You are on page 1of 266

J.L.

Berggren

Episodes
in the
Mathematics
of
Medieval
Islam

Second Edition
Episodes in the Mathematics of Medieval Islam
J.L. Berggren

Episodes in the Mathematics


of Medieval Islam
Second Edition

123
J.L. Berggren
Department of Mathematics
Simon Fraser University
Burnaby, BC
Canada

ISBN 978-1-4939-3778-3 ISBN 978-1-4939-3780-6 (eBook)


DOI 10.1007/978-1-4939-3780-6
Library of Congress Control Number: 2016952523

© Springer Science+Business Media New York 1986, 2016


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part
of the material is concerned, specifically the rights of translation, 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, express or implied, with respect to the material contained herein or
for any errors or omissions that may have been made.

Printed on acid-free paper

This Springer imprint is published by Springer Nature


The registered company is Springer Science+Business Media LLC
The registered company address is: 233 Spring Street, New York, NY 10013, U.S.A.
Preface

Thirty years have elapsed since the appearance of the first edition of this book, and
the author has been much gratified by the positive reaction of professional col-
leagues as well as individuals previously unknown to him who have written to
express their enjoyment of the book.
However, in those 30 years much has happened in the study of the history of
mathematics in the medieval Islamic world. Young scholars, who were just
beginning their careers when Episodes was written (or, indeed, were not yet born),
as well as others, have made a number of important sources, together with pene-
trating studies of the same, available in European languages. Moreover, increasing
attention has been paid to the setting of mathematics in medieval Islamic society,
how it was taught, and how its study was fostered.
But, in addition to taking into account the growth of the field there is another
reason for a second edition of this work, and that is that the first edition did not
cover the role of mathematicians in the Western regions of medieval Islam. Shortly
after the book appeared, a colleague in that part of the world wrote and asked why I
had not included material on mathematics in that region. To this very reasonable
question I could only reply, as Dr. Johnson had to a woman who asked him why a
certain word was missing in his great Dictionary, “Ignorance, madam, sheer
ignorance.”
Since that time, happily, the author’s knowledge of the history of mathematics in
the Maghrib and al-Andalus has improved, thanks mainly to the work of scholars in
that region as well as Barcelona and France who have contributed so much in the
past few decades. And I hope that this second edition will go some way to inform a
wider public of activities of mathematicians in Islam’s western regions.
Finally, I have taken the opportunity of a second edition to correct typographic
(and other) errors in the first edition, and I thank readers and colleagues who have
alerted me to problems in the text. I also thank my editors at Springer, Dahlia Fisch

vii
viii Preface

and Elizabeth Loew for their support and patience in waiting for a book that was
promised so many years ago! Special thanks, however, are due to Dr. Petra G.
Schmidl, who, in the course of translating the second edition for a German version
of the book, made so many useful suggestions for improving it, and to Heinz Klaus
Strick, for editorial work and his suggestion of including images he provided of
postage stamps honoring eminent Muslim mathematicians. Special thanks, too, are
due to Drs. A. Djebbar and J.P. Hogendijk, and other colleagues too numerous to
mention, who answered many of my questions.
I dedicate this second edition to my wife, Tasoula Saparilla Berggren, who so
willingly gave me the time and support necessary to produce this book.

Burnaby, Canada J.L. Berggren


Contents

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1 The Beginnings of Islam . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
2 Islam’s Acquisition of Foreign Science . . . . . . . . . . . . . . . . . . . . . . 2
3 Four Muslim Scientists . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.2 Al-Khwārizmī . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.3 Al-Bīrūnī . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.4 ‛Umar al-Khayyāmī . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.5 Al-Kāshī . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
4 The Sources. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
5 The Arabic Language and Arabic Names . . . . . . . . . . . . . . . . . . . . 25
5.1 The Language . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
5.2 Transliterating Arabic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
Excercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2 Arithmetic in the Islamic World . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
1 The Decimal System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2 Kūshyār’s Arithmetic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.1 Survey of The Arithmetic . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.2 Addition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
2.3 Subtraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.4 Multiplication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
2.5 Division . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3 The Arithmetic of Common Fractions . . . . . . . . . . . . . . . . . . . . . . . 38
4 The Discovery of Decimal Fractions . . . . . . . . . . . . . . . . . . . . . . . . 43
5 Muslim Sexagesimal Arithmetic . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.1 History of Sexagesimals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.2 Sexagesimal Addition and Subtraction . . . . . . . . . . . . . . . . . . 50
5.3 Sexagesimal Multiplication . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
5.4 Sexagesimal Division . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

ix
x Contents

6 Square Roots . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
6.2 Obtaining Approximate Square Roots . . . . . . . . . . . . . . . . . . . 57
6.3 Justifying the Approximation . . . . . . . . . . . . . . . . . . . . . . . . . 59
7 Al-Kāshī’s Extraction of a Fifth Root . . . . . . . . . . . . . . . . . . . . . . . 61
7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
7.2 Laying Out the Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
7.3 The Procedure for the First Two Digits . . . . . . . . . . . . . . . . . 62
7.4 Justification for the Procedure . . . . . . . . . . . . . . . . . . . . . . . . . 65
7.5 The Remaining Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
7.6 The Fractional Part of the Root . . . . . . . . . . . . . . . . . . . . . . . 68
8 The Islamic Dimension: Problems of Inheritance . . . . . . . . . . . . . . 71
8.1 The First Problem of Inheritance . . . . . . . . . . . . . . . . . . . . . . 73
8.2 The Second Problem of Inheritance . . . . . . . . . . . . . . . . . . . . 73
8.3 On the Calculation of Zakāt . . . . . . . . . . . . . . . . . . . . . . . . . . 74
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
3 Geometry in the Islamic World . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
1 Geometrical Constructions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
2 Greek Sources for Islamic Geometry . . . . . . . . . . . . . . . . . . . . . . . . 82
3 Apollonios’s Theory of the Conics . . . . . . . . . . . . . . . . . . . . . . . . . 83
3.1 Symptom of the Parabola . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
3.2 Symptom of the Hyperbola . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
4 Abū Sahl al-Kūhī on the Regular Heptagon . . . . . . . . . . . . . . . . . . 87
4.1 Archimedes’ Construction of the Regular Heptagon . . . . . . . . 87
4.2 Abū Sahl’s Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
5 The Construction of the Regular Nonagon . . . . . . . . . . . . . . . . . . . 92
5.1 Verging Constructions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
5.2 Fixed Versus Moving Geometry . . . . . . . . . . . . . . . . . . . . . . . 94
5.3 Abū Sahl’s Trisection of the Angle . . . . . . . . . . . . . . . . . . . . 94
6 Construction of the Conic Sections . . . . . . . . . . . . . . . . . . . . . . . . . 95
6.1 Life of Ibrāhīm b. Sinān . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
6.2 Ibrāhīm b. Sinān on the Parabola . . . . . . . . . . . . . . . . . . . . . . 97
6.3 Ibrāhīm b. Sinān on the Hyperbola . . . . . . . . . . . . . . . . . . . . . 98
7 A Problem in Geometrical Optics . . . . . . . . . . . . . . . . . . . . . . . . . . 99
8 Geometry with a Rusty Compass . . . . . . . . . . . . . . . . . . . . . . . . . . 104
9 Al-Mu’taman b. Hūd’s Book of Completion . . . . . . . . . . . . . . . . . . 111
10 Practical Geometry of Measurement . . . . . . . . . . . . . . . . . . . . . . . . 113
Excercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
Contents xi

4 Algebra in the Islamic World . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121


1 Problems About Unknown Quantities . . . . . . . . . . . . . . . . . . . . . . . 121
2 Sources of Islamic Algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
3 Al-Khwārizmī’s Algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
3.1 Basic Ideas in Al-Khwārizmī’s Algebra . . . . . . . . . . . . . . . . . 124
3.2 Al-Khwārizmī’s Discussion of x2 þ 21 ¼ 10x . . . . . . . . . . . . . 125
4 Thābit’s Demonstration for Quadratic Equations . . . . . . . . . . . . . . . 126
4.1 Thābit’s Demonstration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
5 Abū Kāmil on Algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
5.1 Similarities with al-Khwārizmī . . . . . . . . . . . . . . . . . . . . . . . . 130
5.2 Advances Beyond al-Khwārizmī . . . . . . . . . . . . . . . . . . . . . . . 130
5.3 A Problem from Abū Kāmil . . . . . . . . . . . . . . . . . . . . . . . . . . 131
6 Al-Karajī’s Arithmetization of Algebra . . . . . . . . . . . . . . . . . . . . . . 133
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
6.2 Al-Samaw’al on the Law of Exponents . . . . . . . . . . . . . . . . . 135
6.3 Al-Samaw’al on the Division of Polynomials . . . . . . . . . . . . . 137
7 Al-Samaw’al on the Table of Binomial Coefficients . . . . . . . . . . . . 140
8 Algebra in the Maghrib . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
8.1 Ibn al-Bannā’ on Quadratic Equations . . . . . . . . . . . . . . . . . . 144
8.2 Algebraic Notation in the Maghrib . . . . . . . . . . . . . . . . . . . . . 145
9 ‛Umar al-Khayyāmī and the Cubic Equation . . . . . . . . . . . . . . . . . . 146
9.1 The Background to ‛Umar’s Work . . . . . . . . . . . . . . . . . . . . . 146
9.2 ‛Umar’s Classification of Cubic Equations . . . . . . . . . . . . . . . 147
9.3 ‛Umar’s Treatment of x3 þ mx ¼ n . . . . . . . . . . . . . . . . . . . . . 148
9.4 The Main Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
9.5 ‛Umar’s Discussion of the Number of Roots . . . . . . . . . . . . . 150
10 The Islamic Dimension: The Algebra of Legacies . . . . . . . . . . . . . . 152
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
5 Trigonometry in the Islamic World . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
1 Ancient Background: The Table of Chords and the Sine. . . . . . . . . 155
2 The Introduction of the Six Trigonometric Functions . . . . . . . . . . . 160
3 The Seventh Trigonometric Function . . . . . . . . . . . . . . . . . . . . . . . 163
4 Abū a1-Wafā”s Proof of the Addition Theorem for Sines . . . . . . . . 166
5 Naṣīr al-Dīn’s Proof of the Sine Law . . . . . . . . . . . . . . . . . . . . . . . 170
6 Al-Bīrūnī’s Measurement of the Earth . . . . . . . . . . . . . . . . . . . . . . . 173
7 Trigonometric Tables: Calculation and Interpolation . . . . . . . . . . . . 175
8 Auxiliary Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
9 Interpolation Procedures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
9.1 Linear Interpolation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
9.2 Second-Order Interpolation . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
10 Al-Kāshī’s Approximation to Sin(1°) . . . . . . . . . . . . . . . . . . . . . . . 182
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
xii Contents

