1, 0, − 13 , 0, 15 , 0,
− 17 , 0, . . . and 0, 0, 0, 0, . . .
3. Sometimes, division is problematical.
198 13 Analysis’ Detachment from Reality—and the Introduction of the Actual Infinite. . .

Some properties can be salvaged, if more numbers of this new kind are accepted.
(a) Besides the converging sequences of rational numbers, one should also
accept the sequences of their reciprocals (which are the “definitely” diverging
sequences of rational numbers). The reason is this: analysis demands the
division of numbers. Therefore, if one accepts “infinitely small” numbers, the
arithmetical laws demand “infinitely large” numbers, too. If a = 0 is a number,
so must be a1 .
(b) Nevertheless, there are shortcomings that cannot be overcome: no number of
this new kind, which is a sequence containing zero as a term, can be a divisor
(at least in case of there being infinitely many zeros). In other words:
This new set of numbers does not come with a general concept of division.

It is a question of personal taste whether or not one finds this acceptable. (There
is an analogous state of affairs in regard to Weierstraß’ numbers—however, they
can be divided by a particular procedure! Maybe there also exists a procedure
to divide by “infinitely small” numbers for this new kind of number?)
This should suffice. It was our aim to explore the possibilities of Cantor and his
colleagues defining new numbers in another way from what they actually did. Of
course, we have to concede that they did not see these possibilities. It was not until
the second half of the twentieth century that such ideas have emerged.

The Formal Character of Cantor’s Analysis

Frege’s criticism (p. 193) made it clear: Cantor’s analysis bears a formal character—
although not to the same degree as Heine’s. The reason is: Cantor is unable to
declare WHAT his numbers b, b , . . . are. To Frege, Cantor replies: they are things
of thought. However, in 1872, Cantor was unable to specify, of what particular
kind these “things of thought” are. Whereas his teacher Weierstraß tried to create
mathematical concepts fit for real applications, Cantor was content with “things of
thought”. Not Cantor, but Weierstraß justified the introduction of each new concept
of number with recourse to real activities .
It took another 23 years before Cantor succeeded in proposing a first draft of
what is called “set-theory” nowadays. The first sentence of his treatise from 1895 is
widely known:
By “set” (in German: “Menge”) we mean any collection M of definite, clearly distinguished
objects m of our intuition or our thinking (which will be called “elements” of M) of a whole.
It will be symbolized thus:

M = {m} .
The Form of the Real Numbers, Invented by Traugott Müller, Joseph Bertrand,. . . 199

There had been earlier attempts, but not until Cantor’s ideas did they lead to
a system. During the twentieth century, these ideas developed gradually into “set-
theory”, or more exactly: into different set-theories.
Set-theory is a theory of pure thought, without any chains to reality (and even without any
postulation thereof).

Ever since Cantor published his first sketch of a Set-theory, the real numbers have
been constructed within a set-theory. (Other attempts did not gain wide recognition.)
That is why it is legitimate to characterize Cantor’s concept of number as a “formal”
one—a concept that does not want to be related to anything approaching reality.
This is a decisive change within the philosophical style of analysis. This change
happened between Weierstraß and Cantor, and it was driven further by the creation
and the implementation of Topology into the field of analysis, which is still the case
today.

The Form of the Real Numbers, Invented by Traugott Müller,


Joseph Bertrand, and Richard Dedekind

The Basic Problem

On 20 March 1872, Richard Dedekind (1831–1916) wrote the following:


The statement is so frequently made that the differential calculus deals with continuous
magnitude, and yet an explanation of this continuity is nowhere given. [. . . ] To secure a real
definition of the essence of continuity I succeeded on 24 November 1858. [. . . ] I could not
make up my mind to its publication, because the theory itself had little promise.

However, on March 14 and 20, the two treatises by Heine and Cantor with their
construction of the real numbers came into Dedekind’s hands, and this made him
publish his own ideas in a commemorative that appeared on 26 April 1872.
Subsequently, the text was written in a hurry. This is readily apparent. Later,
Bertrand Russell presented the idea much more elegantly.
Even his choice of words turned out to be somewhat clumsy. He called “stetig”
(“continuous”) what is normally called in German “kontinuierlich” (in English they
are the same! see p. 108).

The Basic Idea

Clear-sightedly and from a historical distance: what was the actual problem? What
was really and “securely” known about the “numbers” in 1872?
200 13 Analysis’ Detachment from Reality—and the Introduction of the Actual Infinite. . .

Well, with regard to analysis, nobody will doubt the nature of the “natural”
numbers 1, 2, 3, . . . Similarly with the “fractions” ( 12 , 13 , 23 , 37 , . . . ) that had been
used in trade and commerce for centuries.
√Really
√ problematic were (and are!) the other “numbers”: “negatives”, “roots”
( 2, 3 5, . . . ), or even more extravagant numbers like, for instance pi (π ) etc. What
was known about them?
At least, this: they can be indefinitely approximated by the unproblematical
numbers (the fractions). The problematical numbers are placed, somewhat cosily,
in between the unproblematical ones. We have already looked at this (p. 168).
Consequently, all things considered, the following idea will appear not too far-
fetched:
Let us classify the fractions according to whether they are smaller or greater than such a
problematic number!

The most important thing with this idea, as with any good idea is: one must
HAVE it! Nothing else can get off the ground without this flash of genius. However,
for those who succeed, everything will follow.
The idea is this: define such a problematic number by classifying the known
numbers—into whether they are smaller or whether they are greater than the one
under consideration. Isn’t this idea much easier than Cantor’s—isn’t it closer to
hand?
Yes, it is! This is proved by the fact that it came up much earlier than Cantor’s—
for: it is a fact, Richard Dedekind was not the first who thought of it.

Traugott Müller, the Very First Man Who Invented the


“Dedekind-Cuts”, in 1838
A Mathematical Outsider

Traugott Müller (1797–1862) was a very gifted mathematician. His father was
a priest, and the son studied first theology and then mathematics and science
in Leipzig between 1817–22. In the early years, when Müller was gradually
introduced in traditional German grammar schools (called “Gymnasien”)—in
which, originally, only languages were taught—, Müller worked as the headmaster
at the Real-Gymnasien of Gotha (1836–44) and Wiesbaden (1844–62). Besides
other books, he published a Textbook of General Arithmetic for grammar schools,
comprising 555 pages, in 1838; a second edition appeared in 1855.

Müller’s Educational Convictions

Müller’s aim was to make Arithmetic as important as Geometry in education. Both


fields “should keep the mind engaged to an equal extent”. This aim led Müller
to write his highly systematized textbook of Arithmetic, which he started with a
systematic classification of the concept of number.
The Form of the Real Numbers, Invented by Traugott Müller, Joseph Bertrand,. . . 201

Müller’s Concept of Irrational Number

Having introduced the fractions (without sign), Müller presented the classical proof
for the existence of “incommensurable” quantities. From this he concluded:
Since, for two incommensurable quantities A and B, we always have

p A p+1
q < B < q .

A
Whatever the integer q might be, we are not able to represent the quotient B exactly,
neither by an integer nor by a fraction. However, we can imagine q as large as we wish
and, consequently, make q1 smaller than any given number. That is why it is possible to
p p+1
approximate this quotient A
B endlessly, whereby A
B − q as well as q − A
B is always less
than q1 . This makes A
B the limit of these two fractions, which will change for any new q.

And then follows a breathtaking (somewhat long-winded) sentence:


If we would be able to show that these limits, if they were joined by the usual operation
signs, can be transformed in the same way as were proved for the integers and later for the
fractions, then it would be legitimate to give them the name number with the same good
reasons as was done with fractions, thereby raising the former theorems to more generality.
This will be done in the following paragraphs.

Müller continues:
Unlike the former numbers, these limits will be called irrational numbers, whereas the
former ones are named rational numbers.

Next, Müller proves the equality of his irrational numbers to be the identity and
shows the validity of the commutative law for addition and multiplication for them.
However, Müller’s definition of the relation < is wrong.
Nevertheless, we can state:

(a) In 1838, Traugott Müller formulated the construction, which today bears the name of
Dedekind (“Dedekind-cut”), in absolute clarity.
(b) Müller took this idea to define a new kind of number, beyond the rational numbers.

An Incidental Idea, by Bertrand in 1849

Eleven years later, a textbook on arithmetics for scientists was published in Paris.
Its author was Joseph Louis François Bertrand (1822–1900), who was, like Müller,
a grammar schoolteacher but became in 1856 a university professor. It is almost
certain that he was unaware of Müller’s book, but, amazingly, he also formulated
Müller’s idea:
An irrational number cannot be defined by means of a unity, like a quantity. That is why we
take for its definition the specifications, which rational numbers are smaller and which are
greater than it . . .
202 13 Analysis’ Detachment from Reality—and the Introduction of the Actual Infinite. . .

. . . which is the irrational number. Quite casually and without pretentiousness, all
necessary aspects are stated here.
The logician might object: how can something known to be “smaller” or
“greater” than something else, which is unknown? However, this is nit-picking.
Bertrand presents the idea very clearly: he divides the “rational” numbers into two
classes, the class of the “smaller” and that of the “larger” ones.
Spelling out all the details remained a task for Dedekind 13 years later. But even
so, the idea was formulated.
Bertrand continued in this casual style. Addition is left out—what could possibly
not be clear about it? However, multiplication is treated more explicitly:
If the multiplicator is irrational, we need a new definition [of multiplication]. We call the
product of a [rational] number A with an irrational number B a number, which is smaller
than the product of A with any arbitrarily chosen rational number, which is greater than B
and which is greater than the product of A with any arbitrarily chosen rational number,
which is smaller than B.

The product of a rational number r with an irrational number q cannot be


rational. And how shall we determine this irrational product? Clearly, by specifying
the “smaller” and the “greater” rational numbers. All this is stated by Bertrand.
All well explained. However, Dedekind presents just the same, but with more
pomposity.

The Dramatic Version, Given by Dedekind in 1872

Richard Dedekind gives a much more dramatic presentation of this simple idea—
even without considering his geometrical preliminaries.
First, Dedekind defines the concept of “cut” (which nowadays bears his name;
by “system R”, Dedekind means all the rational numbers):
If there is given any separation of the system R of rational numbers into two classes such
that every number a1 of the first class A1 is less than every number a2 of the second class
A2 [. . . ], then for brevity we shall call such a separation a cut and designate it by (A1 , A2 ).

This is the crystal-clear definition, which was also given, and in nearly the
same precision, by Traugott Müller in 1838, and, somewhat more incidentally, in
1849, by François Bertrand.—What follows is Dedekind’s description of such a
problematical number:
Whenever, then, we have to do with a cut (A1 , A2 ) produced by no rational number,
we create a new, irrational, number α, which we regard as completely defined by this
cut (A1 , A2 ). From now on, and for the sake of brevity, we shall say that the number α
corresponds to that cut or that it creates this cut.

What Müller declares to be a “legitimate” number and what Bertrand plainly


defines to be a number is highly stylized by Dedekind to be a “creation”. It’s the
tone that makes the music. Frege’s response we do already know (p. 193).
The Form of the Real Numbers, Invented by Traugott Müller, Joseph Bertrand,. . . 203

The arithmetical operations are given by Dedekind this way:


If c is any rational number, we put it into the class C1 , provided there are two numbers one
a1 in A1 and one b1 in B1 such that their sum a1 + b1  c; all other rational numbers shall
be put into the class C2 .
This separation of all rational numbers into the two classes C1 , C2 evidently forms a
cut, since every number c1 in C1 is less than every number c2 in C2 .

These are, yet again, the same ideas that Müller and Bertrand gave but garnished
with a lot of pomp and in much more detail.

The Elegant Version, Given by Russell, in 1919

Half a century later, Russell (1872–1970) put things in a rather elegant and simple
way. (The Nobel laureate for the literature, Bertrand Russell, wrote this book during
his six month imprisonment for his militant anti-militarism.) Felix Hausdorff (1868–
1942) got the same idea earlier but did not explicate it with the same elegance.
(i) Russell started with the concept of dividing the rational numbers into two
classes, where the class of the smaller numbers has no maximum and the class of
the bigger one no minimum. (ii) Then he defined the concept of “limit” (to be “of
utmost importance”): the “upper limit” in the case that the class of smaller numbers
has a maximum and the “lower limit” in the opposite case. Thereafter, he proceeded
as follows (emphasized words added):
Let us observe that an irrational is represented by an irrational cut and a cut is represented
by its lower section.
Definition (segment). Let us confine ourselves to cuts in which the lower section has no
maximum; in this case we will call the lower section a “segment”.
Then those segments that correspond to ratios are those that consist of all ratios less
than the ratio they correspond to, which is their boundary[2] , while those that represent
irrationals are those that have no boundary.
Segments, both those that have boundaries and those that do not, are such that, of any
two pertaining to one series[3] , one must be part of the other; hence, they can all be arranged
in a series by the relation of whole and part, that is the relation .
A series in which there are Dedekind gaps, i.e., in which there are segments that have
no boundary, will give rise to more segments than it has terms, since each term will define
a segment having that term for boundary, and then the segments without boundaries will be
extra.
We are now in a position to define a real number and an irrational number.
Definition (real number). A “real number” is a segment of the series of ratios in order of
magnitude.
Definition (irrational number). An “irrational number” is a segment of the series of
ratios which has no boundary.
Definition (rational real number). A “rational real number” is a segment of the series of
ratios which has a boundary.

2 emphasis added
3 The modern term is: “sequence”.
204 13 Analysis’ Detachment from Reality—and the Introduction of the Actual Infinite. . .

Thus a rational real number consists of all ratios less than a certain ratio, and it is the
rational real number corresponding to that ratio. The real number 1, for instance, is the class
of all proper fractions [all negatives included].

Russell simplifies Dedekind’s concepts significantly:


(i) Russell realizes: Dedekind does not really work with two classes A1 , A2 of
rational numbers, but only with one: A1 ; for A2 is defined by A1 : A2 = R \ A1
(the set-theoretical difference).
(ii) Whether A1 has a boundary or not is important. (This boundary may belong to
A1 or to R \ A1 .)
(iii) Consequently, Dedekind’s use of the ordered pair (A1 , A2 ) (at least: in notation)
is empty talk. Actually, Dedekind does not work with pairs (ordered or not), but
only with single sets of rational numbers (of a certain kind).