6 Spherical Trigonometry in the Islamic World . . . . . . . . . . . . . . . . . . . 189


1 The Ancient Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
2 Important Circles on the Celestial Sphere . . . . . . . . . . . . . . . . . . . . 192
3 The Rising Times of the Zodiacal Signs . . . . . . . . . . . . . . . . . . . . . 195
4 Stereographic Projection and the Astrolabe . . . . . . . . . . . . . . . . . . . 197
5 Telling Time by Sun and Stars . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
6 Spherical Trigonometry in Islam . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
7 Tables for Spherical Astronomy . . . . . . . . . . . . . . . . . . . . . . . . . . . 211
8 The Islamic Dimension: The Direction of Prayer . . . . . . . . . . . . . . 216
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
7 Number Theory and Combinatorics in the Islamic World . . . . . . . . . 223
1 Number Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
1.1 Representing Rational Numbers as Sums of Squares . . . . . . . 224
1.2 Figured Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226
1.3 Magic Squares . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228
2 Combinatorics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
2.1 Enumerating Words of k Distinct Letters in an Alphabet
of n Letters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... 235
2.2 Ibn Mun‘im on Counting Arabic Words
of at Most Ten Letters . . . . . . . . . . . . . . . . . . . . . . . . . . .... 236
2.3 Ibn al-Majdī on Enumerating Polynomial Equations . . . . .... 243
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... 245
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
Chapter 1
Introduction

1 The Beginnings of Islam

The Muslim calendar begins in the year A.D. 622, when Muḥammad fled from his
hometown of Mecca, very near the west coast of the Arabian peninsula, to Medīna,
a city about 200 miles to the north. The doctrines of one God, called in Arabic Allāh
(= The God), which he announced had been revealed to him by the angel Gabriel,
had created considerable dissension in Mecca. This was because Mecca was at that
time a thriving center of pilgrimage whose chief attraction was a shrine called the
Ka‛ba, dedicated to the worship of many gods. Eight years later Muḥammad
returned in triumph to Mecca, an event which marked the beginning of the spread of
the religion of Islam, based on the idea of submission to the will of God, which is
the meaning of the Arabic word Islām.
When Muḥammad returned to Mecca in A.D. 630, and even when he died in
632, Islamic contributions to the sciences lay in the future. The first “worlds to
conquer” were not intellectual but the actual lands beyond the Arabian peninsula,
and the Muslims proved as successful in these physical conquests as they were to
prove later in the intellectual ones. To give a detailed account of the great battles
and the generals whose tactics carried so many days is not our purpose here, and we
can only mention that it was Syria and Iraq to the north of the peninsula, whose
foreign rulers were heartily detested by the local populations, that fell soonest to the
Arab armies and their battle cry, “Allāhu akbar” (God is greatest). By 642, the
conquest even of Persia was complete, and Islam had reached the borders of India.
A few years before this, the general ‛Amr ibn al-‛Aṣ had conquered first Egypt and
then all of North Africa, driving the Byzantine armies before him.
Soon, the new religion had spread from the borders of China to Spain, only to be
stopped in France by the victory of Charles Martel near Tours in 732. Although
much has been made of this battle, it is relevant to recall the words of an eminent
historian, Phillip Hitti, in The Arabs: A Short History, where he says:

© Springer Science+Business Media New York 2016 1


J.L. Berggren, Episodes in the Mathematics of Medieval Islam,
DOI 10.1007/978-1-4939-3780-6_1
2 1 Introduction

Later legends embellished this day …, greatly exaggerating its importance. To the
Christians it meant the turning point in the military fortunes of their eternal foe. Gibbon,
and after him other historians, would see mosques in Paris and London, where cathedrals
now stand, and would hear the Koran instead of the Bible expounded in Oxford and other
seats of learning, had the Arabs won the day. To several modern historical writers this battle
of Tours is one of the decisive battles in history.
In reality the battle of Tours decided nothing. The Arab–Berber wave, already almost a
thousand miles from its starting place in Gibraltar, had reached a natural standstill. It had
lost its momentum and spent itself. Although this defeat near Tours was not the actual cause
of the Arab halt, it does set the farthest limit of the victorious Moslem arms.

The political center of the great empire of early Islam was Damascus, a city so
beautiful that Muḥammad, when he saw it, turned back saying that he wanted to
enter Paradise only once. Here, the caliphs, the succesors of Muḥammad as a
political and military leader, held court. They were members of the Umayyad
family (see Plate 1), but in 750 to the Caliphate was seized by a new, Persian family
known as the ‛Abbāsids, whose power-base was in the Eastern lands.
However, one of the Umayyad princes escaped and fled to northwest Africa,
known in Arabic as al-Maghrib (the Arabic word for ‘west’), whence he invaded
and eventually conquered most of Spain, known in Arabic as al-Andalūs. He took
the throne name ‘Abd al-Rahmān and ruled al-Andalūs from 755–788.

2 Islam’s Acquisition of Foreign Science

With this slight account of the early military and political history of Islam, we may
turn to the beginnings of scientific activity in that civilization, for as early as the time
of the Umayyads in the 730s in Sind (modern Pakistan) and Afghanistan astronomical
treatises, based on Indian and Persian sources, were written in Arabic. Moreover, a
civilization as extensive and sophisticated as that of the Umayyads certainly needed
people who were experts in both computation and geometry. In the earliest times there
is evidence that much of that expertise came from non-Islamic subjects, but G. Saliba
has provided evidence that scientific works were translated into Arabic and studied
early in the eighth century, if not before. And, according to him, it was the Arabization
of the Umayyad administration that led unemployed speakers of Greek and Syriac to
begin translating more advanced works from their languages into Arabic.
The usual account of the acquisition of foreign science in Islamic lands stresses
the role of the ‘Abbasids, beginning with the arrival, during the reign of the
‘Abbāsid caliph, al-Manṣūr, of a delegation from Sind which arrived in Baghdad.
The delegation included an Indian who was versed in astronomy and who helped
al-Fazārī translate a Sanskrit astronomical text. The resulting work is the Zīj
al-Sindhind, which contains elements of many astronomical traditions, including
mathematical methods using Sines.
The caliph al-Manṣūr, who had Baghdad built as his new capital, ordered that the
work commence at a time on 30 July, 762 that his astrologers considered to be
2 Islam’s Acquisition of Foreign Science 3

Plate 1 The Umayyad mosque in Damascus where timekeepers (muwaqqit) such as Ibn al-Shātir
and al-Khalīlī worked. From the minaret the faithful are called to prayer five times a day, which
times are astronomically defined. There is the additional requirement that the one saying the prayer
faces in the direction of Mecca, and the functioning of such a mosque depended on a certain
amount of nontrivial mathematics and astronomy

auspicious. (One of these astrologers was al-Fazārī, mentioned above.) The


astrologers must have done their work well, for Baghdad did indeed flourish, both
as a commercial and intellectual center. Thus, during the reign of Hārūn al-Rashīd,
whose glittering rule (786–809) is portrayed in the tales of The Thousand and One
Nights, a library was constructed, and here one could doubtless find both originals
and translations of scientific works in Sanskrit, Persian and Greek, works whose
contents inspired and instructed the first Islamic scientists. Even more stimulus was
given to scientific activity by the Caliph al-Ma’mūn, who was a patron of a number
of important activities and projects that brought the ancient sciences into Islamic
culture. Among other activities, he patronized one of the first important Muslim
mathematicians, al-Khwārizmī, who dedicated his Algebra to him. He also spon-
sored a geodetic survey to determine the length of a degree of a meridian in order to
4 1 Introduction

find the circumference of the earth. (That the earth was a sphere had been known to
the educated since the time of Plato, early in the fourth century B.C.E.) And he
supported the construction of astronomical instruments and their use in observation
of the motions of the heavenly bodies. In addition, Islam’s acquisition of the ancient
sciences enlisted the efforts and support of many wealthy and influential families—
such as the Banū Mūsā (Fig. 1) and the Barmakids, to name but two.
However, a culture’s acquisition of intellectual material from an alien culture is a
complex process and not accomplished in one place by a few individuals.
And D. Gutas has argued that Muslim acquisition of foreign learning during the
early years of the ‘Abbasid reign was as much due to political and religious
problems facing the early caliphs as it was to simple intellectual curiosity and love
of learning for its own sake.
To say that, however, is not to diminish the achievement of these early caliphs,
East and West, or the influence of what they did. For example, al-Ma’mūn’s ini-
tiatives inspired a similar initiative in the Maghrib, by Ibrāhīm II (reigned 875–
902), one of the last rulers of a pro-‘Abbasid dynasty in the Maghrib, the
Aghlabids. And according to the eleventh-century historian Ṣā‘id al-Andalūsī it was
during the reign of ‘Abd al-Raḥmān that the study of the foreign sciences came to
al-Andalūs, and a number of Umayyad rulers in al-Andalūs after him supported the
sciences. Among them were ‘Abd al-Raḥmān, II (822–852) and, the most famous of
this dynasty, ‘Abd al-Raḥmān, III, whose reign, duriong the 50 years from 912 to
961, created what has been called “the golden age” of his kingdom.
The translators, who were employed by wealthy families as well as by caliphs,
exploited the potential inherent in the Arabic language of expressing subtle varia-
tions on an idea by using a set of more or less standard variations of a basic, root
form of a word. They thereby created what was to become the language of scholars
from the Pyrenees to the borders of China.
To obtain copies of ancient manuscripts patrons sent missions to foreign lands to
purchase copies of their most important works. A good example of the dedicated
search for foreign books is the troubles that one of the ninth-century translators,
Ḥunayn ibn Isḥāq, took to find a copy of a book by a Greek medical writer named

Fig. 1 A stamp from Syria commemorating the three Banū Mūsā, showing them engaged in
geometry and astronomy. (The faces of medieval Islamic mathematicians shown in this and other
stamps are, of course, necessarily the product of the artist’s imagination.)
2 Islam’s Acquisition of Foreign Science 5

Galen. (The word “ibn” in the middle of Ḥunayn’s name is Arabic for “son of,” and
in the sequel we shall follow the modern custom of abbreviating it by the single
letter “b.”) Ḥunayn says first that his colleague, Gabriel, went to great troubles to
find it, and then, according to Rosenthal’s translation, “I myself searched with great
zeal in quest of this book over Mesopotamia, all of Syria, in Palestine and Egypt
until I came to Alexandria. I found nothing, except in Damascus, about half of it
(Galen’s book). But what I found was neither successive chapters nor complete.
However, Gabriel (also) found some chapters of this book, which are not the same
as those I found.” (Four hundred years later, this search for foreign science would
be repeated, but this time it would be Europeans traveling in Islamic lands in search
of precious Arabic scientific manuscripts.)
Having, above, explained earlier what “ibn” in an Arabic name means, and since
we shall now meet quite a number of such names, it seems worthwhile to explain
how one reads Arabic names.
A child of a Muslim family will receive a name (in Arabic ism) like Muḥammad,
Ḥusain, Thābit, etc. After this comes the phrase “son of so-and-so,” and the child
will be known as Thābit ibn Qurra (son of Qurra) or Muḥammad ibn Ḥusain (son of
Ḥusain). The genealogy can be compounded. For example, Ibrāhīm ibn Sinān ibn
Thābit ibn Qurra, carries it back to the great-grandfather. Later in life, one might
have a child and then gain a paternal/maternal name (kunya) such as Abū/’Umm
‛Abdullāh (the father/mother of ‛Abdullāh). Next comes a name indicating the tribe
or place of origin (nisba), such as al-Ḥarrānī, “the man from Ḥarrān.” At the end of
the name might come a tag (laqab), it being a nickname such as “the goggle-eyed”
(al-Jāḥiz) or “the tent-maker” (al-Khayyāmī) or a title such as “the orthodox”
(al-Rashīd) or “the blood-shedder” (al-Saffāḥ). Putting all this together, we find the
name of one of the most famous Muslim writers on mechanical devices had the full
name Badī‛ al-Zamān Abū al-‛Izz Isma‛īl b. al-Razzāz al-Jazarī. Here Isma‘īl, the
son of the rice merchant, has earned the laqab “Badī‛al-Zamān,” i.e., “prodigy of
the Age,” certainly a title a scientist might wish to earn. His nisba, al-Jazarī,
signifies a person hailing from al-Jazīra, the region between the upper reaches of the
Tigris and Euphrates rivers.
In the above explanation of Arabc names, we mentioned “Thābit ibn Qurra,”
who was one of the most important figures in early Islamic mathematics. He was a
protégé of the three brothers known in Arabic as the Banū Mūsā (the sons of
Moses), whom we mentioned earlier. These brothers, who traveled even to the
Byzantine world to buy books, also conducted their own researches on mathematics
and mechanics and were patrons of Thābit b. Qurra from Ḥarrān (the modern
Turkish Diyār Bakr) in northern Mesopotamia.
Thābit, who lived from 836 to 901, had a gift for languages and gave Arabic
some of its best translations from Greek. He was a member of the sect of star
worshippers who called themselves Ṣabians (after a sect sanctioned in Book 2,
verse 63 of the Qur’ān) to escape forced conversion to Islam, for as polytheists,
their religious beliefs would have been abhorrent to Muslims. According to one
account, the Banū Mūsā discovered Thābit’s linguistic talents when they met him as
a money changer in Ḥarrān on their travels and brought him back with them to
6 1 Introduction

Baghdad to work with them in their researches. In addition to his skills as a


translator, Thābit’s talent for mathematics increased the store of beautiful results in
that science, and his talents as a medical practitioner earned him a place of honor in
the caliph’s retinue.
There were a number of other skilled translators in this early period of Islamic
science, who were patronized both by wealthy citizens and rulers. Among them were
Ḥunayn b. Isḥāq, whose search for manuscripts we have already mentioned, as well
as his son Isḥāq b. Ḥunayn, Qusṭā b. Lūqā, a Christian from the Lebanese town of
Baalbek, and al-Ḥajjāj b. Maṭar. The chart below lists some of the Greek mathe-
matical authors and their works that we shall refer to in this book, the names of those
who translated them into Arabic and the approximate dates of the translations.

ARABIC TRANSLATIONS FROM GREEK

Author Title Translator Date/Comments


Euclid The Elements Al-Ḥajjāj b. Time of Harūn al-Rashīd and
Maṭar al-Ma’mūn.
Isḥāq b. Ḥunayn Late ninth century
Thābit b. Qurra Died in 901.
The Data Isḥāq b. Ḥunayn
The Optics
Archimedes Sphere and Cylinder Isḥāq b. Ḥunayn Revised a poor early ninth
Thābit b. Qurra century translation.
Measurement of the Thābit b. Qurra Used commentary of Eutocios
Circle
Heptagon in the Thābit b. Qurra Unknown in Greek
Circle
The Lemmas Thābit b. Qurra
Apollonios The Conies Hilāl al-Ḥimṣī,
Aḥmad b. Mūsā,
Thābit b. Qurra
Diophantos Arithmetic Qusṭā b. Lūqā Died 912
Menelaos Spherica Ḥunayn b. Isḥāq Born 809

The fate of Apollonius’s Conics is instructive. The Arabic bibliographer al-Nadīm tells
us that “... after the book was studied it was lost track of until Eutocius of Ascalon made
a thorough study of geometry.… After he had collected as much of this volume (Conics)
as he could he corrected four of its sections. The Banū Mūsā, however, said that the
volume had eight sections, the part of it now extant being seven, with a part of the eighth.
Hilāl b. Abī al- im ī translated the first four sections with the guidance of Aḥmad b.
Mūsā, and Thābit b. Qurra al-Ḥarrānī [translated] the last three.”

Chart 1
3 Four Muslim Scientists 7

3 Four Muslim Scientists

3.1 Introduction

Like any other civilization that of Islam was not unwavering in its support of
scientists. For example, in the tenth century the scholar al-Sijzī, writing from an
unnamed locality, complained that where he lived people considered it lawful to kill
mathematicians. (Perhaps this was because most mathematicians were also
astronomers, and hence astrologers.) However, whatever hardships the vagaries of a
particular ruler might cause in one area were generally compensated for by a
generous and enthusiastic patron elsewhere, so that, on the whole, mathematicians
and astronomers in Islam could expect both honor and support. For example,
although the son of ‘Abd al-Raḥmān, III in al-Andalūs strongly supported scientific
investigation, his son, Hishām, was deposed by a coup, led by his chamberlain, who
did his best to stamp out interest in the sciences. But, as Ṣā‘id, tells us, some
half-century later, “The present state, thanks to Allah, the Highest, is better than
what al-Andalūs has experienced in the past; there is freedom for acquiring and
cultivating the ancient sciences and all past restrictions have been removed.”1
And, during the time of persecution of scientists in al-Andalūs, the Egyptian
ruler al-Hākim, of whom we shall say more in Chap. 5, founded a library in 1005
called Dār al-Ḥikma. In addition to providing a reading room and halls for courses
of studies, al-Ḥākim paid librarians and ensured that scholars were given pensions
to allow them to follow their studies.
Islamic civilization thus produced, from roughly 750 to 1450, a series of
mathematicians who have to their credit the completion of the arithmetic of a
decimal system that includes decimal fractions, the creation of algebra as one would
learn it in a good high school course today, important discoveries in plane and
spherical trigonometry as well as the systematization of these sciences, and the
creation of elegant procedures for finding numerical solutions of equations. This list
is by no means exhaustive, and we shall detail not only the above contributions but
others as well.
Since the men who made these contributions are probably not well known to the
reader, we begin with some biographical material on four whose names will appear
repeatedly in the following pages. One is Muḥammad b. Mūsā al-Khwārizmī, who
was active in al-Ma’mūn’s Baghdad. The second is Abū a1-Rayḥān al-Bīrūnī,
whose long life bridged the tenth and eleventh centuries, and whose learning and
creative intellect are still impressive. The third, born shortly before al-Bīrūnī died, is
the celebrated ‛Umar al-Khayyāmī, and the fourth, whom a contemporary described
as “the pearl of the glory of his age,” is Jamshīd al-Kāshī, whose work in
Samarqand raised computational mathematics to new heights. Taken together, these

1
From other remarks of Ṣā‘id it appears that the study of astronomy suffered more than that of
mathematics during the earlier period, for he tells us that books on mathematics and medicine were
spared from destruction.
8 1 Introduction

men represent the breadth of interest, the depth of investigation, and the height of
achievement of the best of Islamic scholars.

3.2 Al-Khwārizmī

The springs that fed Islamic civilization came from many lands. Symptomatic of
this is the fact that the family of its greatest early scientist, the Central Asian
scholar, Muḥammad ibn Mūsā al-Khwārizmī, came from the old and high civi-
lization that had grown up in the region of Khwārizm. This is the ancient name for
the region of Uzbekistan around Urgench, a city near the delta of the Āmū Daryā
(Oxus) River on the Aral Sea.
Al-Khwārizmī served the Caliph al-Ma’mūn and is connected to a later caliph,
al-Wāthiq (842–847), by the following story told by the historian al-Ṭabarī. It
seems that when al-Wāthiq was stricken by a serious illness he asked al-Khwārizmī
to tell from his horoscope whether or not he would survive. Al-Khwārizmī assured
him he would live another 50 years, but al-Wāthiq died in 10 days. Perhaps,
al-Ṭabarī tells this story to show that even great scientists can make errors, but
perhaps he told it as an example of al-Khwārizmī’s political astuteness. The hazards
of bearing bad news to a king, who might mistake the bearer for the cause, are well
known. We shall see in the case of another Khwārizmian, al-Bīrūnī, that he too was
very astute politically.
Al-Khwārizmī’s principal contributions to the sciences lay in the four areas of
arithmetic, algebra, geography and astronomy. In arithmetic and astronomy, he
introduced Hindu methods to the Islamic world, and his exposition of algebra was
of prime importance in the development of that science in Islam. Finally, his
achievements in geography earn him a place among the ancient masters of that
discipline.
His arithmetical work The Book of Addition and Subtraction According to the
Hindu Calculation introduced the very useful decimal positional system that the
Hindus had developed by the sixth century A.D., along with the ten ciphers that
make our system so convenient. His book was the first Arabic arithmetic to be
translated into Latin, and its influence on Western mathematics is illustrated by the
derivation of the word algorithm. This word is in constant use today in computing
science and mathematics to denote any definite procedure for calculating some-
thing, and it originated in the corruption of the name al-Khwārizmī to the Latin
version algorismi.
Al-Khwārizmī’s book provided Islamic mathematicians with a tool that was in
constant—though not universal—use from the early ninth century onward. From
the oldest surviving Arabic arithmetic,2 Aḥmad al-Uqlīdisī’s Book of Chapters,
written ca. A.D. 950, to the encyclopedic treatise of 1427 by Jamshīd al-Kāshī, The