This is the fourth construction of the “real” numbers, which makes use of the
rational numbers or fractions. That this idea was published three times, probably
independently of each other, in the middle of the nineteenth century, seems to show
at least two facts:
(a) Textbooks for schools were not read by scientists in those times.
(b) Around 1872, German mathematicians were not used to reading French text
books—which was obviously an omission.
However, the first point might not be completely correct. For Cantor cited a
schoolbook in 1883! And even more surprisingly: he cited Müller’s textbook for
grammar schools in its second edition! He did this in his introduction, where he
mentioned the rational numbers, in which Cantor continued to discuss the concepts
of real numbers (“which are known to me”), given by Weierstraß, Dedekind, and
himself.
We know already that Cantor misunderstood Weierstraß’ concept, but what about
Müller’s? The second edition of Müller’s book also contains his concept of real
number, which at bottom is the same as Dedekind’s. This is true, although in the
second edition it is not as clearly presented as in the first edition. Now: did Cantor
not understand Müller’s definition of the real numbers? (If so, this would strengthen
the suspicion that mathematicians are not reliable readers of alien ideas.) Or did he
not want to discredit Dedekind’s fame of being the (first and sole) inventor of this
concept?

The Evaluation of This Solution by Tannery in 1886

Because nobody knew about Weierstraß’ special concept of real number in the
nineteenth (and twentieth!) century (there may have been one single exception: Otto
Hölder (1859–1937)), only two of them became public knowledge. But which of
them was thought to be the proper one for the foundations of analysis?
In 1886, Jules Tannery (1848–1910) published his textbook Introduction to the
Theory of Functions in One Variable, which became influential in the following
The Axiomatic Characterization of the Real Numbers by David Hilbert in the. . . 205

years. Therein he defines the real numbers with reference to Bertrand. Later, Tannery
explained that he had not had access to Dedekind’s treatise from 1872, but he had
known about Heine’ detailed treatise, that is: Cantor’s concept.
Tannery’s evaluation is of interest: he valued the constructions of Cantor and
Heine to be “too arbitrary”, and therefore he did not accept them. A very interesting
judgement! According to Tannery, the construction of the real numbers as “cuts”
within the class of the rational numbers is the more natural way, free of arbitrariness.
Tannery accepted the fact that the operations using “cuts” are far less elegant than
those with sequences—not to speak of the proofs of the arithmetical laws! Today
we may guess: if Tannery had known about Weierstraß’ construction of the real
numbers, he might have preferred it to the others because of the better coherence of
the formation of its concepts, the simplicity of its means, and the more general
way of defining addition and multiplication. Instead, there might have arisen a
fruitful discussion about the real differences between direct and indirect arithmetical
operations. Perhaps, analysis would have taken on a different shape?

The Axiomatic Characterization of the Real Numbers by


David Hilbert in the Years 1899 and 1900

The Situation in 1872: Two Definitions of One Subject

Müller as well as Bertrand and Dedekind on the one side, and Cantor on the
other, had published two constructions of the real numbers. Although they differ
significantly from each other, both constructions were based on the “fractions”.
Why do both constructions lead to the same result? Moreover, do they actually produce the
same result? And if so, what exactly is the meaning of “the same”?

So far as I am aware, this question was not formulated in print before Hilbert’s
Axioms. Clearly, the fact is not evident and a proof for this sameness is far from
evident. To my knowledge, no mathematician had any doubts then. The fundamental
belief in the unity of mathematics might have been the underlying sentiment ruling
out uncertainty.

Hilbert’s Axiomatic System—the First Attempt, 1899

Without having been posed beforehand, the problem was solved 27 years later by
David Hilbert (1862–1943). He presented a characterization of the “real” numbers.
Hilbert approached the problem differently from earlier mathematicians. He
argued for a logical perspective. As for Geometry, Hilbert defined the task for
mathematics to offer “a logical analysis of our spatial intuition”. In the process
206 13 Analysis’ Detachment from Reality—and the Introduction of the Actual Infinite. . .

of this, he presented in 1899, quite casually, a logical examination of the analytical


concept of number. Hilbert wrote:
In the beginning of this chapter we are going to give a short discussion of complex number
systems, which will be helpful later on in the presentation.
The real numbers in their entirety are a system of things having the following properties:
...

Hilbert’s wording “complex number system” shows the influence of Weierstraß.


However, what follows is completely alien to Weierstraß’ way of thought:
twelve “theorems of operation”

(like: a · 1 = a or a(bc) = (ab)c)


as well as four “theorems of order”

(e.g.: if a > b, then we have a + c > b + c)


and the “Theorem of Archimedes”.

That was all!


Or was it not, after all? Immediately after his book was published, Hilbert
realized two mistakes. First, he omitted a very important “property”—we shall come
to this soon. Secondly, in his first characterization of the system of real numbers,
Hilbert called the properties under discussion “theorems”. However, a “theorem” in
mathematics is a proposition that must be proved. Nonetheless, Hilbert is unable to
prove these twelve “theorems”! How could he? Hilbert has nothing at all at hand for
what he may prove “properties” thereof! Quite contrarily, Hilbert takes the opposite
way. Hilbert states some “properties”—and then proclaims: I shall call all objects
that bear these properties to be the “system of real numbers”. Without any argument:
there you go! Par ordre du mufti.
At the first attempt, Hilbert made two mistakes: a methodological one (his talking
about “theorems”) as well as one of skill (the omission of a central property). Even
the great mathematician David Hilbert was not infallible.

The Axiomatic System of Hilbert—the Second Attempt in 1900

One year on, Hilbert published an article, wherein he removes those defects (without
hinting at these changes).
In the new paper, Hilbert repeats his theorems word for word—but calls them
“axioms”. Only their order is, in two cases, changed. Moreover, Hilbert changes
and supplements titles. Now there are six “axioms of operation”, six “axioms of
calculation”, as well as four “axioms of order”. Also, there are two “axioms of
continuity” introduced: as the last “theorem” (“Theorem of Archimedes”, now:
“Axiom of Archimedes”), and as eighteenth axiom (emphases are added):
Axiom of Completeness. It is not possible to add to the system of numbers another system of
objects such that the preceding axioms are fulfilled for the resulting new system. In short:
The Axiomatic Characterization of the Real Numbers by David Hilbert in the. . . 207

the numbers constitute a system of objects which is not capable of any enlargement, if all
the axioms hold.

This decides the question that divided the two collaborating colleagues Cantor
and Heine and in which Dedekind sided with Heine’s position by a pure postulation.
Hilbert decided that Dedekind and Heine were right and that Cantor was wrong.
The problem was the following. Cantor as well as Heine generated from
the system of rational numbers a new system of numbers (today called: “real”
numbers)—although with different methods.
In both cases, the question arose: what happens, if the new system, i.e. the real
numbers, is taken as a starting point, instead of the “rational” numbers, and if we
then generate in the same manner yet another system of new numbers?
Heine formulated and proved a theorem: the numbers that are subsequently
generated “are not new, but coincide with those which were generated at first”. (That
this theorem contradicts Heine’s opinion that numbers were signs, as a matter of
course, went unnoticed.)
Ironically, Cantor disagreed. It is true, he conceded that the two newly generated
“systems of number to some extend coincide”, but he wanted “to cling to the
conceptual differences of the two systems”. In this case, Cantor’s manner of forming
concepts was very pedantic.
Dedekind, on the other side, felt himself unable “to yet detect the advantage of
this mere [sic] conceptual difference” (emphases added).
Now Hilbert decrees: this conceptual difference, which was “only” noticed by
Cantor, is, when considered purely mathematically, insignificant. Basta!
Today, the two “axioms of continuity” are replaced by a single one, which
is designed very differently: the “Axiom of Completeness”. The “Theorem of
Archimedes” is then proved (with the help of the unlimitedness of the natural
numbers).

Advantage and Disadvantage of Hilbert’s Axiomatic System

With his axioms, Hilbert distilled out those properties of the “real” numbers that
have been generally accepted ever since. (Maybe not completely apt, as it seems
today.) This was the first decisive step on the path to structural mathematics,
which dominated analysis world-wide after the Second World War, and which was
disseminated by the fictitious author called “Nicolas Bourbaki”.
The power of this perspective is obvious: the important properties (of the number
system) are clearly expressed. This helps to study the variants of the number system
and their special attributes. The same with other structures.
However, Weierstraß’ construction of the real numbers proves that Hilbert’s
method is not universally applicable! Weierstraß’ system has no concept of
division—nonetheless, it is possible to divide these numbers! Weierstraß’
construction unmistakably proves: an indirect operation is fundamentally different
208 13 Analysis’ Detachment from Reality—and the Introduction of the Actual Infinite. . .

from a direct one. If you want to divide two numbers, you need a common system
to represent them (e.g. the decimal system)—otherwise you will not be able to
describe a procedure formally (which is one of trial and error!) in order to execute
the calculation! That is why Weierstraß’ construction has proved that:
There exists a version of the real numbers, which does not comply with Hilbert’s axiomatic
system.

This fact seems to narrow the scope of Hilbert’s axiomatical method and to point
to a system missed out by structural mathematics. Maybe “pure” mathematics in the
footsteps of Hilbert, that is to say: structural mathematics, is not at all as general
as it is claimed to be? Is there some pure mathematics which is not structural
mathematics? And if so, what is it?
Hilbert did not stop at just listing these advantages. He went on to demand: the
“real” numbers must be defined in this axiomatic way.
That’s a bit much! Immediately, Gottlob Frege, whom we already know as a
harsh philosophical critic, entered the scene. In a short correspondence between
27 December 1899 and 22 September 1900, both exchanged their opinions without
reaching any consensus.
Frege asked Hilbert about the justifications for his axioms. Taken from a
traditional philosophical point of view (which we set out to call “substancial”),
axioms must express “the basic facts of intuition”, and therefore they are in need
of a legitimation.
Hilbert firmly rejected such a demand (and thereby, as a consequence, the
traditional philosophical view). Hilbert said: I merely baptize certain properties
(Hilbert calls them “characteristics”) to be “axioms”. This was merely a “matter of
taste”. In his opinion, a concept cannot be derived from a thing, but concepts could
only be specified by “their interrelatedness”—as it is done by an axiomatic system.
On December 29th , 1899, Hilbert explained this by an illustrative formulation:
If I take my points to be any system of things, e.g. the system: love, law, chimney-sweeps,
and if I only assume all my axioms to be the relations between these things, my theorems
would be valid for these things, e.g. the Theorem of Pythagoras.

Frege opposed this and objected to it (thereby referring to different kinds of


geometry; in some of them the Theorem of Pythagoras is valid, and in others it
is not, e.g. in the spherical geometry):
Only the wording of the theorem is the same; the mental content (Gedankeninhalt) is in
each geometry another. It would not be correct to call a special case of the Theorem of
Pythagoras “the Pythagoras”; because if we have proven it in a special case, we have not
proven the Pythagoras.

The conception of mathematics, propagated by Hilbert against that of Frege, is


called by me “relational”. It is opposed to the “substancial” conception, which was
indicated by Weierstraß. Therefore, we should not be surprised that Hilbert’s method
of judgement does not apply to Weierstraß’ concepts. It is really a pity that Frege
did not know about Weierstraß’ concept of numbers—because he could seriously
have driven him into a corner.
The Price of Success: The Inclusion of the Actual Infinite in Mathematics 209

The New Social Duty of Mathematics—Concerning Hilbert’s


Ideas

In a public speech, given in Zürich in 1917 during the Great War, Hilbert resumed
this topic (emphases original):
If we compile the facts of a certain, more or less extensive field of knowledge, we easily
realize that these facts are capable of being arranged. This arrangement always takes place
with the help of a certain framework of concepts, such that each single object corresponds
to a concept of this framework and each single fact of this field corresponds to a logical
relation between these concepts. This framework of concepts is nothing else than the theory
of this field of knowledge.

This way, Hilbert relocates traditional mathematics in a fundamentally new way


in our society: the subjects of mathematics are no longer, as before, the facts (of
thinking), but it is the ordering of any facts within any “field of knowledge”,
whatsoever.
In doing so, he elevated mathematics, towards the end of the great upheaval of civilization,
the Great War, to the general security police of science. From now on, mathematics did
not have to deal with “things” and their “relations”, but with the appropriate “framework
of concepts”: for the logical structuring of “every fact in a given field of knowledge”. This
dismisses traditional mathematics as a science of things (concepts).

By the way, it was Dedekind who had already demanded this basic change within
mathematics more than half a century earlier. As early as 1854, Dedekind stated in
his habilitation lecture:
This twisting and turning of the definitions for the sake of the found laws or truths, in which
they play a role, constitutes the highest skill of the systematist.

Dedekind did not say why “truths” are “laws”. He also concealed from where
these “laws” or “truths” should derive their authority. It seems clear that Dedekind
takes the “concepts” to make up that “framework” talked about by Hilbert in his
letter to Frege in 1899 and in his public speech in 1917.

The Price of Success: The Inclusion of the Actual Infinite in


Mathematics

The Plain Fact

By the year 1872, following Dedekind, the idea of defining the “irrational” numbers
with the help of “cuts”, which are divisions of the ENTIRETY of the rational
numbers, was increasingly accepted in mathematics. This had not been the case
in Bertrand’s time, in 1849, and even less so in Müller’s time, in 1838.
210 13 Analysis’ Detachment from Reality—and the Introduction of the Actual Infinite. . .

A “cut” divides the entirety R of the “rational” numbers into two parts A1 , A2 ,
and each (!) of these parts is infinite, i.e. it comprises more elements than can be
expressed by any natural number.
Each of the parts A1 , A2 is the “complement” of the other, i.e. it contains all
elements which the other part does not contain: A2 = R \ A1 and A1 = R \ A2 .
Consequently, each part is completely determined by the other one, and it is enough
to take one of these two parts for the definition of the “irrational” number. This had
been realized by Russell (p. 203): take a “segment” instead of a “cut”, A1 instead of
(A1 , A2 ) = (A1 , R \ A1 ). However, less is not enough!
The definition of a single “irrational” number needs an infinitely large set of “rational”
numbers.

An enormous expenditure! However, also the other two constructions of the


irrational or the real numbers (from Weierstraß and from Cantor) are unable to
dispense with infinite number sets. Maybe there is no other way?
Everybody who accepts one of these constructions as a definition admits to the
existence of the actual infinite in mathematics. That means: she or he acts as if
the infinite set of “rational” numbers under consideration—Russell’s “segment”,
Cantor’s “sequence” or Weierstraß’ “irrational” number—was a well-defined object.