2
The earliest versions of al-Khwārizmī’s Hindu Arithmetic are in Latin.
3 Four Muslim Scientists 9

Calculators’ Key, decimal arithmetic was an important system of calculation in


Islam. By the mid-tenth century Aḥmad b. Ibrāhīm al-Uqlīdisī solved some prob-
lems by the use of decimal fractions in his book on Hindu arithmetic, so that, in a
little over a century, al-Khwārizmī’s treatise had led to the invention of decimal
fractions. These too were used by such Islamic mathematicians as al-Samaw’al ben
Yaḥyā al-Maghribī in the twelfth century to find roots of numbers and by al-Kāshī
in the fifteenth century to express the ratio of the circumference of a circle to its
radius as 6.2831853071795865, a result correct to 16 decimal places.
Arithmetic was only one area in which al-Khwārizmī made important contri-
butions to Islamic mathematics. His other famous work, written before his
Arithmetic, is his Kitāb al-jabr wa l-muqābala (The Book of Restoring and
Balancing), which is dedicated to al-Ma’mūn. This book became the starting point
for the subject of algebra for Islamic mathematicians, and it also gave its title to
serve as the Western name for the subject, for algebra comes from the Arabic
al-jabr. In this book many influences are evident, including Babylonian and Hindu
methods for solving what we would call quadratic equations and Greek concerns
with classification of problems into different types and geometrical proofs of the
validity of the methods involved.
The synthesis of Oriental and Greek elements is typical of Islam, as is the
application of a science to religious law, in this case the thorny problems posed by
Islamic inheritance law. A large part of the book is devoted to such problems, and
here again al-Khwārizmī’s example became the model for later Islamic writers.
Thus, after the time of al-Khwārizmī, Abū Kāmil, known as “The Egyptian
Reckoner,” also wrote on the application of algebra to inheritance problems.
And, in the Maghrib, the very earliest mathematical compositions (written in the
late eighth century in the region around Qairawān) appear to have been devoted to
solving arithmetic problems posed by commerce and inheritance law. It was also
around this time that Hindu arithmetic made its appearance in that region, with a
work, On Hindu Reckoning, by Abū Sahl of Qairawān.
Finally, we must comment on al-Khwārizmī’s contribution to the science of
cartography. He was part of the team of astronomers employed by al-Ma’mūn to
measure the length of one degree along a meridian. Since, as we said earlier, the
educated had known that the earth was spherical it followed that multiplication of
an accurate value for the length of one degree by 360 would lead to a good estimate
for the size of the Earth. In the third century B.C.E, the scientist Eratosthenes of
Alexandria, who was the first scientist to be appointed Librarian of the famous
library in that city, used this idea with his knowledge of mathematical astronomy to
obtain an estimate of 250,000 stades for the circumference of the Earth. This was
later shortened by an unknown author to 180,000 stades, a figure far too small but
adopted by the Greek astronomer, Claudius Ptolemy, in his Geography.
We know that the Hellenistic stade is approximately 600 feet but this was not
known to the caliph al-Ma’mūn. As al-Bīrūnī says in his Coordinates of Cities,
10 1 Introduction

al-Ma’mūn “read in some Greek books that one degree of the meridian is equivalent
to 500 stadia…. However, he found that its actual length [i.e., the stade’s] was not
sufficiently known to the translators to enable them to identify it with local stan-
dards of length.” Thus al-Ma’mūn ordered a new survey to be made on the large,
level plain of Sinjār some 70 miles west of Mosul, and two surveying parties
participated. Starting from a common location one party traveled due north and the
other due south. In the words of al-Bīrūnī:
Each party observed the meridian altitude of the sun until they found that the change in its
meridian altitude had amounted to one degree, apart from the change due to variation in the
declination. While proceeding on their paths, they measured the distances they had tra-
versed, and planted arrows at different stages of their paths (to mark their courses). While
on their way back, they verified, by a second survey, their former estimates of the lengths of
the courses they had followed, until both parties met at the place whence they had departed.
They found that one degree of a terrestrial meridian is equivalent to fifty-six miles. He
(Ḥabash) claimed that he had heard Khālid dictating that number to Judge Yaḥyā b.
Aktham. So he heard of that achievement from Khālid himself.

Again one sees an Islamic side to this project in the involvement of a jurist, for
the law was the Islamic religious law and in this case the jurist (qāḍi in Arabic) was
the chief justice of Baṣra, Yaḥyā b. Aktham. Al-Bīrūnī goes on to say that a second
result was also obtained by the survey, namely 56 23 miles/degree, and in fact
al-Bīrūnī uses this value in his own computations later on.
Al-Khwārizmī’s contribution went beyond this to assist in the construction of a
map of the known world, a project that would require solving three problems that
combined theory and practice. The first problem was mainly theoretical and
required mastery of the methods, such as those explained by Ptolemy in the
mid-second century A.D., for mapping a portion of the surface of a sphere (the
earth) onto a plane in a way that would include some visual features of a map on a
sphere. The second was to use astronomical observations and computations to find
the latitude and longitude of important places on the earth’s surface. The difficulties
involved here are both theoretical and practical. The third problem was to sup-
plement these observations by reports of travelers (always more numerous and
usually less reliable than astronomers) on journey-times from one place to another.
Among al-Khwārizmī’s achievements in his geographical work The Image of the
Earth were his correction of Ptolemy’s exaggerated length of the Mediterranean Sea
and his much better description of the geography of Asia and Africa. With such a
map the caliph could survey at a glance the extent and shape of the empire he
controlled and, perhaps more importantly, advertise to all who saw it the extent of
his power.
Thus, it was that al-Khwarizmi’s legacy to Islamic society included a way of
representing numbers that led to easy methods of computing, even with fractions, a
science of algebra that could help settle problems of inheritance, and a map that
showed the distribution of cities, seas and islands on the earth’s surface.
3 Four Muslim Scientists 11

3.3 Al-Bīrūnī

The Central Asian scholar Abū al-Rāyḥan (Fig. 2) al-Bīrūnī was born in Khwārizm
on 4 September, 973. During his youth at least four powers were contending with
each other in and around Khwārizm, so that in his early twenties al-Bīrūnī spent
much of his time either in hiding or fleeing one king to seek hospitality from
another. Despite these setbacks, however, he completed eight works before the age
of 30, including his Chronology of Ancient Nations, the sort of work necessary to
any astronomer who wanted to use (say) ancient eclipse records and needed to
convert the dates given in terms of some exotic calendar into dates in the Muslim
calendar. He had also engaged in a famous controversy on the nature of light with a
precocious teenager from Bukhara named Abū ‛Alī b. Sinā, known to the West as
Avicenna, and he had somehow found the time and the means to construct, and to
use, large graduated rings to determine latitudes. Also, in cooperation with Abū
al-Wafā’ in Baghdad, he used an eclipse of the moon as a time signal to determine
the longitude difference between Kāth (on the River Oxus) and Baghdad. All these
observations and calculations (and those for longitude are indeed difficult) he used
in a book called The Determination of the Coordinates of Localities, where he
continued the tradition of geographical research in Islam that goes back at least to
al-Khwārizmī. He records in this book that he wanted to do measurements to settle
the discrepancy between the two results he had heard of for the number of miles in a
meridian degree, and he writes:
That difference is a puzzle; it is an incentive for a fresh examination and observations. Who
is prepared to help me in this (project)? It requires a strong command over a vast tract of
land and extreme caution is needed from the dangerous treacheries of those spread over it.
I once chose for this project the localities between Dahistān, in the vicinity of Jurjān and the
land of the Ghuzz (Turks), but the findings were not encouraging, and the patrons who
financed the project lost interest in it.

Fig. 2 This stamp, from Guinea-Bissau shows, in the background, part of al-Bīrūnī’s explanation
of the phases of the moon. (The Portuguese account of al-Bīrūnī’s achievements is incorrect.)
12 1 Introduction

He also discovered two new map projections, one known today as the azimuthal
equidistant projection and the other as the globular projection. (See Berggren
(1982) for details.)
Sometime during his thirtieth year, al-Bīrunī was able to return to his homeland,
where he was patronized by the reigning Shāh Abū al-‛Abbās Ma’mūn. The shah
was pressed on the one side by the local desire for an autonomous kingdom and, on
the other, by the clear fact that his kingdom existed at the sufferance of Sultan
Maḥmūd of Ghazna (in present-day Afghanistan), and he was glad to use the
skillful tongue of al-Bīrūnī to mediate the disputes that arose. Of this al-Bīrūnī
wrote perceptively “I was compelled to participate in worldly affairs, which excited
the envy of fools, but made the wise pity me.”
In 1019, even al-Bīrūnī’s “tongue of silver and gold” could no longer control the
local situation, and the army killed Shāh Ma’mūn. Immediately, Sultan Maḥmūd
invaded and among the spoils of his conquest he took al-Bīrūnī back to Ghazna as a
virtual prisoner. Later on al-Bīrūnī’s situation improved and he was able to get
astronomical instruments and return to his observations.
Maḥmūd’s conquests had already made him master of large parts of India, and
al-Bīrūnī went to India where he studied Sanskrit. By asking questions, observing,
and reading Sanskrit texts he compiled information on all aspects of Indian society
and culture. His work India, which resulted from this observation and study, is a
masterpiece, and is an important source for modern Indologists. His comparisons of
Islam with Hinduism are fine examples of comparative religion, and show an
honesty that is not often found in treatises on other people’s religions. Al-Bīrūnī’s
treatment of Indian religion contrasts starkly with that of his patron Maḥmūd who
carried off valuable booty from Indian temples as well as pieces of a phallic idol,
one of which he had installed as a foot-scraper at the entrance to a mosque in
Ghazna. He finished India in 1030, after the death of Sultan Maḥmūd. When the
succession was finally settled in favor of Mas‛ūd, one of Maḥmūd’s two sons,
al-Bīrūnī lost no time in dedicating a new astronomical work to him, the Mas‛ūdic
Canon. This may have won certain privileges for him, since he was subsequently
allowed to visit his native land again.
Sometime after 1040, he wrote his famous work Gems, in which he included
results of his experiments on the specific gravity of many valuable stones. Much of
this material was used by al-Khāzinī who, in the following century, described the
construction and operation of a very accurate hydrostatic balance.
The breadth of al-Bīrūnī’s studies, if not already sufficiently demonstrated, is
plain from his work Pharmacology, which he wrote in his eightieth year while his
eyesight and hearing were failing. The bulk of the work is an alphabetical listing of
about 720 drugs telling, in addition to the source and therapeutic value of each,
common names for the drug in Arabic, Greek, Syriac, Persian, one of the Indian
languages, and sometimes in other languages as well.
We close this brief biography of al-Bīrūnī with the words of E.S. Kennedy,
whose account in the Dictionary of Scientific Biography we have relied on for most
of the details above: (Plate 2).
3 Four Muslim Scientists 13

Plate 2 A gunbad from Ghazna in Afghanistan, built at the time of Sultan Mas‛ūd, the eleventh
century patron of the polymath al-Bīrūnī

Bīrūnī’s interests were very wide and deep, and he labored in almost all the branches of
science known in his time. He was not ignorant of philosophy and the speculative disci-
plines, but his bent was strongly toward the study of observable phenomena, in nature and
in man … about half his total output is in astronomy, astrology, and related subjects, the
exact sciences par excellence of those days. Mathematics in its own right came next, but it
was invariably applied mathematics.