The State of Affairs Until Now

In Chap. 5 it was shown how much Leibniz resisted accepting the actual infinite in
mathematics.
The two subsequent chapters demonstrated that Johann Bernoulli as well as Euler
took the completely opposite view and calculated using actual infinite quantities.
Euler, especially, presented highly impressive examples for the productive use of
the actual infinite in his “Algebraic Analysis” (pp. 77f).
The remodelling of the “Algebraic Analysis” into the “Analysis of Values” since
1817/21 rendered the reputation of the actual infinite in analysis precarious. The
reason was that as soon as “value”—and, consequently, “number”—becomes basic
notions of analysis, they needed a characterization that is suitable for proofs. It
must become possible to create “numbers” in proofs—which means to create them
conceptually.
However, a concept of “number”, which is appropriate for proofs, existed (before
1872 resp. 1849 or 1838) only in a technical sense: the decimal number.
In the beginning of the thirteenth century, the Hindu–Arabic positional system
was brought into commercial Italy by Leonardo of Pisa (c1170–c1250). By the end
of the fifteenth century, this technique of writing numbers had reached the trading
towns in Germany and was thereby slowly superseding the Roman numerals. The
Rechenmeister taught the computation “by feathers or chalk in numerals” (Adam
Ries (1492–1559); that is written computation) to the businessmen as an alternative
The Price of Success: The Inclusion of the Actual Infinite in Mathematics 211

to computation on the abacus. In this process, the signs “+” and “−” started to
replace the Italian abbreviations “p” and “m”. Christoff Rudolff (c1499–c1545)
was the first to use the decimal comma in his textbook Hurried and Charming
Computation With the Help of the Ornate Rules of Algebra, Which Ordinarily are
Called the Coss, which was printed in Strasbourg in 1525. In 1585, the Dutch
businessman, engineer and physicist Simon Stevin (1548/9–1620) systematically
taught the computation with decimal numbers in his little book De Thiende (“About
the Tenth”).
La Géometrie from Descartes actually shows only rational numbers. But Leibniz
could not get along with them and used decimal numbers in his calculations. In
1821, Cauchy proved the Intermediate Value Theorem by dividing the distance h of
the two limits x0 and X of the domain into m equal parts x0 + m h
, x0 + 2 m
h
, x0 + 3 m
h
,
h 
. . . , X − m ; the newly found subinterval X − x1 = m is divided yet again into
h

intervals of length m1 (X − x1 ) = m12 (X − x0 ) = mh2 , etc. This can be understood as


the way to determine the decimal places of the required root X .
However, nobody actually proposed a figurative representation for “infinite” numbers.

A first proposal for this only appeared in the 1950s (see p. 227).
And we have been aware (p. 177), Weierstraß expressly wanted to exclude
infinity (∞) from analysis, although it was commonly accepted as being a “value”.

The Compromise

In this challenging situation—the urgency of creating a concept of an analytical


“number” that can be used in proofs, and at the same time a great scepticism towards
nearly every kind of the infinite—the analysts (and even Weierstraß) were finally
ready to swallow the bitter pill of accepting the concept of the actual infinite in 1872.
By this year, the majority was ready to make a small concession and accepted the
actual infinite at least conceptually (one or even two infinite sets of rational numbers
make up a real number). However, at the same time, they decisively insisted on
excluding the infinite numbers from analysis.
It is another paradox within the history of analysis that it was Georg Cantor of all
people, the man who had been committed to the recognition of conceptualizing the
actual infinite on a large scale, who found himself with both the obligation to prove
and the capability of proving that “the existence of actual infinitely small quantities
is impossible”. Cantor, being himself one of the greatest advocates for accepting
the actual infinite as a mathematical concept, especially in set-theory, nevertheless
struggled with all his might to exclude the actual infinite from arithmetic and actual
calculations.
Why should this be? Was it not Euler who had already shown what extraordinary
calculations could be carried out with infinite “numbers” within analysis (see p. 77)?
One should not underestimate the influence of Weierstraß’ personal aversion
towards the inclusion of the infinite into analysis—yet, even outside of his sphere
212 13 Analysis’ Detachment from Reality—and the Introduction of the Actual Infinite. . .

of influence (maybe in France) I don’t know of any attempt to establish the infinite
“as a number for calculation”. It wasn’t until 1976 that a mathematician (Detlef
Laugwitz) dared to propagate this in print. Yet, this happened a full two decades
after he had formulated an alternative version of analysis that provided for this.

Literature

Bertrand, J. (1849). Traité d’Arithmétique. Paris: Librairie de L. Hachette et Cie..


Cantor, G. (1872). Über die Ausdehnung eines Satzes aus der Theorie der trigonometrischen
Reihen. Mathematische Annalen, 5, 123–132. cited from Zermelo 1932, pp. 92–102.
Cantor, G. (1887/1888). Mitteilungen zur Lehre vom Transfiniten (2 parts). Zeitschrift für
Philosophie und philosophische Kritik, 91, 92, 81–125, 240–265. cited from Zermelo 1932,
pp. 378–439.
Cantor, G. (1889). Bemerkung mit Bezug auf den Aufsatz: Zur Weierstraß-Cantorschen Theorie
der Irrationalzahlen. Mathematische Annalen, 33, 476. cited from Zermelo 1932, p. 114.
Dedekind, R. (1901). Essays in the theory of numbers. (Chicago: The Open Court Publishing
Company), authorised translation by Wooster Woodruff Beman. https://www.gutenberg.org/
2/1/0/1/21016/.
Dedekind, R. (1854). Über die Einführung neuer Funktionen in die Mathematik. In R. Fricke, E.
Noether, & Ø. Ore (Eds.), 1930–32 (vol. 3, pp. 428–438).
Dedekind, R. (1872). Stetigkeit und irrationale Zahlen. Braunschweig: Vieweg. see also: Fricke,
Noether & Ø. Ore (Eds.), 1930–32 (vol. 3, pp. 315–334).
Frege, G. (1884). Die Grundlagen der Arithmetik. Breslau: Verlag von Wilhelm Koebner. reprints
Breslau 1934, Hildesheim 1961.
Frege, G. (1893, 1903). Grundgesetze der Arithmetik: Begriffsschriftlich abgeleitet (2 vols). Jena:
Pohle. https://gdz.sub.uni-goettingen.de/dms/load/img/?PPN=PPN593233409&DMDID=
DMDLOG_0001;http://gdz.sub.uni-goettingen.de/dms/load/img/?PPN=PPN593233549&
DMDID=DMDLOG_0001.
Fricke, R., Noether, E., & Ore, Ø. (Eds.) (1930–32). Richard Dedekind. Gesammelte mathemati-
sche Werke. (3 vols) Braunschweig: Vieweg.
Gabriel, G., Hermes, H., Kambartel, F., Thiel, C., & Veraart, A. (Eds.) (1976). Gottlob Frege:
Nachgelassene Schriften und wissenschaftlicher Briefwechsel. (Vol. 2). Hamburg: Felix Meiner
Verlag.
Gabriel, G., Kambartel, F., & Thiel, C. (Eds.) (1980). Gottlob Freges Briefwechsel mit D. Hilbert,
E. Husserl, B. Russell, sowie ausgewählte Einzelbriefe Freges. No 321 of the Philosophische
Bibliothek. Hamburg: Felix Meiner Verlag.
Gericke, H. & Vogel, K. (1965). Simon Stevin: De Thiende (Die Dezimalbruchrechnung). Frankfurt
am Main: Akademische Verlagsgesellschaft.
Hankel, H. (1867). Vorlesungen über die complexen Zahlen und ihre Functionen in zwei Theilen.
I. Theil. Theorie der complexen Zahlsysteme, insbesondere der gemeinen imaginären Zahlen
und der Hamilton’schen Quaternionen nebst ihrer geometrischen. Leipzig: Leopold Voss.
https://www.urn:nbn:de:bvb:12-bsb10081922-7.
Hausdorff, F. (1914). Grundzüge der Mengenlehre. reprint New York: Chelsea Publishing Com-
pany, 1949.
Heine, E. (1872). Die Elemente der Functionenlehre. Journal für die reine und angewandte
Mathematik, 74, 172–188.
Hilbert, D. (1899). Grundlagen der Geometrie. In Fest-Comitee (Ed.). Festschrift zur Feier der
Enthüllung des Gauß-Weber-Denkmals in Göttingen. Leipzig: Teubner.
Hilbert, D. (1900). Über den Zahlbegriff. Jahresbericht der Deutschen Mathematikervereinigung,
8, 180–184.
Literature 213

Hilbert, D. (1917). Axiomatisches Denken. cited from Hilbert 1932–35, (Vol. 3, pp. 146–156).
Hilbert, D. (1932–35). Gesammelte Abhandlungen, 3 vols. reprint: New York: Chelsea Publishing,
1965.
Laugwitz, D. (1976). Unendlich als Rechenzahl. Der Mathematikunterricht, 22(5), 101–117.
Laugwitz, D. (1978). Infinitesimalkalkül: Kontinuum und Zahlen. Eine elementare Einführung in
die Nichtstandardanalysis. Mannheim, Wien, Zürich: Bibliographisches Institut.
Müller, J. H. T. (1838). Lehrbuch der allgemeinen Arithmetik für Gymnasien und Realschulen,
nebst vielen Übungsaufgaben und Excursen. Halle: Verlag der Buchhandlung des Waisen-
hauses. https://sammlungen.ub.uni-frankfurt.de/urn/urn:nbn:de:hebis:30:4-103650.
Müller, J. H. T. (1855). Lehrbuch der allgemeinen Arithmetik für höhere Lehranstalten. Erstes
Heft. Second revised edition. Halle: Verlag der Buchhandlung des Waisenhauses.
Ries, A. (1522). Rechenbuch auff Linien vnd Ziphren . . . . Frankfurt: Chr. Egen. Erben, 1574.
reprint: Darmstadt: Hoppenstedt
Rudolff, C. (1525). Behennd vnnd Hübsch Rechnung durch die kunstreichen regeln Algebre / so
gemeincklich die Coß genen̄t werden . . . Argentorati [= Straßburg]: Vuolfius Cephaleus.
Russell, B. (1923). Bertrand Russell (1919): Einführung in die mathematische Philosophie. In
Philosophische Bibliothek (Vol. 536). Hamburg: Felix Meiner Verlag, 2002. German: Emil
Julius Gumbel.
Tannery, J. (1886). Introduction à la théorie des fonctions d’une variable. Paris: A. Hermann.
Zermelo, E. (Ed.) (1932). Georg Cantor. Gesammelte Abhandlungen mathematischen und
philosophischen Inhalts. Hildesheim: Georg Olms Verlagsbuchhandlung, reprint 1966.
Chapter 14
Analysis with or Without Paradoxes?

In this chapter, we shall introduce attempts to base the calculus on a different


concept from that of Weierstraß.

Built on Very Thin Ice: Cantor’s Diagonal Argument

The Presentation of Evidence

It is quite astonishing to discover which precarious arguments are sometimes


accepted as being mathematically sound.
An outstanding example thereof is the (time and time again, by experts and non-
experts alike) celebrated “proof for the existence of the uncountable infinity”. This
bizarre “proof” runs as follows (there should be added some technical details in case
the number ends with infinitely many nines, but this is just technical and does not
touch the essential reasoning; therefore, it can be omitted):
Suppose we could enumerate the numbers between 0 and 1, written as decimals.
If this is done, we arrive at a list

b1 = 0.a11 a12 a13 a14 . . .


b2 = 0.a21 a22 a23 a24 . . .
b3 = 0.a31 a32 a33 a34 . . .
..
.

The akn are digits from 0 to 9.—Possibly, we have

b59284 = 0.5000 . . .

According to our assumption, the list includes each number between 0 and 1.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 215
D. D. Spalt, A Brief History of Analysis,
https://doi.org/10.1007/978-3-031-00650-0_14
216 14 Analysis with or Without Paradoxes?

The surprising “argument” that should supply the contradiction to our supposition
is as follows:
• Take the number c = 0.c1 c2 c3 c4 . . ., defined by
c1 = any number, but not a11 ;
c2 = any number, but not a22 ;
c3 = any number, but not a33 ;
etc.
• This number c differs obviously from all bm . (This c deviates in the m-th digit from
bm , as this digit of bm is = amm = cm .)
• Consequently, the list of the bm does not contain c.
• But this contradicts the initial assumption that we had a list bm of all numbers between
0 and 1.
• Contradiction. Presumed end of proof.

Every modern book on the mathematical infinite presents this “proof”, including
the book Proofs from THE BOOK from Martin Aigner and Günter M. Ziegler from
1998 as well (pp. 92f). And certainly the authors are excited: star writer David Foster
Wallace (1962–2008) e.g. called this proof “both ingenious and beautiful—a total
confirmation of art’s compresence in pure math” (p. 255). An appraisal such as
“ingenious” calls for attention, no doubt about that!

The Impotence of This Reasoning

Let us pause for thought and ask: does this actually convince anyone?
Let us work out the details. At once we realise:
The above “proof” is not necessarily valid.
(This is no news. I only repeat a well-known reasoning.)—Therefore, I ask two
questions.
(1) How many decimal numbers with n digits in between 0 and 1 do exist?
The answer is straightforward:

1. If n = 1, we have exactly 10, to wit: 0.0; 0.1; 0.2; . . . 0.9.


2. If n = 2, we have exactly 100 = 102 , to wit: 0.00; 0.01; 0.02; . . . 0.99.
etc.
n. Consequently, in the general case n, we have 10n numbers.

And now my second question:


(2) How many digits of c are safely determined by this so-called proof?
Built on Very Thin Ice: Cantor’s Diagonal Argument 217

The answer to this question is evident: exactly n—one for each digit under
consideration. But what is the consequence? Plainly this: the number c defined by
this professed proof certainly differs from the first n numbers in the list—but this
does not demonstrate CONCLUSIVELY that it is not included in the list at all; instead,
it only proves: c = bm as well as m > n, nothing else.
To be practical, let n = 2. We write down the list of the 102 = 100 numbers in
between 0 and 1 with two digits after the decimal point:

b1 = 0.00 , b2 = 0.01 b11 = 0.10 ... , b100 = 0.99 .


b3 = 0.02 b12 = 0.11
... , ... ,

Let us take c = 0.12. Then we know c = b1 and c = b2 ; but we have c = b13 .


c is not to be found among the first n = 2 numbers but later on it is included in the
list.
This reasoning imputes a (MOST) EXTREMELY STRANGE hidden assumption to
this “both ingenious and beautiful” “proof”. It is this one:

10∞ = ∞ .

For in fact it is proclaimed: The number c is not contained in the list of the bm .
What is true, however, is just this: The number c is not among the first infinitely
many (“∞”) numbers of the list, but only subsequently.
However, who is demanding to calculate with infinity in this way? It is true, e.g. Weierstraß
authorized this—i.e. 10∞ = ∞—but at the same time he demanded to exclude infinity
(“ ∞”) from analysis (p. 177).
But Euler did use infinity in his calculations. He even assumed i > i − 1 (p. 77) and
moreover

10i > i

for infinitely large i.

There is no obligation to think like Euler and also none to think like Weierstraß.
In mathematics we do not deal with dogmas but with arguments.
From Chap. 5 we know:
There exists no conclusive or ultimate reasoning for dealing with the
actual infinite in analysis.
218 14 Analysis with or Without Paradoxes?

The Origin of the “Diagonal Argument”

The “proof”, which is described above, was invented by Georg Cantor. Cantor
demonstrated thus in 1891 for the third time:
Theorem. There exist infinite sets which cannot be related one-to-one
to the set of natural numbers.
Cantor designed his proof slightly differently. (He took only two different digits
instead of ten as we did.—Strikingly, this proof, given for binary numbers instead
of decimal numbers, is not only not conclusive but also wrong—why exactly? And:
is it permissible to make a proof about the real numbers dependent on the mode of
their representation?)

The True Understanding of the “Diagonal Argument”

It was Bertrand Russell who unearthed the rational essence of Cantor’s proof.
The true theorem of Cantor’s proof is this:
Theorem. The number of subsets of a set is larger than the number of the
elements of this set.
Let us take an example. The given set may contain two elements, a and b:

M = { a, b }.

Then M has the following 22 = 4 subsets, first of all the empty set { } = ∅:

∅, { a }, { b }, { a, b }.

Hence, there are more subsets than elements: 4 > 2.


Now we faithfully present Russell’s general proof of his theorem (changed,
however, are the names of the sets—which Russell calls “classes”). Like Cantor,
Russell starts with a list of subsets signified by the elements. Russell calls this list a
“one-one correlation R ”.
When a one-one correlation R is established between all the members of M and some of its
sub-classes, it may happen that a given member x is correlated with a sub-class of which it
is a member; or, again, it may happen that x is correlated with a sub-class of which it is not
a member.

We interrupt Russell and illustrate his set-theoretical construction with the help
of our example, and that twice.
Built on Very Thin Ice: Cantor’s Diagonal Argument 219

1. Illustration. The list of the correspondences of the elements and the subsets of M =
{a , b} might be

a → { a } , b → { a , b } .

Then we have for both elements a and b that they correspond to a subset
that contains them as an element.
2. Illustration. A second list might be:

a → ∅ , b → { b } .

In this case, only element b corresponds to a subset, which contains itself as


an element, but element a does not.

Now we continue with Russell’s proof.


Let us form the whole class, N say, of those members x which are correlated with sub-
classes of which they are not a member.

In our first illustration, we have N = ∅; in the second, it is N = {a}. We continue


with the proof. Now follows the decisive sentence (I give it in italics):
This is a sub-class of M, and it is not correlated with any member of M.
For taking first the members of N , each of them is (by the definition of N ) correlated
with some sub-class of which it is not a member, and is therefore not correlated with N .
Taking next the terms that are not members of N , each of them (by the definition of
N ) is correlated with some sub-class of which it is a member, and therefore again is not
correlated with N .
Thus no member of M is correlated with N .
Since R was any one-one correlation of all members with some sub-classes, it follows
that there is no correlation of all members with all sub-classes.

It is extremely impressive how a few words can express a really entangled


reasoning. Those who are able to understand this demonstration from Russell may
be pleased with their capability of abstract thought.
If you love formalism, you may write down this proof in the following manner
(usually “2M ” denotes the set of all subsets of M; “M \ N ” indicates the difference
of the sets M and N):

R
Let R : M −→ 2M ; therefore M  x −→ R x ∈ 2M .
Let N = {x|x∈
/ Rx }  M. Then we have for all x ∈ M : R x = N.
Proof :
1. Let x ∈ N . Then x ∈
/ Rx, and consequently, N = R x (for N  x ∈ Rx) .

2. Let x ∈ M \ N . Then we have x ∈


/ N ; therefore, x ∈ R x , and again R x = N .
(In both cases, an element of one set is shown,
which is missing in the other set.)
1 & 2: x ∈ M ⇒ R x = N , as was claimed.
220 14 Analysis with or Without Paradoxes?

In this way, Russell established:


Theorem. 2n is always larger than n—even if n is infinite.
This implies that the infinite cardinal numbers do not have a maximum.

The Significance of the “Diagonal Argument”

That is why the so-called Diagonal Argument from Cantor is a method of proof
in set-theory. It gives the means of proving e.g. that there always exist larger
“cardinalities”, even in the infinite.
The “numbers” in set-theory are “cardinal” numbers. In analysis they do not
feature (at least not in traditional analysis, in calculus). Calculus needs “numbers
for calculations”.
Therefore, Russell’s proof that 2n > n (which holds in set-theory) does not
contradict Weierstraß’ principle that 10∞ = ∞. Quite the contrary: if one accepts
Cantor’s Diagonal Argument as permissible reasoning in analysis, you must also
accept Weierstraß’ dictum that 10∞ = ∞, for otherwise the Diagonal Argument
fails! This is explained above.
So, if you think analysis (in this sense) and set-theory as one thing, you will be convinced
that 10∞ = ∞ (in analysis) and that 2n > n for all n, including infinite numbers (in
set-theory).

Some will like this . . . But in some vague sense, this does not look altogether
consistent. Luckily, one is not forced to think in this style.
For the record: on the case of numbers used for calculating, i.e. in analysis, you
may accept the Diagonal Argument. Or you may not. In any case, it is not conclusive
there. And if you want to relate it to the facts connected to the decimal numbers (for
which it CLAIMS validity), it is even less conclusive. For we have seen:
The Diagonal Argument demands the acceptance of an argument regarding infinity,
which is invalid in any finite case.

As we know, Leibniz once argued in the same way, thereby surprising Johann
Bernoulli (p. 55).
It might cross one’s mind to interpret this argument as a proof by induction:
• Start: choose c1 = a11 .
• Inductive step: Let cn = ann . Then choose cn+1 = an+1,n+1 .
—Finished.
An objection to this kind of reasoning is obvious: it is presupposed that the first
part of each number 0.c1 c2 . . . cn formed in this way is contained in the list of the bm :
0.c1 c2 . . . cn = bm ; of course, it is m > n. So what? If any first part of c is
contained in the list of the bm —why should this hold no longer for c itself?
If you do not accept the Diagonal Argument in analysis (to repeat: this is as
legitimate as is the opposite), you are accepting the validity of
Paradox I: Conditionally Convergent Series 221

10i > i

for infinite numbers used in calculations. Finally, we shall demonstrate where this
kind of thinking will lead to.

Paradox I: Conditionally Convergent Series

Let us finish our detour into set-theory and return to analysis. There are certain
curiosities that have arisen during the last three and a half centuries that caught the
attention of some analysts.

1− 1
2 + 1
3 − 1
4 + 1
5 − + . . . = ln 2 .

Let us now tackle some of these curiosities in more detail, using the title “para-
doxes”. The creator of this name will be uncovered later.
In order to prepare for this, we now return to Riemann. We know: Riemann is
always one for offering a surprising perspective.

A Mathematical Monstrosity: The Riemann Theorem on


Rearrangements of 1854

Riemann proved the following oddity, nowadays called the “Riemann Theorem on
Rearrangements”, in his habilitation thesis of 1854:
Theorem. If a convergent series is no longer convergent when its terms are made
positive, the terms can be rearranged in such a way that the series converges on any
arbitrarily chosen value.

The proof seems to be straightforward. One assumes a convergent series of


numbers:


ak = s .
k=1

Infinitely many terms have the sign +, and infinitely many have the sign −. We call
the positive terms of the series b, the negative ones c:



k=∞ ∞

k=∞
bi = ak as well as ci = ak .
i=1 ak >0 and i=1 ak <0 and
k=1 k=1
222 14 Analysis with or Without Paradoxes?

So we have bi > 0 as well as ci < 0. Neither of those series converges:



k=∞ ∞

k=∞
bi = ak = ∞ as well as ci = ak = −∞ .
i=1 ak >0 and i=1 ak <0 and
k=1 k=1


If both series had finite sums, then the series |ak | would also converge, which
contradicts our assumption. If only one of these two series converges, the initial
series could not converge. That is why none of our partial series converges.
Now Riemann’s reasoning: let D, say D > 0, be an arbitrarily given value.
Then we approach D gradually—at first from below, then from above; and so on,
alternating. That is to say, we choose the smallest number n1 with


n1
n1
m1
bk > D ; thereafter, the smallest number m1 with bk + ck < D ,
k=1 k=1 k=1

etc. The difference to D will never be more than the absolute value of the term that
is the one before the last sign change. But ak converges, and therefore, bk as well
as |ck | are decreasing with increasing k below any given quantity; and so does the
difference to D—which seems to prove all.
This theorem is a monster. Weierstraß obviously did not like it, and he indeed
mentions the notion of conditional convergence (see p. 184), but he ignores this
concept in his own analysis.
Why is this theorem a monster? Because it assumes that a mathematician is able
to decide infinitely many times AT WILL to change between the partial series of the bi
and ci back and forth. “An infinite number of arbitrary choices is an impossibility”
says Russell. Infinitely many actual decisions have nothing in common with reality.
Note: A proof may be easy, but the proven statement may be absurd.

Mitigation

A theory of infinitely many single choices in analysis certainly is an extreme. An


essential mitigation is to think about a regular change of infinitely many terms of
a series. We might have more chance of acceptance if we concentrate on regular
mathematical constructions—although Weierstraß might still have taken a different
view.
Let us take an example. We start with a series A, halve every term, add both
series term-by-term, and regain the initial series, although slightly rearranged.
We have:

A= ln 2 = 1 − 1
2 + 1
3 − 1
4 + 1
5 − 1
6 + 1
7 − 1
8 + −...
Paradox II: Methods of Summation 223

Therefore, 1
2 ln 2 = 1
2 − 1
4 + 1
6 − 1
8 + −... ,

and consequently B= 3
2 ln 2 = 1 + 1
3 − 1
2 + 1
5 + 1
7 − 1
4 + + − ...
= [rearrangement of] ln 2.

Do we have A = B, i.e. ln 2 = 3
2 ln 2?
According to Weierstraß, this conclusion is a “fallacy” (p. 177). Why?—Is the
Riemann Theorem on Rearrangements an answer?—Is this answer satisfactory?

Paradox II: Methods of Summation

For divergent series like

A = 1 − 1 + 1 − 1 + 1 − 1 + −...

sometimes other summation methods are in use, e.g. the so-called method of
C-1 summation:
s 1 + s 2 + s 3 + . . . + sn
S C1 = lim .
n→∞ n
In case of series A, the partial sums are s1 = 1; s2 = 1 − 1; s3 = 1 − 1 + 1;
s4 = 1 − 1 + 1 − 1 etc., that is:
s1 = s 3 = s 5 = . . . = 1 ,
s2 = s 4 = s 6 = . . . = 0 ,

and consequently, the terms of S C 1 are


s1
p1 = 1 = 1,
s1 +s2
p2 = 2 = 1
2 ,
s1 +s2 +s3
p3 = 3 = 2
3 ,
s1 +s2 +s3 +s4
p4 = 4 = 2
4 = 1
2 ,
s1 +s2 +s3 +s4 +s5
p5 = 5 = 3
5 ,
...
i.e. p2n−1 = n
2n−1 ,

p2n = n
2n = 1
2 ,
all together: S C 1 = lim pn = 1
2 .
n→∞
224 14 Analysis with or Without Paradoxes?

With this method, we do get:

A = 1 − 1 + 1 − 1 + 1 − 1 + −... = 1
2 !?

Paradox III: The Convergence of Function Series

Let us take the geometric series

s(x) = 1 − x + x 2 − x 3 + x 4 − + . . .

The well-known way to calculate the sum s(x) is to add sn (x) and x · sn (x):

sn (x) = 1 − x + x 2 − x 3 + x 4 − x 5 + − . . . + (−1)n−1 x n−1

x · sn (x) = x − x 2 + x 3 − x 4 + x 5 − + . . . + (−1)n−2 x n−1 + (−1)n−1 x n

(1 + x) · sn (x) = 1 + (−1)n−1 x n

1+(−1)n−1 x n
and sn (x) = 1+x if x = −1.

The consequence is
If |x| < 1 we have s(x) = lim sn (x) = 1
1+x ,
n→∞

if x=1 we have s(1) = 1


2 (1 − 1 + 1 − 1 + − . . . ), which does not exist.

Yet, “does not exist” is not a beautiful result of a calculation.

Paradox IV: The Term-by-Term Integration of Series

Let us take the series of functions


nx
fn (x) = with 0  x  1 .
1 + n2 x 4

As lim nx
= 0 if 0  x  1 , we usually conclude for the limit n → ∞:
n→∞ 1+n2 x 4

nx
f (x) = lim fn (x) = lim = 0. (‡‡)
n→∞ n→∞ 1 + n2 x 4
Paradox IV: The Term-by-Term Integration of Series 225

But if xn = √1
n
(and consequently lim xn = lim √1
n
= 0), the value of f (x) at
n→∞ n→∞
the value 0 = lim √1
n
is
n→∞
  √ √
n n
lim fn √1 = lim 2 = lim =∞!
n→∞ n n→∞ 1+ n2 n→∞ 2
n

We remember: in these cases Cauchy concludes ∞ to be a value of the function


f (x) at the value x = 0 (p. 124). However, present-day analysis handles things
differently and demands:

n·0 0
f (0) = lim fn (0) = lim = = 0,
n→∞ n→∞ 1 + n2 · 0 1

unambiguous and crystal-clear.


If we have, following ‡‡, f (x) = 0 for all 0  x  1, integration is easy:

1 1
f (x) dx = 0 · dx = 0 .
0 0

But let us integrate term-by-term:

1 1
nx
fn (x) dx = dx .
1 + n2 x 4
0 0

Substituting z = nx 2 gives dz = 2nx dx and we arrive at

n z=n
dz
= 1
· = 1
· arctan z = 1
· arctan n .
2 1 + z2 2
z=0
2
0

With the help of f (x) = lim fn (x), we conclude


 1  1
f (x) dx = lim fn (x) dx = lim 1
· arctan n = π
.
0 n→∞ 0 n→∞ 2 4

This contradicts the previous result. Consequently, the first equality in the last line
is wrong—the others being indubitable. So, for this series, it is not possible to
interchange integration and limit.
Conclusion: We need a theorem to tell us, in which cases the interchange
of integration and limit is allowed. Naturally, it would be more attractive if the
calculation itself would make the problem visible. Some analysts enjoy calculating
more than hunting for theorems that allow them to calculate.
226 14 Analysis with or Without Paradoxes?

Is an Analysis Without Those Paradoxes Possible?

A Source from the Years 1948–53

Curt Schmieden (1905–91) was an eminent mathematician. In 1934, he became a


professor at the Technical University of Darmstadt, where he was awarded a chair
in mathematics in 1937, and in 1957/58 he was the rector of this university. In a
manuscript dating from 1948–53, Schmieden commented on examples of this kind
(ALL emphases added):
Such examples could be given endlessly; the deeper one penetrates into mathematics, the
more such paradoxes emerge—even if one completely disregards set-theory.
The most thrilling aspect of this kind of mathematics is that in spite of those paradoxes
one always arrives at a “true” result for each concrete problem. Consequently, it does
not surprise that some mathematicians decide at some point that the naive intuition about
infinity does no longer suffice in some way or other, whereas working along definite rules in
regard to infinity makes it possible to tame those quantities.
Nevertheless, there remains a feeling of discomfort. Also the question remains,
WHETHER IT IS NOT POSSIBLE AFTER ALL TO CONSTRUCT ANALYSIS IN SUCH A
WAY THAT IT WORKS ALONGSIDE FIXED RULES AS IT HAS BEEN DOING SO FAR, AND
THIS IN SUCH A WAY THAT the kinds of aforementioned PARADOXES DO NOT EMERGE .
Thus our naive intuition which according to Goethe, more often than not wishes “to
represent the infinite with the help of the finite” [1] , would be granted.