3.4 ‛Umar al-Khayyāmī

‛Umar al-Khayyāmī must be the only famous mathematician to have had clubs
formed in his name. They were not, however, clubs to study his many contributions
to science, but to read and discuss the famous verses ascribed to him under the title
of The Rub‛āyāt (Quatrains) which have been translated into so many of the
14 1 Introduction

world’s languages. Indeed, outside of the Islamic world ‛Umar is admired more as a
poet than as a mathematician, and yet his contributions to the sciences of mathe-
matics and astronomy were of the first order.
He was born in Nishāpūr, in a region now part of Iran, but then known as
Khurasān, around the year 1048. This is only shortly before the death of al-Bīrūnī,
at a time when the Seljuk Turks were masters of Khurasān, a vast region east of
what was then Iran, whose principal cities were Nishapur, Balkh, Marw and Tūs.
His name, “al-Khayyāmī,” suggests that either he or his father at one time practiced
the trade of tentmaking (al-khayyām = tentmaker). In addition, he showed an early
interest in the mathematical sciences by writing treatises on arithmetic, algebra and
music theory, but, beyond these facts, nothing is known of his youth. The nice story
of a boyhood pact with a schoolmate, who later was known as Niẓām al-Mulk and
became a minister in the government of the ruler Malikshah, to the effect that
whichever of them first obtained high rank would help the other is not supported by
the dates at which these men lived. In fact, most scholars believe that ‛Umar died
around 1131, so if he had been Niẓām’s schoolmate, he would have had to be
around 120 years old when he died in order for the story to fit the known dates for
Niẓām al-Mulk. Better founded is the report of the biographer Ẓāhir al-Dīn
al-Bayhaqī. He knew al-Khayyāmī personally and describes him as being both
ill-tempered and narrow-minded. Of course, al-Khayyāmī had examined al-Bayhaqī
as a schoolboy in literature and mathematics, so it may be that he did not get to
know him under the best of circumstances.
We also know that in 1070, when ‛Umar wrote his great work on algebra, he was
supported by the chief judge of Samarqand, Abū Ṭāhir. In this work, ‛Umar sys-
tematically studied all the kinds of cubic equations and used conic sections to
construct the roots of these equations as line segments obtained from the inter-
sections of these curves. There is evidence that ‛Umar also tried to find an algebraic
formula for these roots, for he wrote that “We have tried to express these roots by
algebra but have failed. It may be, however, that men who come after us will
succeed.” The candor of this passage, and its recognition of being part of a tradition
of inquiry that will continue after one’s own death, bespeaks, al-Bayhaqī aside, a
modest and civilized man.
During the 1070s, ‛Umar went to Isfahan (see Plate 3), where he stayed for
18 years and, with the support of the ruler Malikshah and his minister Niẓām
al-Mulk, conducted a program of astronomical investigations at an observatory. As
a result of these researches, he was able, in 1079, to present a plan to reform the
calendar then in use. (One wonders if there is not an echo of this achievement,
together with a nice reference to squaring the circle, in the quatrain “Ah, but my
calculations, people say,/Have squared the year to human compass, eh?/If so by
striking out/Unborn tomorrow and dead yesterday.”) ‛Umar’s scheme made eight of
every 33 years leap years, with 366 days each, and produced a length for the year
closer to the true value than does the present-day Gregorian calendar.
Another important work of ‛Umar’s was his Explanation of the Difficulties in the
Postulates of Euclid, a work composed in 1077, two years before he presented his
calendar reform. In this treatise, ‛Umar treats two extremely important questions in
3 Four Muslim Scientists 15

Plate 3 The entrance to the Friday Mosque in Isfahan, Friday being the day when Muslims gather
together in the mosque to pray and, perhaps, to hear a khutba (homily). Parts of the mosque date
back to the time of ‛Umar al-Khayyāmī. The tile designs on the twin minarets are calligraphic
renderings of the word Allāh (The One God), and calligraphy borders the geometric patterns of
arabesques on the facade

the foundations of geometry. One of these, already treated by Thābit ibn Qurra and
Ibn al-Haytham (known to the West as Alhazen), is the fifth postulate of Book I of
Euclid’s Elements on parallel lines. (In fact, Toth provides evidence that remarks in
16 1 Introduction

Fig. 3 CA and DB are equal straight lines perpendicular to a given straight line, AB, at points A
and B respectively. CD is a straight line joining C and D. It is required to show,without the use of
the parallel postulate, that the angles at C and D are right

various writings of Aristotle imply that mathematicians before Euclid investigated


this question.) ‛Umar bases his analysis on the quadrilateral ABCD of Fig. 3, where
CA and DB are two equal line segments, both perpendicular to AB, and he rec-
ognizes that in order to show that the parallel postulate follows from the other
Euclidean postulates it suffices to show that the interior angles at C and D are both
right angles, which implies the existence of a rectangle. (In fact, the two may be
shown to be equivalent.) Although Ibn al-Haytham, who flourished around 1010,
preceded ‛Umar in using this method of attack on the problem, ‛Umar took issue
with Ibn al-Haytham’s use of motion in geometry. A century and a half later Naṣīr
al-Dīn al-Ṭūsī adopted ‛Umar’s quadrilateral when he wrote his treatment of
Euclid’s parallel postulate. Then, one of his followers, usually referred to as
“Pseudo-Ṭūsī,” wrote an exposition of Euclid’s elements. This work, published in
Arabic in Rome in 1594 under al-Ṭūsī’s name, was translated into Latin and
influenced both J. Wallis and G. Saccheri’s work on the postulate in the seventeenth
and eighteenth centuries.
The other topic that ‛Umar treated in his discussion of the difficulties in Euclid is
that of ratios. Here, al-Khayyāmī’s achievements are twofold. The one is his
demonstration that a definition of proportion that was elaborated in Islamic math-
ematics, a definition that he felt was more true to the intuitive idea of “ratio,” was
equivalent to the definition Euclid used. The other is his suggestion that the idea of
number needed to be enlarged to include a new kind of number, namely ratios of
magnitudes. For example, in ‛Umar’s view, the ratio of the diagonal of a square to
pffiffiffi
the side ð 2Þ, or the ratio of the circumference of a circle to its diameter (π), should
be considered as new kinds of numbers. This important idea in mathematics
amounted to the introduction of positive real numbers and, as was the case with the
parallel postulate, this was communicated to European mathematicians through the
writings of the pseudo-Naṣīr al-Dīn al-Ṭūsī.
‛Umar once said to a friend that when he died he wanted to be buried in Isfahan,
where “the wind will blow the scent of the roses over my grave.” His wish was
granted and the tomb of Islam’s poet–mathematician has remained there to this day.
3 Four Muslim Scientists 17

3.5 Al-Kāshī

Among the nicknames that were sometimes bestowed on mathematicians or astron-


omers in the Islamic world was that of al-ḥāsib (= the reckoner). Strangely, however,
the man who may be the most deserving of that title seems never to have gotten it, but
rather bore the name Ghiyāth al-Dīn Jamshīd al-Kāshī (Fig. 4); but, before we speak of
his remarkable calculations, we must tell what is known of his life.
He was born in the latter half of the fourteenth century in the Persian town of
Kāshān, some 90 miles north of ‛Umar’s tomb in Isfahan, but we know nothing
of his life until the year 1406 when, his own writings say, he began a series of
observations of lunar eclipses in Kāshān. In the following year, also in Kāshān, he
wrote a work on the dimensions of the cosmos, which he dedicated to a minor
prince. Seven years later, in 1414, he finished his revision of the great astronomical
tables written 150 years earlier by Naṣīr al-Dīn al-Ṭūsī, and he dedicated this
revision to the Great Khan (Khāqān) Ulūgh Beg, a grandson of Tamurlane, whose
capital was situated at Samarqand (see Plates 4 and 5). In the introduction to these

Fig. 4 Iranian stamp featuring an artist’s image of al- Kāshī, with an astrolabe as a background
‘halo.’
18 1 Introduction

Plate 4 The tomb of Tamurlane in Samarqand. Tamurlane was a man of considerable intellectual
attainments, as well as being a great military strategist, and his grandson, Ulūgh Beg, was a
generous patron of learning and the arts in the first half of the fifteenth century in Samarqand

tables he speaks of the poverty he endured and how only the generosity of Ulūgh
Beg allowed him to complete the work. Then in 1416, two years later, he finished a
short work on astronomical instruments in general, dedicated to Sulṭān Iskandar (a
Black Sheep Turk and member of a dynasty rival to that of the offspring of
Tamurlane), and a longer treatise on an instrument known as an equatorium. This
instrument is, in essence, an analog computer for finding the position of the planets
according to the geometrical models in Ptolemy’s Almagest, and its utility is that it
allows one to avoid elaborate computations by manipulating a physical model of
Ptolemy’s theories to find the positions of the planets. See Fig. 5 for a picture of
al-Kāshī’s equatorium.
The writing on the equatorium marked the end of al-Kāshī’s career as a wan-
dering scholar, and the next we hear of him is as a member of the entourage of
Ulūgh Beg, to whom he had dedicated his Khāqānī tables. Exactly when al-Kāshī
arrived in Samarqand we do not know, but during the year 1417 Ulūgh Beg began
building a madrasa (school) there, whose remains still impress visitors to the site,
and, on its completion, began construction on an observatory (see Plates 6 and 7).
Two letters from al-Kāshī to his father have survived and give us a rare glimpse
into details of the intellectual life at Ulūgh Beg’s court. In one of these letters
al-Kāshī describes at length the accomplishments of Ulūgh Beg who, he says
(translated in Kennedy et al., p. 724)
3 Four Muslim Scientists 19

Plate 5 A modern bust of al-Kāshī’s patron Ulūgh Beg at the museum in Samarqand. Ulūgh Beg
was himself an accomplished astronomer whose astronomical tables were used in Europe into the
seventeenth century

“has by heart most of the glorious Qur’ān … and every day he recites two chapters in the
presence of (Qur’ān) memorizers and no mistake is made. He knows (Arabic) grammar
well and he writes Arabic composition extremely well, and likewise he is well posted in
canon law; he has knowledge of logic, rhetoric, and elocution, and likewise of the Elements
(of Euclid?) , and he himself cultivates the branches of mathematics, and this has reached
the extent that one day while riding he wanted to determine the date, which was a Monday
of [the month of] Rajab, between the fifth and the tenth in the year eight hundred and
eighteen (A.H.), as to what day it was of the (astronomical) season of the year. From these
very given data, by mental computation, and from horseback, he determined the true
longitude of the sun (correct) to degrees and minutes. When he came back he asked this
humble servant about it. Truly, since in mental computation the quantities must be retained
by memory and others determined, and there is a limit to one’s strength of retention, he (i.e.
I) was not able to extract it to degrees and minutes, but contented myself with degrees.”