Schmieden, being “fully aware of the imperfections of this attempt” stated some
foundational principles along those lines. Among them was the following that is
obviously inspired by Johann Bernoulli and Euler but which also surpasses both
conceptually:
Reason forcibly demands that there must exist infinitely large numbers, which can only be
meaningfully defined as follows:
An infinitely large natural number is a natural number, which cannot be
grasped by an unlimited continuation of the process of counting, for counting
is necessarily bound to the degree of finiteness.
Consequently, to arrive at such a “number”, it requires, figuratively speaking, a jump, or
as a metaphor: infinitely large numbers form the “horizon of finiteness”.

The idea is:


If we conceive an infinitely large number called  (capital omega), reached by such
a jump, obviously nothing speaks against but everything for it that we are allowed
to calculate with such a “number” in just the same way as with a usual number.
Basically, Johann Bernoulli and Euler had shared this opinion. A little later Schmieden gives
the following definition:

ω = 1 (small omega) is an infinitely small number which is smaller than each


positive rational number and which is defined by the level of finite numbers. (ω is
a “zero number”.)

1 This is the exact translation of Goethe’s words.


Is an Analysis Without Those Paradoxes Possible? 227

In the same way as for any infinitely large number, there exists one of a larger order[2] ,
and there exists for any zero number one of a smaller order[2] .
Therefore, our ordinary zero known from conventional analysis is only allowed to be
used as another notation of the identity a ≡ a.

The latter would have pleased Frege (p. 195), and Weierstraß, too, would have
given his approval in regard to such precision in relation to equality.
Finally, there is Schmieden’s foundational equation:
We define

lim n =  ;
n→∞

each other limit has to be related to this expression.

(The sign “∞” is not a good choice here and should have been avoided
altogether.)
It is fairly evident that it is possible to construct from infinitely large natural
numbers suitable rational numbers, e.g. 1 or 1 + 1 etc. The essential question is:
Is this enough for analysis? After all, we now have “infinitely near” numbers, and
this also in the neighbourhood of each rational number. In Schmieden’s words:
Around each ordinary rational number there exists an ω-sphere which does not include any
other ordinary rational number, but which already contains infinitely many rational numbers
of the second class [i.e. level] in the lowest ω-level.
With the help of the enlarged notion of rational numbers, which inevitably follows from
the introduction of , we reach the seemingly paradoxical result that analysis, if founded
on -numbers, gets along with these [-]rational numbers, and even more—that no other
numbers can occur, because each number defined via the usual limit arises in our system as
an -rational number.

Here we have the essence of Schmieden’s idea: instead of the real numbers
(following Müller or Bertrand or Cantor or Dedekind), we take “-rational”
numbers, e.g. q ± 1 or q ± 5 −7+3
2
3 3
, etc. These can be used in calculations in the
 −4
usual manner. The “-rational” numbers should be enough because—as we know
since Euler—analysis consists essentially of calculations, especially of calculations
with the infinite.

Schmieden Dissolves the Paradoxes

It was the aim of Schmieden to dissolve such “paradoxes” as shown above with the
help of more precise calculations—i.e. with his -numbers. Schmieden states:
We especially inspect those instances where analysis puts up signs with the warning: here
it is forbidden to calculate as in the finite case!

In what follows, representative examples of his method are shown.

2 emphases added
228 14 Analysis with or Without Paradoxes?

Resolving paradox I (p. 227).


Take the series

ln 2 = 1 − 1
2 + 1
3 − 1
4 + 1
5 − 1
6 + 1
7 − 1
8 + −...

Schmieden now takes care in regard to the infinite (“+ − . . .”) and calculates
precisely:



(−1)n−1
ln 2 = n =1− 1
2 + 1
3 − 1
4 + 1
5 − 1
6 + 1
7 − 1
8 + −... + 1
−1 − 1
 =: A .
n=1

Here  is assumed to be even. But careful: “+ − . . .” has different meanings in the


last two lines of formulae. (a) In the first line, “+−. . .” indicates “and so on” without
last term. (b) In the second line, “+ − . . .” marks a gap, like in 1 + 12 + 14 + . . . + 21n .
Now Schmieden rearranges the series A and supposes that  divided by 4 leaves
no remainder. (Why not? If it is useful! At a pinch, make  = 4.) He then realizes
the sum formula of the rearranged series:


/4
 
1
4n−3 + 1
4n−1 − 2
4n = 1+ 13 − 12 + 15 + 17 − 14 ++−. . .+ −3
1
+ −1
1
− 1/2 =: B .
n=1

We see:
It is true, A and B have the same positive terms (all odd denominators, including  − 1),
but the second half of the negative terms of A (even denominators) are missing in B, up to
:

1 1 1 1
/2

− − − − ... − =− 1
=: −C ,
/2 + 2 /2 + 4 /2 + 6  2n

n= 4 +1

and we have

A=B −C.

There you are! That is why we have B > A!


B is not at all a rearrangement of A, for it has infinitely fewer (negative) terms.

Schmieden even calculates the value of this difference C. This he does by


integration—using the fact that in case of a continuous function and an infinitely
small interval the estimated value of the area under this function only differs from
the integral by an infinitely small amount. We denote this just like we did in the case
Is an Analysis Without Those Paradoxes Possible? 229

of Johann Bernoulli (p. 62) with ≈ and observe that ω = 1


 is infinitely small. We
are going to calculate


/2

C= 1
2n · ω
ω .
n= 
4 +1

Schmieden defines z = 2nω, consequently dz = 2ω (as n always grows by 1),


ω
and therefore 2nω = 12 dz 1
. But 2nω is continuous (in the sense of ε-δ). Lower limit:
 z  
z = 2nω = 2 4 + 1 · ω = 2 + 2ω, upper limit: z = 2nω = 2 2 · ω = 1, and
1

therefore:

1
C≈ 1
2 · dz
z ≈ 1
2 · (ln 1 − ln 12 ) = 12 (ln 1 − ln 1 + ln 2) = 1
2 ln 2 .
z= 21 +2ω

So we get in fact

B −C ≈ 3
2 ln 2 − 1
2 ln 2 = ln 2 = A
resp. A − B = −C ≈ − 12 ln 2 ,

as we expected. Nothing of the kind 3


2 ln 2 = ln 2!
We record:
The investigation of regular “rearrangements” of conditionally convergent series leads
to comparable results: seemingly paradoxical equations of conventional analysis are
explained in calculations with -numbers through the omission of infinitely many
terms with a finite sum.

Besides, the concept of “conditional” (i.e. not “absolute”) convergence is


dispensable. Instead, all series are treated equally, like finite sums. (This would have
pleased Weierstraß.)
Interestingly, in their publication from 1958, Schmieden and Laugwitz hint
at the fact that “nevertheless, the difference between absolute and conditional
convergence remains important for technical calculations, insofar as absolutely
convergent series do allow arbitrary rearrangements without further considerations”.
Obviously, this sentence is a concession by the authors and is probably due to the
pressure exerted by the editors of the journal—because essentially it contradicts the
foundational view of the authors: in their way of precisely calculating, all series are
treated equally.
Resolving paradox II (p. 223).
Schmieden considers the C-1 Summation of the series

A = 1 − 1 + 1 − 1 + 1 − 1 + −...
230 14 Analysis with or Without Paradoxes?

from p. 223. There we examined the expression

s 1 + s 2 + s 3 + . . . + sn
S C 1 = lim .
n→∞ n
Schmieden presents this series more conveniently via  instead of lim and ∞:

s 1 + s 2 + s 3 + . . . + s
AC 1 = .

This fraction he writes down in more detail. Thereby he pays attention to the
following: (i) The first 1 of series A is included in all partial sums sk ; this amounts
 . (ii) The second 1 of series A is only included in  − 1 of
to the first term being 
the partial sums sk (i.e. not in s1 ); this gives the second term as −1
 ; etc.
−(−2) (−1)+1
AC 1 = 
 − −1
 + −2
 − + . . . + (−1)  + 

= 1 − (1 − ω) + (1 − 2ω) − + . . . + (−1) · 2ω + (−1)+1 · 1ω

He then considers the difference of A − AC 1 step-by-step, going from term to


term, as follows:

A − AC 1 = 
0 −ω +2ω −3ω +4ω ... +(−1)+1 (−ω)
0
  
−ω
  
ω
  
−2ω
  

  
...
  

− 2 ω = − 12 if  even, then A=0
=
+ −1
2 ω ≈ +2
1
if  odd, then A = 1.

All this is very accurate. Note the sum A has the value 12 , but we have AC 1 = − 12
if  is even, if not we have AC 1 ≈ 12 . The sum A ALWAYS has one of the values 1
or 0.—No paradox whatsoever.
The following question presents itself:

What constitutes the sum AC 1 really?


Schmieden just calculates. For the general

A = a0 + a1 + a2 + . . . + a−1
Is an Analysis Without Those Paradoxes Possible? 231

he takes, as we have already seen,


−1
AC 1
= 1
 sn = 1
 (a0 + ( − 1)a1 + ( − 2)a2 + . . . + ( − ( − 1))a−1 ) .
n=0

Especially, if

A(x) = 1 + x + x 2 + . . . + x −1 ,

he substitutes an = x n and gets


−1  
AC 1 (x) = 
1 1  · 1 + ( − 1)x + ( − 2)x 2 + . . . + ( − ( − 1))x −1
sn = 
n=0

= 1 − x(1 − ω) + x 2 (1 − 2ω) − + . . . + (−1)−1 x −1 (1 − (1 − )ω)


−1
−1
= (−1)n x n + ω (−1)n+1 nx n
n=0 n=0


−1
= s (x) + ω (−1)n+1 nx n .
n=0

Therefore, we have


−1
s (x) − A C1
(x) = ω (−1)n nx n = K .
n=0

K generates the convergence of s (x) to AC 1 (x), but only in the infinite: As the
factor ω in K shows, a finite number of terms of K will have sum ≈ 0.
Resolving paradox III (p. 224).
From the formula

1 + (−1)n−1 x n
sn (x) = if x = −1 ,
1+x

Schmieden directly concludes


 
s (x) = 1
1+x 1 + (−1)−1 x  . (§§)

1. If |x| < 1 and |x| ≈ 1, we get x  ≈ 0, and consequently,

s (x) ≈ 1
1+x .
232 14 Analysis with or Without Paradoxes?

2. If x = 1 , we get for an even or odd number  of terms:



  1 if  is odd
s (1) = 1
2 1 + (−1) −1 =
0 if  is even.

3. If x < 1 and x ≈ 1, the tricky x  is calculated by means of a trick. Like Euler,


Schmieden takes for ν > 0:

x := 1 − ξ ων

where ξ is finite and positive. He gets (remember from p. 77: instead of


n
lim 1 + xn = ex , Schmieden has of course (1 + ωx) ≈ ex ) with the help
n→∞
of the following subsidiary calculation:


    ⎨≈ 1
⎪ if ν > 1 ,
≈ e−ξ ω
ν−1
x = 1 − ξ ων = 1 − ω · ξ ων−1 e−ξ

= if ν = 1 ,


⎩≈ 0 if ν < 1 ,

and finally for §§ (as 1 + x ≈ 2), the result


⎧  

⎨ 2 1 + (−1)
1
⎪ −1 if ν > 1 ,
−ξ

s (x) ≈ 2 1 + (−1)
1 −1 e if ν = 1 ,


⎩1 if 0 < ν < 1 .
2

The case |x| > 1 is judged by Schmieden as being “nonsensical” (Fig. 14.1).

<
Fig. 14.1 The function s (x) = 1 − x + x 2 − + . . . in [0, 1], at right for x ≈ 1
The First Formal Version of a Nonstandard-Analysis in the Year 1958 233

Everything is exactly explained: if ν = 1 and:


(i)  is odd, the function s (x) increases from the value 12 near x = 1 as
 
e-function 12 1 + e−ξ up to the value 1 for x = 1 (i.e. up to ξ = 0).
(ii)  is even, the function decreases from the value 12 near x = 1 as
 
e-function 12 1 − e−ξ up to the value 0 for x = 1.
In other words: depending on whether  is odd or even, the value of the function
near x = 1 decreases from 12 to 0 or increases from 12 to 1. This is by no means
indeterminate! Of course, some calculation is needed—in mathematics, you do not
get anything without work. The distinction between  being odd or even shows the
different possibilities of the results of your calculation.
Resolving paradox IV (p. 224).
Only Eq. ‡‡ is thorny. Schmieden considers yet again the details and studies

x
f (x) = .
1 + 2 x 4

As before, he takes x = ξ ων and gets




⎪0 if ν > 1 ,





⎪ ξ if ν = 1 ,



⎪ −3

⎪ ξ  , (0 < α < 2) , if 13 < ν < 1,
α
 ν ξ ων−1 ξ 1−ν ⎨
ξ 1−ν
f ξ ω = = ≈ as 4 2−4ν ∼ 
1−ν−2+4ν
1+ξ ω4 4ν−2 1+ξ 4 2−4ν ⎪
⎪ 1+ξ



⎪ = 3ν−1



⎪ −3

⎪ξ if ν = 13 ,



0 if 0 < ν < 13 .

Therefore near 0 in the case 1


3  ν  1, we unambiguously obtain

f = f !

Schmieden also deals with delta-functions (indeed, if you know -numbers,


you have true delta-functions!). This would lead us astray, but one example of an
elementary delta-function will nevertheless be shown below.

The First Formal Version of a Nonstandard-Analysis in the


Year 1958

In 1954, the then young Detlef Laugwitz (1932–2000) was handed a manuscript
by Carl Friedrich von Weizsäcker (1912–2007) to report on it in his seminar in
234 14 Analysis with or Without Paradoxes?

Göttingen. After Laugwitz’ affirmative reaction to it, Weizsäcker brought him into
contact with the author Curt Schmieden. This was the beginning of a very fruitful
cooperation, and in 1962, at the age of 29, Laugwitz took up a position as chair in
mathematics at the University of Darmstadt.
Four years earlier, in 1958, and after some quarrels behind the scenes, the
renowned journal Mathematische Zeitschrift had published an article by Schmieden
and Laugwitz in its volume 69. Later it turned out to be the first publication in a
field nowadays called “nonstandard-analysis”.
In it, the two authors expressively emphasize that:
It should be noticed that not only a rebuilding or a mere modification of the conventional
analysis emerges here, but that a true enlargement is also produced.