It is perhaps because of Ulūgh Beg’s enlightened patronage of learning that


al-Kāshī goes on to call Samarqand a place where “the rams of the learned are
20 1 Introduction

Fig. 5 Al-Kāshī’s equatorium set for finding the longitude of a planet. The outer rim shows the
Arabic names for the zodiacal signs, counterclockwise from Aries, the word just above “epicycle.”
Figure adapted from Kennedy, The Planetary Equatorium. Reprinted by permission of Princeton
University Press

gathered together, and teachers who hold classes in all the sciences are at hand, and
the students are all at work on the art of mathematics.”
However, he leaves no doubt in his father’s mind that he is the most able of all
gathered there. He first tells how he found for the scholars at Samarqand the
solution of difficulties involved in laying out the star-map of an astrolabe. He then
lists another triumph (Kennedy et al., p. 726):
Furthermore, it was desired to set up a gnomon on the wall of the royal palace and to draw
the lines of the equal hours on it (the wall). Since the wall was neither in the line of the
meridian nor in the east–west line, no one had done the like of it (before) and (they) could
not do it at all. Some said that it can be done in one year, i.e., as soon as the sun reaches the
beginning of a zodiacal sign, on that day let an observation be made for each hour, and a
mark made until it is finished. When this humble servant arrived it was commanded that
this humble servant (lay out the lines), which was completed in one day. When it was
examined by use of a big astrolabe it was found to be in agreement and proper arrangement.
3 Four Muslim Scientists 21

Plate 6 The observatory, as the scientific institution we know today, was born and developed in
the Islamic world. Here is a part of the sextant (or perhaps quadrant) at the observatory in
Samarqand where al-Kāshī worked. It was aligned in the north–south direction and was 11 meters
deep at the south end. Thus an astronomer sitting between the guide rails could have seen the stars
crossing the meridian even in the daytime while assistants sitting on either side held a sighting
plate through which he could observe the transits of heavenly bodies. It was at this observatory that
the greatest star catalog since the time of Ptolemy was compiled

One can imagine what al-Kāshī thought (and probably said) about the persons he
mentions who could only draw the lines for a sundial not facing one of the cardinal
directions by waiting for a year and marking off the shadows month-by-month.
It is from this period in Samarqand, which is from about 1418, that al-Kāshī’s
greatest mathematical achievements come. One of these is his spectacular calcu-
lation, in 1424, of a value for 2π which, when he expresses it in decimal fractions, is
correct to 16 decimal places. In order to achieve this accuracy, he calculated the
perimeters of inscribed and circumscribed polygons, in a given circle, having
805,306,368 sides. What makes the achievement especially impressive is that
al-Kāshī states in advance how close he wants his approximation to be and then
carefully plans how accurate each stage must be so that what we would call
round-off errors do not accumulate as he goes through the series of root extractions
22 1 Introduction

Plate 7 This mausoleum houses the tomb of al-Kāshī’s colleague at Samarqand, Qāḍīzādeh
Rūmī, one of the few colleagues he seems to have had some respect for. The building derives its
basic structure from a favorite pattern in Muslim architecture—a dome surmounting an octagonal
base which is itself resting on a square

necessary to arrive at the final result. Al-Kāshī phrases his requirement for accuracy
by stating that he wants the value to be so accurate that, when it is used to calculate
the circumference of the universe according to the ancient dimensions, the result
would not differ from the true value by more than the width of a horse’s hair.
Although this treatise on π bears no dedication, the work he completed two years
later, a compendium of arithmetic, algebra and measurement called The
Calculators’ Key, is dedicated to Ulūgh Beg, and as, the crowning achievement of
Islamic arithmetic, is truly a gift fit for a king. Among its many jewels is a sys-
tematic exposition of the arithmetic of decimal fractions, an invention al-Kāshī
claims as his own, and a beautiful algorithm for finding the fifth root of a number
3 Four Muslim Scientists 23

which he demonstrated on a number on the order of trillions. Such was the


excellence of the work that, according to the Persian scholar Muḥammad Ṭāhir
Ṭabarsī, al-Kāshī’s book was the standard arithmetic and algebra text in the Persian
madrasas until the seventeenth century. It is also of interest that a copy of The
Calculators’ Key found in the British Museum was made by the great-great-great
grandson of al-Kāshī.
Finally, among the works al-Kāshī mentions in the preface to The Calculators’
Key is one on solving a cubic equation to obtain a value for Sin(l°). A remarkable
feature of al-Kāshī’s method, whose details we explain in Chap. 5, is that the
calculator can repeat the procedure; each time using the last result obtained, and so
obtain numbers as near to the true value of the root of the equation as he pleases.
Al-Kāshī’s remarkable career ended on the morning of 22 June 1429 when he
died at the observatory he had helped to build. In the preface to his own astro-
nomical tables, written some eight years after al-Kāshī’s death, Ulūgh Beg refers to
al-Kāshī as “the admirable mullah, known among the famous of the world, who had
mastered and completed the sciences of the ancients, and who could solve the most
difficult questions.”
These words are a fitting epitaph, not only for al-Kāshī but for any of the great
mathematicians of Islam, and they may end this brief biographical section. (We
shall, of course, give biographical details about some of the figures mentioned later
in the book as they appear.)

4 The Sources

Most sources for the history of Islamic mathematics are treatises running from a few
pages to several hundred pages, written with ink and usually on paper (see Plate 8).
Typically, several treatises are bound together to form a volume, called a codex,
many of which have interesting histories. For example, the owners’ marks on the
title page of a manuscript now at the library of the great shrine at Meshhed in Iran,
which was copied in 1462–1463, show that in the seventeenth century it was in the
library of Shāh Jahān in India, the Mughal emperor and patron of the sciences.
There are a variety of notations from the late seventeenth century, still at the court
of the Mughals. Then, by the nineteenth century the manuscript had reached
Meshhed, at a library called the Fāḍilīya, whose contents were recently incorporated
into those of the shrine. Some scholars have conjectured that the manuscript came
to Meshhed as part of the booty brought back to Iran by Nādir Shāh when he
defeated the Mughals in 1739. Certainly it would not be the unique instance of
scientific manuscripts being among the spoils of war, for al-Ma’mūn obtained
Greek manuscripts from the Byzantines by the terms of peace treaties.
In former times Arabic scientific manuscripts could be bought very cheaply, on
the open market or otherwise. In addition, European collectors could hire an
Arabic-writing scribe to copy old manuscripts for them. In this way, through
purchase, theft, gift and copying, many large collections were built up by
24 1 Introduction

Plate 8 This page of a manuscript housed in the Dār al-Kutub (“abode of books”) in Cairo is from
a treatise by Ibrāhīm b. Sinān on the area of a parabolic section. This copy was made by the scholar
Muṣṭafa Sidqī in the early eighteenth century. Perhaps Sidqī’s competence in mathematics was one
reason the diagrams are so carefully done. (One often finds very nicely written manuscripts with
spaces carefully left for diagrams which were never drawn.) Photo courtesy of Egyptian National
Library
4 The Sources 25

Europeans and these collections were in turn donated to or purchased by European


libraries. Thus at Berlin, Dublin, the Escorial in Spain, Leiden, London, Oxford,
and Paris (to name only some of the main sites) there are large collections of Arabic
manuscripts accessible to scholars, and there are also several large collections in
both the U.S., Russia and such Central Asian Republics as Uzbekistan and
Tajikistan.
In the Islamic world, there are of course large collections of manuscripts all the
way from Morocco through Afghanistan to India and S.E. Asia. Many of these
libraries are generous in providing access to the collections, but in other cases local
politics and nationalism combine with a lack of adequate catalogs to make access
difficult for scholars.
The study of Arabic manuscripts presents the same problem as does the study of
Greek manuscripts, namely that one is usually dealing with copies, many times
removed, of a vanished original; but, additional difficulties arise because of the fact
that, unlike Indo-European languages, written Arabic indicates only consonants and
long vowels. Thus, the title of al-Kāshī’s book in Arabic is miftāḥ
al-ḥisāb, and the second word, reading in the direction of Arabic writing from right
to left, could be read either as ḥisāb or as husāb, where the first means “arithmetic”
and the second “calculators.” If the context does not provide clues, and if there are
no special marks to indicate which is intended, there is simply no way to tell.
A further problem arises because very different letters are distinguished only by one
or more dots, and often the dots are omitted. For example, the letters are,
respectively, “J,” “H” (approximately), and “Kh,” and so, for the famous mathe-
matician depending on whether the has a dot written underneath or above,
his name will be read “al-Karajī” or “al-Karkhī.” In the first case it implies he came
from Iran and in the second case from Iraq, but, since scholars say about as many
manuscripts support one reading as another, we shall probably never know the
origin of one of the greatest Islamic algebraists.
Despite these difficulties, the last few decades have seen an increased interest in
the study of all aspects of the mathematical sciences in the Islamic world. And, even
though we are far from having the same proportion of carefully edited texts of
major works that students of Greek mathematics have at their disposal, we do
possess at least the main outlines, and a good number of details, of the story that
occupies the following chapters.

5 The Arabic Language and Arabic Names

5.1 The Language

The Arabic language is a member of the group of languages known as the Semitic
group, a group that includes Hebrew, Ethiopian, Babylonian and Phoenician. All
members of this group share the characteristic that most of their words are formed
26 1 Introduction

around roots of three consonants. For example, the consonants “k t b” in that order
carry the meaning of writing. Kataba means “he wrote,” and if the first and second
vowels are changed to “u” and “i” then one has kutiba, “it was written.”
The roots of Arabic words are quite unfamiliar to speakers of Western languages
and one learning Arabic must do an immense amount of memorization to learn an
entire vocabulary with almost no similarities to English. (One of the few areas
where there is any help from English is a few terms in mathematics and astronomy
we have borrowed from Arabic.) However, one feature of the Arabic language that
does ease the acquisition of its vocabulary is that derived forms of the root are
formed in a standard way, independently of the root. To illustrate this consider the
three roots k t b = writing, r ṣ d = observing, ḥ s b = computing. If XYZ is any root
then XāYiZ denotes a person performing the action of the root, so that kātib is a
scribe, rāṣid a watcher and ḥāsib a calculator (or an astronomer, who had to do so
many calculations). The place where the action corresponding to the root is done is
denoted by the form maXYaZ, so that maktab is an office or a desk and marṣad is
an observatory. As a last example miXYaZ is an instrument for doing something so
that mirṣad is a telescope. Arabic has standard forms not only for such concrete
notions as the above but also for more subtle variations of the basic meaning, and
this feature made it possible for the early Arabic translators to find Arabic equiv-
alents for a wide range of concepts in Greek, Persian, and Indian science and
philosophy.

5.2 Transliterating Arabic

The reader perhaps noticed in the examples above that some vowels are written with
a bar, which linguists call a macron, over them. In fact, the Arabic alphabet has 28
letters, all of them consonants. Short vowels are indicated by the marks “–” (for “a”)
and “–” (for “u”) placed above the consonant and “–” for “i” placed under the
consonant. Apart from the Muslim scriptures, the Qur’ān, short vowels are written
only when the possibility of an ambiguity occurs to the writer. However, since long
vowels are used to distinguish one standard derived form from another they must be
indicated. This is done in Arabic by using three letters to indicate the three long
vowels, “aliph” (ā), “waw” (ū) and “yā” (ī). It is done in transcription by the macron.
-
, placed above the letter.
Since English has only 21 consonants to Arabic’s 28, some special devices are
necessary to transliterate Arabic to English. (“Transliteration” means using one
system of representing sounds—e.g., the English alphabet—to represent another—
e.g., the Arabic alphabet.) For the names in this book we are using the system of
transliteration explained in Haywood and Nahmad (except that we do not use any
underlining). This is why the reader will see in transliterated Arabic words not only
the consonants ḥ, ṣ, ḍ, ṭ, ẓ but also h, s, d, t, z. Corresponding letters, h and ḥ for
example, may be pronounced conveniently, if incorrectly, as identical. (Obviously
the sounds must have something in common.) Or the reader may consult Tritton
5 The Arabic Language and Arabic Names 27

where phonetic equivalents of all the Arabic consonants are given. The only other
signs we shall comment on are ’ and ‛. The first of these represents a “hamza” and is
pronounced as a glottal stop, the clicking sound we make when emphasizing the
word “am.”’. The second represents the eighteenth letter, “‛ayn,” and is pronounced
as a sort of short growl, deep in the throat. It has no English equivalent—or even
approximant.