Right at the start, they showed that their new analysis incorporates the species of
Dirac’s delta-functions as true “functions”, without the necessity of modifying the
notion of a function. As an example of a delta-function, they presented

1 
δ(x) = · .
π 1 + 2 x 2

The typical features of a delta-function are easily confirmed: δ(x) ≈ 0, except when

+∞
x ≈ 0—where δ(x) is infinitely large, and δ(x) dx = 1.
−∞

The Foundation in the Year 1958

Laugwitz formulated some of Schmieden’s principal ideas in the language of the


then fashionable algebra. He referred to the definitions of Cantor. Cantor had
defined convergent sequences of rational numbers as novel numbers, and he had
named these sequences “limits” (p. 191). The duo Schmieden/Laugwitz proposed
to take the totality of sequences of rational numbers and to define EACH of these
sequences as an individual “number”—independent of their convergence, just: all
these sequences! As a name for the sequences, they chose “Limit” (as Cantor did,
but with a capital “L” as a distinction) as well as “-number”.

(an )n = a = Lim an ,
n=

especially (n)n =  , ( n1 )n = ω and also ((−1)n )n = (−1) .

That is to say, Schmieden’s “jump” from the process of counting 1, 2, 3, . . .


to  (see p. 226) now obtains as its final destination the complete sequence  =
(n)n ! Even this divergent sequence defines an -number, of course an infinitely
large one.
The preceding number to , i.e. −1, is the sequence (n−1)n = (0, 1, 2, 3, . . .);
its successor, i.e.  + 1, is the sequence (2, 3, 4, . . .), etc.
The First Formal Version of a Nonstandard-Analysis in the Year 1958 235

The discombobulating facts in favour of conventional analysis (that the new


numbers comprise zero divisors and therefore do not constitute a field—and
consequently division is not always possible—as well as the missing of their “total
order”) were clearly articulated in this chapter, but not a single word of valuation
was added.
For example, we have

(0, 1, 0, 1, 0, 1, . . .) · (1, 0, 1, 0, 1, 0, . . .) = 0 ,

although neither factor being 0 = (0, 0, 0, . . .). And, moreover, we are not able to
call one of these two factors greater than the other—although obviously they are
different.

Further Peculiarities of the New Analysis in the Year 1958

The authors also clearly identified those facts that appear in this new “enlargement”
of analysis differently from in conventional analysis:
• There are three different equivalence relations as orders of magnitude:
1. “finitely equal”, 2. “having the same magnitude” and 3. “having the same
order of magnitude”.
• Some theorems are stated and proved, which are wrong in conventional
analysis but true in the new enlargement.
1. Every limit exists. This is just the new definition of “number”.
In detail (in the case of sequences of -numbers, we write their “index”
(p)
above): The limit of a sequence of -numbers (a )p is defined for each
“positive integer” as well as for each “infinitely large” -number g , i.e.
g = (gn )n where gn > gk > 0 if n > k. The limit

(p) (g )
Lim a = a  = b
p=g 

is defined as the sequence of components


(gn )
bn = an ,

(p)
i.e. by the “diagonal sequence” defined from the sequence (a )p of -
(p)
numbers
   of the constant sequence a =
a .—Examples: (a) In case
(p)
1
n n = ω, we have lim a = n1 = a = ω, as it should be;
 p= n
(p) (p)
and lim a = 2n 1
= 12 ω. (b) The sequence b with the pth term
p=2 n
236 14 Analysis with or Without Paradoxes?

   
(p) (p) (p)
b = 1
p·n n is lim b = 1
n·n n = 1
· = ω2 and lim b =
  p= p=2

2n·n n = 2· = 12 ω2 —right?


1 1

2. Any two limits are interchangeable. The essential reason for this is that
limits are only indicated but not calculated.
p,q g ,h
In detail: Given a double sequence (a )p,q of -numbers an n , the limit
g ,h
is given by the sequence of the components ann n .
Example:

1
a p,q = p .
1+ q

(The given -numbers are constant sequences of rational numbers with the
constant components 1+1p/q .)
Conventionally, we have

lim lim a p,q = 0 , but lim lim a p,q = 1 .


q→∞ p→∞ p→∞ q→∞

However, for -numbers, independently from the order, we have

1
Lim a p,q = .
p=g
q=h
1 + hg

If you choose p = g (conventionally: p → ∞) for taking the first limit,


you obtain

1
g ,
1+ q

i.e. an infinitely small number for each finite q = 1, 2, 3, . . . Yet, if you


then choose for the other limit h = g , you nevertheless get the finite
value

1
g = 1
g = 1
1+1 = 1
2 ,
1+ h 1+ g
 

1
which says that the infinitely many small numbers g nevertheless have
1+ q
a finite limit of the value 12 .
3. Divergent series are of equal rank to the others, as in the new version of
analysis the calculations are the same: in the finite and in the infinite.
As an example, the famous series:


  
  1
+1
1 1 1 1 1
= − = − =1− ≈ 1.
p(p + 1) p p+1 p p +1
p=1 p=1 p=1 p=2
Finale 237

A convergent series on the left is split into the difference of two divergent
series, and then this difference is calculated to be a finite value. Conven-
tionally, an absolute no-go!
4. The limit function of a sequence of continuous functions is continuous. The
authors’ comment: “This theorem has no simple analogy in conventional
analysis”. There is no reference made to the “Cauchy Sum Theorem”
(p. 138). As an example, the sequence of functions x n , for the closed
interval 0  x  1, is given (already treated above on p. 140).
• The Weierstraß Approximation Theorem was also obtained for the new version
of analysis.
• Differentiation and integration were briefly mentioned. However, a “differ-
entiable” function does not need to have a derivative for each value! (“The
‘derivative’ might be ‘irrational’”.)
• The Mean Value Theorem is proved (for “normal” functions).
Naturally, the notion of “function” is precarious. Of course, complete freedom in
the sense of Bolzano (p. 111) could not be permitted. After all the starting point of
Schmieden/Laugwitz is conventional analysis (which they call “usual”)—and not a
completely new, extravagant shape of analysis. That is why they restrict themselves
to such functions that are “already completely defined, if their values are given for
the rational numbers of the domain”. Sensibly, the description of the values of the
function has to be given in a standardized mathematical language.

Finale

Curt Schmieden was a practitioner of calculating. As a consequence, he had the idea


of practicing analysis by calculating: no complicated general theorems that state
the rules of calculation or rule out certain techniques of calculation—just calculate
and, besides, treat the infinite (i.e. the specifically analytical) in just the same way
as the finite. (He expressed it this way: “calculating analysis”.) The result of the
calculation has to be assessed according to the requirements of the problem at the
end of that calculation.
That this is permissible was shown to him by his calculations (“exactly as in
the finite”—what should be wrong with this?) as well as by his results. To anchor
his technique of “Omega Analysis”—an enlarged version of “Value Analysis”—
reliably within the current foundational principles of mathematics, he left to others.
His enthusiastic colleague and co-worker Detlef Laugwitz gained credit for his
precise algebraic foundation of Schmieden’s ideas. This was first published in the
most general version in an article from 1958, which was signed by both of them.
In 1978, Laugwitz published a stronger algebraic construction, developed from
Schmieden’s essential ideas, and in 1986, he presented a new version, which
incorporates some features of formal logic.
238 14 Analysis with or Without Paradoxes?

Laugwitz was always aware that Schmieden’s -numbers are not a field because
one cannot divide by all non-zero numbers. This seemed to be a problem, especially
at a time, when all mathematics was dominated by Nicolas Bourbaki’s idea of
structure. Subsequently, in 1978, Laugwitz constructed, following others, a field
∗K by algebraic means. Later, in 1986, Laugwitz pretended to, but did not clearly

define a set K with -numbers. Instead of defining this, he changed to a formal


language and developed a theory where “the rules of an ordered field” are valid.
These attempts to represent Schmieden’s ideas within the actual methodology of
calculus could not do justice to Schmieden’s fundamental concerns. In a field that
includes -numbers, it must be settled, whether /2 is a natural number or not. The
mathematician might lack the means to decide which it is—but it will definitely be
one of the two. But this does not fit to Schmieden’s essential idea. Schmieden really
left open both possibilities—and if it were necessary, he distinguished both cases.
Methodical stipulations of this kind—even more if they were bound to be unknown
in every detail—were surely not his aim.

Foundational Problems

This first publication of a nonstandard analysis by these two men was soon followed
by versions from other authors.
The first was by Abraham Robinson (1918–74), a model theorist who also coined
the name “nonstandard-analysis”. This approach to nonstandard analysis requires
the knowledge of logics as a precondition to deal with continuous functions.
Willem A. J. Luxemburg (1929–2018) confined himself to algebraic construc-
tions, essentially relying on ultra-filters for his definition of a field with “infinite”
numbers.
In 1977, Edward Nelson (1932–2014) published a first axiom system for non-
standard analysis. Other versions followed.
This means that another mathematical theory (logics, universal algebra) is
always needed to provide nonstandard analysis with suitable “numbers”. This is
not satisfactory and is also a severe obstacle for an easy acceptance of this novel
approach to analysis, unless you simply confine yourself to an axiom system.

Axiomatics

When inspecting the actual textbooks of analysis, it appears that there is no adequate
consideration paid to the foundations. Generally, the real numbers are not introduced
constructively, as was taught by Müller, Bertrand and Dedekind or by Cantor and
Heine, but instead, following Hilbert, axiomatically. This helps to save time.
Finale 239

Well, possibly, this fits excellently to the new role of mathematics as a security
police for all sciences, which was promulgated by Hilbert in 1917 (p. 209) in order to
teach the unconditional (at least provisional) acceptance of arbitrary axiom systems
detached from any informal substantial justifications.
If it is so, there is then no argument to be put forward against the idea of
taking an axiom system for nonstandard numbers (today usually called “hyper-
real numbers”)—possibly apart from the fact that the elaboration of a nonstandard
analysis differs in some respects from standard analysis. (This has been shown
with the help of some “paradoxes” as well as generalized: see from p. 235.) This
challenges one’s independent thought as well as impedes the usage of textbooks:
you always have to check which version of analysis the chosen author likes.
Standardization of the curriculum facilitates lecturing, especially if many students
have to be educated. In small circles of specialists, subtle discussions will be eased.

A Path to Independency

Schmieden’s intention had been to rescue conventional analysis from its “para-
doxes” as well as to give it a better form: fewer theorems, more computations.
An opposite development started in the 1980s. It continued for some decades in
small circles. Under the (for the laymen astonishing) title “constructive nonstandard-
analysis”, the theory was developed further as a special field in its own right in which
more advanced methods (like sheafs and topoi) were implemented. This leads to a
departure from the first origins of nonstandard analysis as well as from classical
calculus, but it undoubtedly created some marvellous mathematics.
After the discovery of Weierstraß’ construction of the real numbers in 2016, it
may well be possible to develop a new kind of approach to nonstandard analysis.

Nonstandard-Analysis and the History of Analysis

Not Schmieden and Laugwitz, but Robinson was the one who immediately started to
relate the new theory to historical texts. The theorem nowadays called the “Cauchy
Sum Theorem” (p. 138) came up fairly early for discussion. In his book from 1963,
Robinson gave an interpretation of Cauchy’s writings. His result was: Cauchy’s
theorem is correct if we add one of the two additional assumptions: (a) the series is
uniformly convergent or (b) the family (sn (x))n of partial sums is equicontinuous in
the interval.
Robinson’s idea of translating historical mathematical texts into the
language of Nonstandard-analysis impressed the philosopher of science Imre
Lakatos (1922–74). In a lecture, which appeared in print only after his death,
he claimed that “Cauchy’s theorem was true and his proof as correct as an informal
proof can be”. According to Lakatos, the Cauchy Sum Theorem does therefore
240 14 Analysis with or Without Paradoxes?

not need any additional assumptions to gain validity. Thus Lakatos contradicted
Robinson.—The further development (i.e. that Lakatos was right in quite another
sense from what he thought) was traced on pp. 130f.

A Satisfying Finish

It was the aim of my dissertation of 1981 to substantiate Lakatos’ thesis. In contrast


to him, I based my arguments on the Darmstadt version of Nonstandard-analysis.
Today I am aware that my approach (see p. 132) was mistaken.
My doctor father Laugwitz was intrigued by my work, and it inspired him to
undertake his own detailed studies of former analysts, esp. of Cauchy’s works,
thereby producing quite a few articles on the topic.
My studies of Bolzano’s mathematics, which started in 1986, showed me the
inadequacy of Lakatos’ methodology in regard to the understanding of historical
mathematical writings. As a consequence, I had to change my approach. In winter
1990, I turned again to the study of Cauchy’s analysis, this time from a new
perspective. Prof Dr Laugwitz tried to hinder my work with a maximum of vigour.
After he failed, I lost his favour and consequently all local academic support (and
sadly, more than that). The Darmstadt University showed itself incapable of an
unbiased clarification of this issue until today.
If Prof Dr Laugwitz had been factually successful at his time (not only institu-
tionally), this book could not have been written.