Exercises

Note: These exercises are suggestions for library research and are not intended to be
answered on the basis of the information in this chapter. This is also true of several
of the exercises in the following chapters.
1. Write a brief account of the lives and works of any of the following: (1) The
Banū Mūsā, (2) al-Kindī, (3) Kamāl al-Dīn Fārisī or (4) Naṣīr al-Dīn al-Ṭūsī.
2. Write a short paper giving an account of the lives and works of some of the
translators mentioned in this chapter, or of other important translators you have
come across in your reading.
3. Write a short account of the major astronomical observatories in Islam.
4. Where were the major centers of political and military power in the Islamic
world between the years 700 and 1400?
5. What are the main features of the calendar in use in Baghdad in the tenth
century?
6. How did the man who called out the times for the five daily prayers know when
it was time?

Bibliography

The following texts are basic references, which are useful for
further reading on any of the topics in this book

Gillespie, C. C. et al. (eds.). Dictionary of Scientific Biography (16 vols.). New York: Charles
Scribner’s Sons, 1972–80. Now supplemented by Noretta Koertge (ed.), New Dictionary of
Scientific Biography. 8 vols. Detroit: Charles Scribner’s Sons, Gale/Cengage Learning, 2008.
This latter work updates many of the articles in the Dictionary on scientists of medieval Islam.
Hockey, Thomas; et al. (ed.): Biographical Encyclopedia of Astronomers (2 vol’s). Springer: New
York 2007.
Rozenfel’d, Boris A. and Ekmeleddin İhsanoğlu. Mathematicians, astronomers and other scholars
of Islamic civilisation and their works (7th-19th c.) Istanbul:Research Centre for Islamic
History, Art, and Culture, 2003.
Sezgin, F. Geschichte des arabischen Schrifttums (Vol. 5 (Mathematics) and Vol. 6 (Astronomy)).
Leiden: E. J. Brill, 1974 and 1978, resp.
28 1 Introduction

The preceding two works are standard bio-bibliographical surveys of Arabic literature
Storey, C. A. Persian Literature: A Bio-Bibliographical Survey. Luzac and Co., 1927.
Wensink, H. et al. (eds.). The Encyclopaedia of Islam, 2nd ed. Leiden: E. J. Brill, 1960–2009. The
third edition, EI3 (edited by Marc Gaborieau et al.) of this basic scholarly reference work is
currently being published.

Some important general sources for further study of the history


of mathematics in Islam are the following

Kennedy, E. S. “The Exact Sciences in Iran under the Saljuks and Mongols”. In: Cambridge
History of Islam, Vol. 5. Cambridge, U.K.: Cambridge University Press, 1968, pp. 659–679.
Kennedy, E.S., et al. 1983. Studies in the Islamic Exact Sciences. Beirut: American University of
Beirut Press.
The above-named book contains a collection of papers on topics in the history of mathematics,
astronomy and astrology, and geography in Islam by one of the most distinguished workers in
the field, as well as by his students
Rashed, R. (ed.). 1996. Encyclopedia of the History of Arabic Science, Vols. I–III. London and
New York: Routledge.
Sidoli, N. and G. Van Brummelen (ed's). From Alexandria Through Baghdad: Studies in the
Ancient Greek and Medieval Islamic Mathematics Sciences in Honor of J.L. Berggren. New
York: Springer, 2014.

The following bear more specifically on the material in Chapter 1

Bagheri, M. 1977. “A newly found letter of al-Kāshī on scientific life in Samarkand.” Historia
Mathematica 24: 241–256.
Berggren, J. L. 1996. “Islamic acquisition of the foreign sciences: A cultural perspective,” in
Tradition, transmission, transformation: Proceedings of Two Conferences on Pre-modern
Science Held at the University of Oklahoma, ed. F. Jamil Ragep and Sally P. Ragep (Leiden: E.
J. Brill, 1996), pp. 263–83.
Berggren, J.L. 1985. “Nine Muslim Sages”, Hikmat 1 (No. 9) (1979), and “Mathematics in
Medieval Islam.” Hikmat 2 (12–16): 20–23.
Both articles above contain biographical sketches of important mathematicians of the Islamic
world
Berggren, J.L. 1981. “Al-Bīrūnī on plane maps of the sphere.” Journal for the History of Arabic
Science 5: 47–96.
Haywood, J.A., and H.M. Nahmad. 1990. A New Arabic Grammar, 2nd (revised) ed. London:
Lund Humphries.
Hitti, Phillip. 1968. The Arabs: A short history, 5th ed. New York: St. Martin’s Press.
This is an abridgement of his History of the Arabs
Katz, V. (ed.) 2007. Mathematics of Egypt, Mesopotamia, China, India and Islam: A source book.
(Princeton, N.J.: Princeton University Press. (The section on Islam, by the present author, deals
almost entirely with the eastern part of the Islamic world. For sources from al-Andalūs and the
Maghrib consult the following entry.).
Bibliography 29

Katz, V. (ed.) 2016. Source Book in the Mathematics of Medieval Europe and North Africa.
Princeton, NJ: Princeton U. Press. (This source book contains selections of medieval
mathematics from the Latin West, mathematics written in Europe in Hebrew, and a section by
the present writer on mathematicians of al-Andalūs and the Maghrib.).
Kennedy, E. S. “A Letter of Jamshid al-Kāshī to His Father”, Orientalia 29 (1960), 191–213. This
is reprinted in pp. 722–744 of Kennedy et al. (above).
Kennedy, E, S. 1960. The Planetary Equatorium of Jamshīd Ghiyāth al-Dīn al-Kāshī (d.1429)
(Princeton Oriental Studies 18). Princeton 1960.
Pedersen, J. 1984. The Arabic Book (transl. by G. French). Princeton, NJ: Princeton University
Press.
Rosenthal, F. 1975. The Classical Heritage in Islam. Berkeley and Los Angeles, CA: University of
California Press.
The above collection of texts and accompanying commentary illustrates the extent of medieval
Islam’s acquaintance with the classical world
Ṣā‘id al-Andalusī, Science in the Medieval World: “Book of the Categories of Nations” (trans. and
edited by S. I. Salem and Alok Kumar). Austin, Texas: U. of Texas Press, 1991.
Saliba, G. 2007. Islamic science and the making of the European Renaissance. Cambridge, MA
and London: MIT Press.
Toth, Imre, “Non-Euclidean Geometry Before Euclid”, Scientific American, Nov. (1969), 87–95.
Toomer, G. J. 1984. “Lost greek mathematical works in Arabic Translation”. The Mathematical
Intelligencer 6 (No. 2), 32–38.
Tritton, A.S. 1975. Arabic (Teach Yourself Books). London: Hodder and Stoughton.
Chapter 2
Arithmetic in the Islamic World

1 The Decimal System

Muslim mathematicians were the first people to write numbers the way we do, and,
although we are the heirs of the Greeks in geometry, the part of our legacy from the
Muslim world is our arithmetic. This is true even if it was Hindu mathematicians in
India, probably a few centuries before the rise of Islamic civilization, who began
using a numeration system with these two characteristics:
1. The numbers from one to nine are represented by nine digits, all easily made by
one or two strokes.
2. The rightmost digit of a numeral counts the number of units, and a unit in any
place is ten of that to its right. Thus, the digit in the second place counts the
number of tens, that in the third place the number of hundreds (which is ten
tens), and so on. A special mark, the zero, is used to indicate that a given place is
empty.
These two properties describe our present system of writing whole numbers, and
we may summarize the above by saying the Hindus were the first people to use a
cipherized, decimal, positional system, “Cipherized” means that the first nine
numbers are represented by nine ciphers, or digits, instead of accumulating strokes
as the Egyptians and Babylonians did, and “decimal” means that it is base 10.
However, the Hindus did not extend this system to represent parts of the unit by
decimal fractions, and, since it was the Muslims who first did so, they were the first
people to represent numbers as we do. Quite properly, therefore, we call the system
“Hindu–Arabic”.
As to when the Hindus first began writing whole numbers according to this
system, the available evidence shows that the system was not used by the great
Indian astronomer Āryabhata (born in A.D. 476), but it was in use by the time of his
pupil, Bhaskara I, around the year A.D. 520. (See Van der Waerden and Folkerts
for more details.)

© Springer Science+Business Media New York 2016 31


J.L. Berggren, Episodes in the Mathematics of Medieval Islam,
DOI 10.1007/978-1-4939-3780-6_2
32 2 Arithmetic in the Islamic World

News of the discovery spread, for, about 150 years later, Severus Sebokht, a
bishop of the Nestorian Church (one of the several Christian faiths existing in the
East at the time), wrote from his residence in Keneshra on the upper Euphrates river
as follows:
I will not say anything now of the science of the Hindus, who are not even Syrians, of their
subtle discoveries in this science of astronomy, which are even more ingenious than those
of the Greeks and Babylonians, and of the fluent method of their calculation, which sur-
passes words. I want to say only that it is done with nine signs. If those who believe that
they have arrived at the limit of science because they speak Greek had known these things
they would perhaps be convinced, even if a bit late, that there are others who know
something, not only Greeks but also men of a different language.