Literature

Aigner, M. & Ziegler, G. M. (1998). Proofs from the book. Berlin: Springer.
Lakatos, I. (1980). Philosophical papers (2 vols). In Mathematics, science and epistemology
(Vol. 1). Cambridge: Cambridge University Press 1980.
Laugwitz, D. (1986). Zahlen und Kontinuum. Eine Einführung in die Infinitesimalmathematik.
Mannheim: Bibliographisches Institut.
Luxemburg, W. A. J. (1962). Non-Standard Analysis. In Lectures on A. Robinson’s Theory of
Infinitesimals and Infinitely Large Numbers. Lecture notes. Pasadena, California: Mathematics
Department, California Institute of Technology.
Luxemburg, W. A. J., & Körner, S. (Eds.) (1979). Selected papers of Abraham Robinson (Vol. 2).
Amsterdam, New York, Oxford: North Holland Publishing Company.
Nelson, E. (1977). Internal set theory: A new approach to nonstandard analysis. Bulletin of the
American Mathematical Society, 83(6), 1165–1198.
Palmgren, E. (1995). A constructive approach to nonstandard analysis. Annals of Pure and Applied
Logic, 73, 297–325.
Palmgren, E. (1996). Constructive nonstandard analysis. Cahiers du Centre de logique, 9, 69–97.
Palmgren, E. (1997). A sheaf-theoretic approach for nonstandard analysis. Annals of Pure and
Applied Logic, 85, 69–86.
Palmgren, E. (1998). Developments in constructive nonstandard analysis. The Bulletin of Symbolic
Logic, 4(3), 233–272.
Literature 241

Palmgren, E. (2001). Unifying constructive and nonstandard analysis. In P. Schuster, U. Berger,


& H. Osswald (Eds.), Reuniting the antipodes—constructive and nonstandard views of the
continuum 2001 (pp. 167–183).
Richman, F. (1998). Generalized real numbers in constructive mathematics. Indagationes Mathe-
maticae, N. S., 9(4), 595–606.
Riemann, B. (1854). Ueber die Darstellbarkeit einer Function durch eine trigonometrische Reihe.
cited from Weber and Dedekind 1953, pp. 227–271.
Robinson, A. (1961). Non-Standard analysis. Indagationes Mathematicae, 23, 432–440. cited from
Luxemburg and Körner 1979, pp. 3–11.
Robinson, A. (1963). Introduction to model theory and to the metamathematics of algebra.
Amsterdam: North-Holland Publishing Company.
Russell, B. (1919). Introduction to mathematical philosophy. In Georg Allen & Unwin, London
(New York: The MacMillan Co.). www.gutenberg.org/ebooks/41654.
Schmieden, C. (1948–53). Vom Unendlichen und der Null. Versuch einer Neubegründung der
Analysis. manuscript, unpublished.
Schmieden, C. & Laugwitz, D. (1958). Eine Erweiterung der Infinitesimalrechnung. Mathematis-
che Zeitschrift, 69, 1–39.
Schuster, P., Berger, U., & Osswald, H. (2001). Reuniting the antipodes: Constructive and
nonstandard views of the continuum. Dordrecht: Kluwer.
Schuster, P. M. (2000). A constructive look at generalized Cauchy reals. Mathematical Logic
Quarterly, 46, 125–134.
Spalt, D. D. (1981). Vom Mythos der Mathematischen Vernunft. Darmstadt: Wissenschaftliche
Buchgesellschaft. 2 1987.
Spalt, D. D. (1966). Die Vernunft im Cauchy-Mythos. Thun und Frankfurt am Main: Harri Deutsch.
Spalt, D. D. (2022). Die Grundlegung der Analysis durch Karl Weierstraß—eine bislang unbekan-
nte Konstruktion der natürlichen und der reellen Zahlen. Berlin: Springer.
Strauß, E. (1880/81). Weierstrass, Einleitung in die Theorie der Analytischen Functionen. Univer-
sität Frankfurt am Main: Archiv. Sign 2.11.01; 170348.
Wallace, D. F. (2003). Everything and more: A compact history of infinity. New York, London: W.
W. Norton (2010)
Weber, H. & Dedekind, R. (Eds.) (1953). Bernhard Riemann, Gesammelte mathematische Werke.
reprint of the second edition 1892. New York: Dover Publications.
Author Index

A 205, 207, 210, 211, 218, 220, 234, 238,


Abel, Niels Henrik (1802–29), 133, 140, 141, 251
156–158 Carnot, Lazare Nicolas Marguerite
Aigner, Martin, 216 (1753–1823), 115
d’Alembert, Jean-Baptiste le Rond Cauchy, Augustin-Louis (1789–1857), viii,
(1717–1783), xix, 91–94, 101, 110, 120 xvii, xix, xxi–xxiii, 32, 108, 112,
Andersen, Kirsti, 43, 45, 131 115–147, 149–160, 163, 164, 177,
Arbogast, Louis (1759–1803), 83 181, 187, 189, 211, 225, 237, 239, 240,
Aristotle, (−384–−322), 171 248–250, 253, 254
Arnold, Bernd, viii Cavalieri, Buonaventura (c1598–1647), 43–46
Arthur, Richard T. W. , 33 Conway, John Horton (1937–2020), 180, 183
de Coriolis, Gaspard Gustav (1792–1843),
129
B Crelle, August Leopold (1780–1855), 96, 97
Barrow, Isaac (1630–77), 66
Bernoulli, Jacob (1654–1705), 32, 51
Bernoulli, Johann (1667–1748), xvii–xix, 32, D
39, 40, 51–66, 69, 70, 72, 79, 80, 83, 92, Dedekind, Richard (1831–1916), xix–xxii,
93, 95, 120, 132, 177, 194, 210, 220, 173, 177, 189, 193, 199–205, 207, 209,
226, 229, 249 238, 251
Bertrand, Joseph Louis François (1822–1900), Descartes, René (1596–1650), xviii, 1–22, 36,
xx, 189, 201–203, 205, 209, 238 39, 43, 48, 59–62, 71, 72, 81, 92, 123,
Björling, Emanuel G. (1808–72), 155, 156 211
Bolzano, Bernard (1781–1848), xix, xxi, Diderot, Denis (1713–84), 91
101–113, 120, 122–124, 128, 129, 146, Dirac, Paul Adrien Maurice (1902–1984),
155, 164, 168, 170, 187, 188, 237, 240, 234
248, 249, 254 Dirichlet, Johann Peter Gustav (Lejeune)
Bos, Henk, 20, 33, 131 (1805–1859), xix, 153, 154, 156,
Bourbaki, Nicolas, xx, 178, 207, 238 168–170, 248, 249, 254
Breger, Herbert, 34

E
C Elias, Norbert (1897–1990), 88
Cantor, Georg (1845–1918), xvii, xix–xxi, Euclid, (c−300), xv, xvii, 4, 10, 13, 16, 18, 27,
61, 160, 171, 173, 189–200, 204, 41, 44, 81, 92

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 243
D. D. Spalt, A Brief History of Analysis,
https://doi.org/10.1007/978-3-031-00650-0
244 Author Index

Euler, Leonhard (1707–83), xvii, xix, 53, Kneser, Adolf (1862–1930), 174, 177
69–88, 91–98, 101–103, 109–111, Koenigsberger, Leo (1837–1921), 169, 170
116–119, 123–125, 127, 128, 131,
132, 136, 143, 149–152, 154, 177, 194,
210, 211, 217, 226, 227, 232, 248, 249, L
251–253 Lacroix, Sylvestre François (1765–1843), 118
Lagrange, Joseph Louis (1736–1813), xix,
94–98, 101, 116–118, 131, 136, 142,
F 143, 149–151, 249, 254
Foster Wallace, David (1962–2008), 216 Lakatos, Imre (1922–74), 132, 134, 239, 240
Fourier, Jean Baptiste Joseph de (1768–1830), Laugwitz, Detlef (1932–2000), xx, 130, 132,
118 144, 160, 161, 163, 164, 197, 212, 229,
Franz I. (II.), emperor (1804–1835), 101 233, 234, 237–240
Frege, Gottlob (1848–1925), 193, 195, 196, Leibniz, Gottfried Wilhelm (1646–1716),
198, 202, 208, 209, 227 xvi–xix, 13, 21–36, 39, 40, 42, 43, 45,
46, 48, 49, 51, 53–60, 63, 65, 66, 69, 70,
73, 78–81, 83, 84, 92, 107, 120, 123,
G 126, 132, 145, 146, 171, 177, 195, 210,
Galilei, Galileo (1564–1642), 2, 43, 45 211, 220, 248, 249
Gauß, Carl Friedrich (1777–1855), 153 Leonardo of Pisa (c1170–c1250), 210
Givsan, Hassan, viii Lorenzen, Paul (1915–94), 144
Grüson, Johann Philipp (1768–1857), 94 Louis XVIII, King (1814/15–1824), 115
Gudermann, Christoph (1798–1852), 158 Luxemburg, Willem A. J. (1929–2018), 238

H M
Hamborg, Otto, 5, 22 Marquis de Condorcet, Marie Jean Antoine
Hankel, Hermann (1839–73), 188, 191 Nicolas de Caritat (1743–1794), 94
Hausdorff, Felix (1868–1942), 203 Mayer, Johann Tobias (1752–1830), 120
Hegel, Georg Wilhelm Friedrich (1770–1831), Menaechmus, (mid −4th cent), 16
34 Monge, Gaspard (1746–1818), 115
Heine, Heinrich Eduard (1821–81), 189, 192, Müller, Traugott (1797–1862), ix, xx, xxii, 189,
193, 197–199, 205, 207, 238 200–205, 209, 238, 251
Hilbert, David (1862–1943), xx, 179, 180, 188,
205–209, 238, 239, 247, 251 N
Hobbes, Thomas (1588–1679), 11 Napoleon I. (1804–14/15), 115
Hölder, Otto (1859–1937), 204 Nelson, Edward (1932–2014), 238
Horn, Juliane, viii, ix Newton, Isaac (1643–1727), xviii, 22, 40,
Horowski, Leonhard, 88 46–48, 94
l’Hospital, Guillaume François Antoine, Nicholas of Cusa (1401–64), 42, 43
Marquis de (1661–1704), 65, 66 Nieuwentijt, Bernard (1654–1718), 65, 66
Humboldt, Alexander Freiherr von
(1769–1859), 153
Hus, Jan (c1370–1415), 102 O
Husserl, Edmund (1859–1938), 174 Oldenburg, Reinhard, viii

J P
Jacobi, Carl Gustav Jacob (1804–51), 153 Pappos (4th cent), 19
Jungius, Joachim (1587–1657), 28 Pavić, Marco, viii
Peano, Guiseppe (1858–1932), 144
Plato, (−427–−347), 16
K Protagoras (c−480–c−421), 43
Kästner, Abraham Gotthelf (1719–1800), 151 Pythagoras(c−580–c−500), xvii
Author Index 245

R Strauß, Emil (1859–92), 173, 174, 177,


Riede, Harald, viii 178
Riemann, Bernhard Georg Friedrich (1826–
1866), xviii, xix, 32, 112, 154, 155,
T
158–164, 169, 171, 172, 185, 188,
Tannery, Jules (1848–1910), 204, 205
221–223, 249, 254
Thiele, Rüdiger, 72
Ries, Adam (1492–1559), 210
Thomas Aquinas (c1225–74), 42
Robinson, Abraham (1918–74), 130, 134,
Torricelli, Evangelista (1608–47), 45, 46
238–240
Rudolff, Christoff (c1499–c1545), 211
Ruhmann, Iris, viii, ix V
Russell, Bertrand Arthur William (1872–1970), Viète, François (1540–1603), xviii, 5–7
199, 203, 204, 210, 218–220, 222, 251,
252
W
Weierstraß, Karl Theodor Wilhelm (1815–
1897), viii, ix, xvii, xix–xxiii, 61, 155,
S 158, 160–163, 167–186, 188–196, 198,
Schmieden, Curt (1905–91), xx, 130, 226–234, 199, 204–208, 210, 211, 215, 217, 220,
237–239, 251 222, 223, 227, 229, 237, 239, 248, 249,
Seidel, Philipp Ludwig (1821–96), 133, 251–254
156–159 Wolfart, Jürgen, viii
Smith, John D., ix
Sonar, Thomas, 83
Spivak, Michael, 102–106, 122 Z
Stevin, Simon (1548/9–1620), 13, 211 Ziegler, Günter M., 216
Subject Index

Symbols Applicability, 193, 198


+ − . . . , 228 Approximation
indefinite, 120, 200
infinite, 120
A uniform, 172
Accounting (does not need numbers), 119 Arbitrarily
Accuracy (any degree of), 182 chosen value, 221, 222
Activity (internal), 21 large, 106
Addition, 3, 175 many, 106
Algebra, 65, 81, 87, 93, 178, 234 small
universal, 238 changes will become, 161
All the lines (indivisibles), 43–46 differences can be made, 104
Analogy of convergence and continuity, error, 26
107–108, 110–111, 129 terms will become, 23, 102, 162
Analysis, 5, 14 values will become, 78, 139
Algebraic, 87, 91–98, 101, 116, 117, 145, Area, 3, 28, 29, 31, 32, 44
149–151, 210 Arithmetic, 4, 8, 12, 16, 19, 39, 79, 81, 86,
as freestyle wrestling, 152–160 176–178, 191, 193, 194, 197, 198,
Fourier, 82 200, 201, 211
Functional, 185 Arithmetical operations, 194, 203, 205
geometrical foundation of, 93 Axiom
logical, 205 of Archimedes, 206
Nonstandard-, xvii, xx, 130–135, 149, 150, of completeness, 206, 207
226–240 Axiomatization of the real numbers by Hilbert,
constructive, 239 205–207
-, 226–237 Axioms, 208
philosophical style of, 199 of arithmetical operations, 206
of the infinite, 65, 70 of continuity, 206, 207
of the infinitely small, 65
of values, 117, 118, 124, 130, 131, 151,
152, 170, 185, 188, 210 C
standard-, 131, 132, 150, 158, 239 Calculatory expression, 72–74, 81–88, 97,
of indivisibles, 40, 42 108–112, 118, 129, 154, 185

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 247
D. D. Spalt, A Brief History of Analysis,
https://doi.org/10.1007/978-3-031-00650-0
248 Subject Index

Calculus, 151 (un)conditional, 184, 191, 221–223,


differential, 70, 74, 75, 86, 87, 95, 96 229–231
integral, 78 continuous, 139, 144, 159
of expressions, 117, 118, 124, 127, 131, 147 definition wrongly formulated?, 160–164
of values, 117, 128, 131, 132, 147 domain of, 171
Cartesian dualism, 8 equicontinuous, 239
Change, 21–23, 26 by Euler, 84–86
increase or decrease, 118 by Leibniz, 23–27
regular, 222 steady, 158
structural, 117 uniform, 133, 157, 158, 171, 239
sudden, 108, 109, 153 usual, 102–105
of magnitude, 118 criterion from Leibniz, 120
of theory, 117, 172 by Weierstraß?, 180–185
Choice (arbitrary), 222 of function series, 136, 224
Circle, 2, 16, 57, 120 Coordinates (axes, system), 15, 17, 31, 33, 35,
intersection of, 16 36, 83, 108, 110, 126
Class, 202–204, 218, 219 origin, 15
sub-, 218, 219 Correlation (one-one), 218, 219
Compact, 185 Correspondence (definite), 154
Completeness, 179, 185 Cut, 205, 210
Concepts, 208 multiplication of, 202
framework of, 209 is a number, 201
Cone, 16, 19
Constant, 63, 70, 71
Construction D
conceptual, 104, 119, 143, 145 Decimal (comma, places, point), 211
geometrical, 18, 60, 189 Decrease below any given quantity, 36
set-theoretical, 218 Decrease indefinitely, 24
Content (mathematical), 172 Dedekind-cut, 202, 201–204
Continuity, 16, 20, 21, 35, 93, 108, 110, 120, Definite, 70, 71, 92
199 Definition
by Bolzano, 108–111 not completely established, 189
by Cauchy, 125, 128, 123–135 substancial, 171
by Dirichlet, 154 theory of, 171
disruption of, 153 Derivability
essence of, 199 continuous, 144
by Euler, 80, 81, 84 free, 144
global, local, 110 Derivation, 93
in no interval, 112–113 by Cauchy, 143, 144
one-sided, 109 Derivative, 96, 102, 143, 169, 170, 237
thoroughly, 154 Determined, 57
of arbitrary functions, 94 law, 136
of the sum of a series of functions, 139 Diagonal Argument, 215–221
Continuum, 36, 37, 40–46 Differentiability, 143
Contradiction, 27, 55 nearly everywhere, 113, 170
free from, 60 nowhere, 170
Convergence, 22, 23, 26, 80–81, 84, 102, Differential, 32, 33, 59–66
190–192 equation, 32
absolute, 184, 229 quotient, 95, 143, 169, 170
arbitrarily slow, 157 first, second, . . . , 96
by Bolzano, 105–107 Directed length, 13, 27, 195
by Cauchy, 125, 135–136, 136–140, Discontinuities, 169
140–142, 157 Discrete, 20, 107, 110, 125
Subject Index 249

Division F
concept of, 198 Fact (mathematical), 192
points of, 145 logical relations of, 209
values of, 145 of intuition, 191, 208
of lines, 4 Fallacy, 176, 194
of the interval, 145 False theorem, 170
Doctrine (main of analysis), 169, 170 Field, 179, 180, 186, 235, 238
Doctrine of Coincidence, 43 ordered, 238
Finite being, 92
Fluent, fluxion, 48, 94
E Formalism, 193, 199
Element, 198 Formula, 1–10
Ellipse, 16, 37 Formulae (Viète’s), 7
Enlargement of analysis (calculus), 130, 234, Freedom, 196
235 Function, 69
Entity (exotic), 134 by Bolzano, 111
Epsilontics, 159, 159–195 capricious, 168–170, 185
Equality, 44, 60–63, 195, 227 by Cauchy, 122–123
definition of, 189, 197 characteristic, 150, 168–169
essence of, 196 compound, 93
is identiy, 195 continuous, 238
is not identity, 60 definite, 154
laws of, 195 delta-, 233, 234
meanings of, 196 differentiable, 237
sign, 5, 9, 63–65, 93, 195 by Dirichlet, 170
second, 61, 62, 64, 120 discontinuous, 157, 169
substancial concept of, 185 essence of, 172
of different things, 195 by Euler, 71
of real numbers, 192, 196, 197 Euler’s second notion, 82
(by Weierstraß), 177 exceptional, 170
by Weierstraß, 175 exponential, 76–79, 232
Equation, 93 by Johann Bernoulli, 70
adjoining domain, 93 by Lagrange, 94
content of, 193 lawless, 172
formal, 12 by Leibniz, 31
as mathematical object, 9 limit, 237
product of, 12 normal, 237
Equivalence relation, 62, 179, 235 one-to-many, 97
Error one-to-one, 97, 154
caused by breaking off, 26 point of, 153
infinitely small, 78 representation of a, 172
partial, 29, 30 by Riemann, 171
total, 30, 31 sine, 126
of approximation, 29 series, 224–225, 231–233
of argumentation, 133, 135, 142, 157 by Weierstraß, 171, 171–172
Essential, 41, 42
component of a theory, 193
property, 188
Euclidian theory, 7 G
Exception Genius (flash of), 200
from the rule, 72, 75, 82 Geometry, 2–4, 7, 8, 59, 81, 205
proves the rule, 98 non-Euclidian, 150
without, 92 spherical, 208
Extension, 8 Group, 178, 185
250 Subject Index

H L
Hyperbola, 16 Law
mathematical, 79
of calculation, 28, 33, 61, 63, 198
I for negative numbers, 79
Identity, 60, 61, 195 of Continuity, 35, 35, 36, 43, 55, 59–65
Imagination, 66 of dependency, 111, 112
Inconsistant way of thinking, 80 of equations, 7
Increase, 66 of equality, 177
infinitely small, 143 of homogeneity, 6, 7, 9, 11, 92
Increment (negative), 109 of Noncontradiction, 43
Indivisible, 42–46 of thought, 7, 34, 35, 193
Inference from the finite to the infinite, s
55 of arithmetic, 177, 193, 205
Infinite, 55, 169 of computation, 175, 211
actual, 211, 217 of differential calculation, 60–65
mathematical, 57 Length (of straight lines), 195
numerical, 55 Limes calculation, 159, 159–164
representation of the in the finite, lim-operator, 158–160
226 Limit, 94, 103, 119, 143, 144, 158, 158–164,
Infinitely little, 76 191, 195, 201, 203, 227, 234–236
Infinitely many (∞), 217 (as a name), 197
Infinity, 85, 92, 217, 220 by Cauchy, 120, 121–122, 124, 124, 125,
(∞) is a value, 80, 125 128
(∞) is not a true number, 177 definite, 194
(∞) is no value, 211 existence of, 143
actual (acceptance of), 209–210 lower, upper, 203
existence of, 51–58 unique, 125, 129, 136
is not a number, 56, 125 s of the theory, 169
uncountable (existence of), 215 Line, 7, 11, 16, 20, 39, 40, 44
Integral (definite), 31, 32, 102, 118, 145, admissible, 18
169 continuous, 16
Cauchy’s, 145–147 curved, 17–19, 30, 31, 33, 35, 65, 81–84,
existence of, 147 93, 109, 110, 145, 171
Fourier’s, 118 directed, 12–13
Leibniz’, 28–32 epressed by an equation, 18
Riemann’s, 32, 169 infinitely small, 46
sign, 32 length of, 15, 36, 39, 81
Interval, 113, 118, 123, 126, 146 moved, 17
closed, 172, 237 principal, 14, 15, 17
divisions of, 145, 146 whole, 40
finite, 117 of intersection, 16
sub, 211 Lines (calculations with), 3
Intuition, 3, 9, 188, 191, 198, 208, 226 Love, law, chimney-sweeps, 208
(about infinity), 226
counter-intuitive, 170
spacial, 205 M
Magnitude, 92, 118, 119, 199
Majority, 131, 157
J Mathematics, 27
Jump (geometrical), 109, 168, 169 essence of, 196
Jump (numerical), 226, 234 formal, 193
Jump (in values), 153 making, 196
Subject Index 251

relational and substancial, 171, 172, 177, is no multitude, 56


185, 192, 208 largest, 53, 55
relocation of, 209 pi, 121, 200
a science of things, 209 Numbers
secure police for sciences, 209, 239 binary, 218
structural, 62, 179–180, 207, 208, 238 cardinal, 220
unity of, 151, 193, 205 infinite, 220
Matter, 8 complex
Meaning, 116 component of, 73, 97
Mean value, 153 (today), 73, 79, 173
Measure, 92, 119, 187, 195 (by Weierstraß), 206
Method, 2, 9, 14, 179, 224 concept of, 119
axiomatical, 208 decimal, 57–58, 76, 168, 175, 210, 211,
for division, 179 215
of calculation, 31 digit, 168, 217
of fluxion, 46–49 generalization of, 174–175
of indivisibles, 40 operations with, 179
of proof, 220 division in two classes, 203, 203, 209
of trial and error, 179 fractional (by Weierstraß), 175
Mind, 8 transformation of fractions, 176
Minority, 134 hyper-real, 134, 239
Model, 193 by Luxemburg, 238
theory, 238 by Schmieden, 226–227
Moment (present), 42 infinite, 211, 220, 222, 238
Monad, 21–22 for calculation, 220, 221
Monoid, 178, 185 existence, 56
Motion, 8, 16, 43, 118, 189 representation of, 211
Multiplication, 175 irrational, 189, 200, 201, 202
of equations, 9 by Bertrand, 201–202
of infinitely large (small) numbers, 77 by Euler, 79
of lines, 3, 4 existence, 189
Multitude, 92 as limit, 121–122
infinite, 56 by Müller, 201–202
by Russell, 203
theorems on (by Weierstraß), 182–183
N natural
Nature, 37, 43 infinitely large, 226
Necessity (metaphysical), 21 set of, 107
Nothing, 47, 63 by Weierstraß, 174
Nought, 61, 66 negative, 3, 27, 28, 36, 79, 109, 191, 192,
for human imagination, 66 200
Number, 3, 8, 20, 199 -, 226–237
arbitrarily large, 125 infinitely near, 227
arbitrarily small (given), 159 infinitely small, 226, 229
assignate a to, 60 ω-sphere, 227
decimal (place of the digit), 168 orders of magnitude, 235
Euler’s, 77 rational, 227
finite, 177 zero, 226
infinite, 51–56, 177 positional system of, 186, 210
natural, 79, 80, 177 real, 188–189
infinitely (large), 77, 78, 198, 226 (assessment by Tannery), 204–205
infinitely small, 76, 132, 134, 140, 194, by Cantor, 189–196, 234
197, 198 by Dedekind, 199–200, 202–203
is a sign [Zeichen], 192 by Hilbert, 205–208
252 Subject Index

numbers (cont.) arbitrary, 16


rational (by Russell), 203 changing, 36, 94, 154
by Russell, 203–204 constant, 70, 71, 72
the simplest way to construct, 180 decreases indefinitely, 36
by Weierstraß, 173–175 by Euler, 70
standard index form, 76 exponential, 76–79
by Weierstraß, 173–180 extensive, 8, 188
finite in size, 54
fixed, 15
O fluent, 47
Object, 195, 198 geometrical, 39, 59
geometrical, 2, 3 indefinite, 71, 75
infinite, 52 indeterminate, 17
legitimate, 169, 185 infinitely small, 36, 94, 96, 132, 143, 195,
mathematical, 9, 57, 107, 121 196
essence of, 171 actual existence is impossible, 211
principal, 81 infinitely smaller, 60
Ontology, 192 measure of, 92
Operation (direct and inverse), 179, 207 nature of, 70
Order, 192 non-geometrical, 70
of infinite numbers, 227 sufficient large, 14
of the world, 128 taken to be arbitrarily small, 106
Ordered pair, 178, 185 unknown, 5–7
Ordinate, 34, 70, 83, 153, 168 vanishing, 94
Oval line, 18–19 variable for the first time, 26
variable passim, especially, 70, 71, 94, 111,
116, 117, 122, 154, 160, 164
P
Parabola, 16, 35
Paradox, 221–233, 239 R
Part, 41, 53 Ratio test, 78
composed of, 92 Reality, 116, 192, 193, 198, 209
exact, 190, 192 Rectangle, 3, 11, 44, 45
Plane (continuous), 11, 16 Relation
Point, 4, 14–20, 29–31, 33–36, 41, 42, 93, 208 one-to-one, 218
cannot be divided, 41 value-to-value, 143
Polynomials (division of), 74 Revolution (conceptual), 117
Postulate Ring, 178, 179
Johann Bernoulli’s First, 60–65 Root (false, true), 12–14
Parallel, 16 Rule, 52, 88
Power series, 75, 87, 95, 96, 172 of calculation, 57, 72, 79
development of a centred at a value, 171 s in regard to infinity, 226
Principle, 22 s of calculation, 226, 237
of similar triangles, 4, 7 Ruler, 44, 92
Problem (Pappos), 14 and compass, 16
Rules for Differentials, 59–66
division, 64
Q product rule, 33–36, 64, 66
Quality, 21, 41, 43
Quantifier, 106
Quantity, 14, 92, 91–92, 118 S
arbitrarily given, 24, 26, 145 Segment (line), 7, 12–14, 20, 31, 39, 59
arbitrarily large, 86 opposed to, 12, 13
arbitrarily small for the first time, 23 Segment (of numbers), 203, 210
Subject Index 253

Sequence, 52, 103, 107 of continuous functions, 139


Cauchy, 103 of products, 146, 147
definitely diverging, 198 of a sequence, 105
summability of a, 104 of a series
of -numbers (limit), 235 by Cauchy, 136, 137, 138
Series independent from grouping the terms
convergent for the first time, 23 (Weierstraß), 182
convergent (equality of), 197 Summability
divergent, 85, 223, 234, 236 by Weierstraß, 181–185
expansion of the function, 142 C-1 Summation, 223–224, 229–231
exponential, 77, 78, 84 Synthesis, 5, 17–20
for the first time, 23
geometric, 224
harmonic, 78, 85–86, 161–163, 181 T
alternating, 109, 221, 223, 228 Tangent
integration of, 224–225 decrease, increase of, 93
particularly remarkable, 106 s (method of), 31–33
rearranged, 228, 229 Term (infinite), 56, 226–233
regular, 222, 229 Theorem
sum of a for the first time, 84 Binomial, 77
summability, 181 Cauchy’s Sum, 138–140, 140, 150, 152,
Taylor, 75, 95, 96 156, 157, 237, 239
terms (law of formation), 160 Fundamental [of Analysis], 169
trigonometric, 118, 153 Fundamental of Functions, 128, 129, 129,
of continuous functions, 138–142, 152–156
237 Intermediate Value, 146, 211
of fractions, 180 Leibniz’, 23–26
of irrational numbers, 180, 182 Mean Value, 237
by Weierstraß, 180–185 Pythagoras’, xvii
Set, 190, 192, 198, 218, 219 Riemann of Rearrangements, 223
concept of, 173 Weierstraß Approximation, 172, 185, 237
infinite, 210 of Archimedes, 206, 207
ordered, 190 of Pythagoras, 208
sub-, 218, 219 Theorem
Set-theory, 199 correct, 182
Sheafs, 239 false, 133, 138
Sign, 96, 125, 147, 159, 180, 193, 201, proved, but doubted, 189
211 truth of, 189
alternating, 24 s of arithmetical operations, 206
Sign [Zeichen], 191–193, 195, 207 Theory of Analytic Functions, 173
Signification (anarchy of), 164 Thought, 7, 8, 41, 43, 193, 198, 199, 209
Spiral, 18 mental content (Gedankeninhalt), 208
Substance, 8, 21, 22 things of, 198
mathematical, 128, 172 Topoi, 239
simple, 21 Topology, 199
Subtraction Traverse (oblique, straight), 44
of a fraction, 183 Triangle, 2, 8, 46
of irrational numbers (by Weierstraß), auxiliary, 35, 36
179 differential, 35
of lines, 3 inequality, 104
Sum Trick of cancellation, 77, 78
partial, 26, 79, 161, 163, 223, 239 Truth, 3, 43, 131, 209
summable, 230 criteria of, 43
true (by Euler), 86–87 mathematical, 131
254 Subject Index

U by Weierstraß, 179
Unit, 92, 192 Variable, 64, 117
Unity, 4, 6–8, 92, 179, 190 continuous, 109
independent, 122, 137
Vibrating string, 83, 93
V View (mathematical), 54, 58, 189
Value, 9, 10, 70, 71, 80, 97, 116 Volume, 195
all possible, 92
certain, 116
change in, 105
definite, 189 W
of a function, 168 Whole, 41, 53, 56, 58, 110, 198
by Bolzano, 111, 129 uniform, 56
by Cauchy, 124, 124–129, 225 Wholeness (of the opposites), 43
by Dirichlet, 153, 168 World (nature of), 41
by Lagrange, 97 Wrong understanding, 171
(notations), 123
by Riemann, 154, 164, 169
unique, 128, 153–156, 164 Z
negative, 143 Zero, 66, 92, 169, 227
possible, 154 as a limit, 36, 154
of the series, 160 is no number, 119
singular, 170 not as denominator, 194
variable, 36 divisor, 235

You might also like