It seems, then, that Christian scholars in the Middle East, writing only a few
years after the great series of Arab conquests had begun, knew of Hindu numerals
through their study of Hindu astronomy. The interest of Christian scholars in
astronomy and calculation was, in the main, due to their need to be able to calculate
the date of Easter, a problem that stimulated much of the Christian interest in the
exact sciences during the early Middle Ages. It is not a trivial problem, because it
requires the calculation of the date of the first new moon following the spring
equinox. Even the great nineteenth-century mathematician and astronomer C.F.
Gauss was not able to solve the problem completely, so it is no wonder that Severus
Sebokht was delighted to find in Hindu sources a method of arithmetic that would
make calculations easier.
We can perhaps explain the reference to the “nine signs” rather than the ten as
follows: the zero (represented by a small circle) was not regarded as one of the
digits of the system but simply a mark put in a place when it is empty, i.e., when no
digit goes there. The idea that zero represents a number, just as any other digit does,
is a modern notion, foreign to medieval thought. This is clearly shown in
al-Khalili’s auxiliary tables for certain combinations of trigonometric functions
depending on two arguments, x and y. In the case of values of x and y that would
produce a value outside the domain of the arcos function al-Khālilī writes “0 0,”
which can only mean ‘no value’ not zero degrees zero minutes.
With this evidence that the Hindu system of numeration had spread so far by the
year A.D. 662, it may be surprising to learn that the earliest Arabic work we know
of explaining the Hindu system is one written early in the ninth century whose title
may be translated as The Book of Addition and Subtraction According to the Hindu
Calculation. The author was Muḥammad ibn Mūsā al-Khwārizmī who, since he
was born around the year A.D. 780, probably wrote his book after A.D. 800.
We mentioned in Chapter 1 that al-Khwārizmī, who was one of the earliest
important Islamic scientists, came from Central Asia and was not an Arab. This was
not unusual, for, by and large, in Islamic civilization it was not a man’s place (or
people) of origin, his native language, or (within limits) his religion that mattered,
but his learning and his achievements in his chosen profession.
The question arises, however, where al-Khwārizmī learned of the Hindu arith-
metic, given that his home was in a region far from where Bishop Sebokht learned
1 The Decimal System 33

of Hindu numerals 150 years earlier. In the absence of printed books and modern
methods of communication, the penetration of a discovery into a given region by no
means implied its spread to adjacent regions. Thus al-Khwārizmī may have learned
of Hindu numeration not in his native Khwārizm but in Baghdād, where, around
780, the visit of a delegation of scholars from Sind to the court of the Caliph
al-Manṣūr led to the translation of Sanskrit astronomical works. Extant writings of
al-Khwārizmī on astronomy show he was much influenced by Hindu methods, and
it may be that it was from his study of Hindu astronomy that he learned of Hindu
numerals.
Whatever the line of transmission to al-Khwārizmī was, his work helped spread
Hindu numeration both in the Islamic world and in the Latin West. Although this
work has not survived in the Arabic original (doubtless because it was superseded
by superior treatises later on), we possess a Latin translation, made in the twelfth
century A.D. From the introduction to this we learn that the work treated all the
arithmetic operations and not only addition and subtraction as the title might
suggest. Evidently al-Khwārizmī’s usage is parallel to the somewhat dated English
description of a child who is studying arithmetic as “learning his sums.”

2 Kūshyār’s Arithmetic

2.1 Survey of The Arithmetic

As we have said, al-Khwārizmī’s book on arithmetic is no longer extant in Arabic,


and one of the earliest works on Hindu numeration whose Arabic text does exist
was written by a man named Kūshyār b. Labbān, who was born in the region south
of the Caspian Sea some 150 years after al-Khwārizmī wrote his book on arith-
metic. Although Kūshyār was an accomplished astronomer, we know very little
about his life, but despite this personal obscurity his works exerted some influence,
and his treatise on arithmetic, whose title means Principles of Hindu Reckoning,
became one of the main arithmetic textbooks in the Islamic world.
Kūshyār’s concise treatise is a carefully written introduction to arithmetic,
divided into two main parts. The first contains, after a brief introduction, nine
sections on decimal arithmetic, beginning with “On Understanding the Forms of the
Nine Numerals.” In this the nine numerals are given in a form standard in the east,
namely:

and the place-value system is explained. Zero is introduced as the symbol to be


placed in a position “where there is no number.” The Arabic word for zero, “ṣifr,”
comes from the verb “ṣafira” which means “to be empty or void” and it is the
source, via French and Spanish, of our word “cipher”. It is even the source, via
34 2 Arithmetic in the Islamic World

Italian, of our word “zero.” The sections following the first are “On Addition,” “On
Subtraction”—with halving treated in a special subsection, “On Multiplication,”
“On the Results of Multiplication,” “On Division,” “On the results of Division,”
“On the Square Root,” and “On Checking,” through casting out nines.
The sixteen sections of the second part contain an explanation of the arithmetic
of a base-60 positional system, but the book concludes with a section that tells how
to find the cube root of a number in the decimal system. The base-60 system, which
is now called a sexagesimal system, was important to astronomers because angles
were measured, and trigonometric functions were tabulated, according to this
system, and because its unified treatment of whole numbers and fractions made
calculations so much simpler. We shall say more of this later.
As we follow Kūshyār’s explanation of the decimal system it is well to bear in
mind that he was explaining arithmetic to people who would be computing not with
pen or paper but with a stick (or a finger) on a shallow tray covered with fine sand,
which we shall refer to as a “dust board.” Because small boards are more conve-
nient to carry around than large ones, it is desirable to have arithmetic algorithms
that do not require writing down several rows of numbers. On the other hand, it is
easy to erase on a dust board, so algorithms that require considerable erasing pose
no problem, and we shall see how the algorithms for addition, subtraction, multi-
plication, division, and extracting square roots were designed with this feature of
the dust board in mind.
In the text of his book Kūshyār writes out, in words, all the names of the
numbers, and it is only when he is actually exhibiting what is written down on the
dustboard that he uses the Hindu-Arabic ciphers. A reason for this may be that
explanations were considered as text, and therefore written out in words, like any
other text. The examples of what was written on the dust board, however, may have
been viewed as illustrations, much like a diagram in a geometrical argument, and
they were there to show what the calculator would actually see on the dust board.

2.2 Addition

As Kūshyār explains this, the numbers to be added are written in two rows, one
above the other, so that places of the same value are in the same column. He gives
the example of adding 839 to 5625 and, unlike our method, begins his addition by
adding from the highest place common to both numbers, in this case the hundreds’
place, down. At each stage the answer obtained so far replaces part of the number
on the top. Figure 1 illustrates his steps, beginning with 56 + 8 = 64, and an arrow
(!) shows that the display on the right replaces, on the dust board, that on the left.
Thus, at any time, there are only two numbers on the dust board, arranged in
columns, and, in the end, the answer has replaced the number on the top. Unlike our
method, the method Kūshyār explains obtains the leftmost digit of the answer first.
2 Kūshyār’s Arithmetic 35

Fig. 1

2.3 Subtraction

Again Kūshyār explains the method by the same numbers, subtracting 839 from
5625, and again he works from left to right. He explains that since 8 cannot be
subtracted from 6 it must be subtracted from 56 to produce 48, so the 56 of 5625 is
erased and replaced by 48. Thus, working from place to place, Kūshyār obtains the
answer, 4786 (Fig. 2). There is no “borrowing” in Kūshyār’s procedure. He simply
notices that, for example, in the last step, since we cannot subtract 9 from 5 we must
subtract it from 95. Just as with addition Kūshyār works from the higher places to
the lower, and at each stage the partial answer appears as part of the number on top.
His treatment of halving, which he considers to be a variant of subtraction, sheds
light on his treatment of fractions. He begins with 5625 (as usual), but this time he
starts on the right (Fig. 3). He says to set down 5625 and then take half of five, which
is two and a half. “Put two in the place of the five and put the 12 under it, thirty.”
He is using here, for fractions, the sexagesimal system, which goes back to the
Babylonians and uses the principle of place value to represent fractions in terms of
multiples of the subunits 1/60, 1/602 = 1/3600, etc. He explains the system more
fully in the second part of the treatise, and here he contents himself with using his
readers’ familiarity both with the local monetary system in which a dirham con-
tained 60 fulūs, and with degrees, in which 1 degree contains 60 minutes. Thus he
tells his reader, in effect, “If you wish to think of 5625 as dirhams (degrees), then

Fig. 2

Fig. 3
36 2 Arithmetic in the Islamic World

think of 52 dirhams as 2 dirhams and 30 fulūs (2 degrees and 30 minutes).” The next
two steps of his calculation, as shown in Fig. 3, are to halve the 2 in the 10’s place
and then the 6 in the 100’s place, and now he must take half of the 5 in the 1000’s
place. Kūshyār values the 5 in terms of the preceding place, and so looks on it as 50
hundreds. Its half is thus 25 hundreds, and so in the last step he adds the 2500 to the
half of 625 to obtain the answer shown in Fig. 3.

2.4 Multiplication

The algorithm for multiplication shows a thorough understanding of the rule for
multiplying powers of 10, for to multiply 243 by 325 Kūshyār requires his reader to
arrange the numerals so the 3 of 325 is directly above the 3 of 243 (Fig. 4). The
total array occupies five columns, because hundreds multiplied by hundreds yields
tens of thousands ((N  100)  (M  100) = N  M  10,000). Since 3  2 = 6, he
places the 6 directly above the 2, i.e., in the ten thousands’ place, and he remarks
that had the product produced a two-digit number (e.g., had it been 4  3 = 12), the
tens’ digit of the product would be placed in the column to the left of the 2. This is
illustrated at the next step where, since 3  4 = 12, he places the 2 of the 12 directly
above the 4 and adds the 1 to the 6 to get 72. Finally, the top 3 is replaced by the
9 = 3  3, since he no longer needs to multiply by it.
Now we will be multiplying 243 by the upper 2, and since this counts “tens” and
not “hundreds” we must, if we are to continue adding the answers to the top row in
the columns above the bottom numerals, shift 243 one place to the right, since the

Fig. 4
2 Kūshyār’s Arithmetic 37

powers of 10 represented by the answers will be one less. Thus, we begin in the
second row of Fig. 4, and as before, the last digit of the lower number (3) stands
under the current multiplier (2). Then, since 2  2 = 4, we add the 4 to 72 to get 76,
and the remaining steps of Row 2 will be clear to the reader who has followed those
of Row 1. Again, a shift to the right automatically lines up the figures so that the
answers are put in the correct place. We thus have Row 3 of Fig. 4, and the only
thing to be careful of is that on the last multiplication of a given sequence (for
example the final “5  3 = 15”) one does not add the final digit of the product to the
last multiplier (in this case, 5), but, instead, uses it to replace the multiplier.

2.5 Division

This operation offers Kūshyār no more trouble than multiplication, as the division
of 5625 by 243 shows. In Fig. 5 the divisor, 243, is written at the bottom and the
dividend, 5625, is written above that, its highest place written above the highest
place of the divisor. The first digit of the quotient, 2, is obtained by an estimate and
is written in the column above the last digit of the divisor, 243. In Fig. 5 the first
three steps show the process of subtracting 2  243 = 486 from 562. In this case, the
“2,” since it is written above the tens’ place of the dividend, means 20, and the
positioning automatically puts it in the right column. This process is shown in the
first four boxes of Fig. 5, and the fourth box says that 5625 – 20  243 = 765 Now
Kūshyār moves the divisor one column to the right, so that the next digit of the
quotient will be correctly aligned. The second row calculates 765 – 3  243 = 36,
and so, calculating digit by digit, beginning with the one in the highest place,
Kūshyār obtains the quotient (23) and remainder (36).
This answer, 23 + 36/243, is correct but raises the further question, “How big is
the fraction 36/243?” After all, an astronomer doing calculations with angles, or a
judge dividing up a sum of money as an inheritance, needs the answer in a usable
form. Thus, a standard chapter in many arithmetics is one explaining how to express
a fraction a/b in terms of some other subunits l/c, where c is a number appropriate to

Fig. 5
38 2 Arithmetic in the Islamic World

what is being measured. For example, if we were measuring lengths in feet and
inches we would take c = 12, but Kūshyār proceeds to solve this problem for
c = 60.
Of course, if 36/243 = n/60, then n = 36  60/243, and this division produces a
quotient of 8 and a remainder of 216. So, if we think of the first remainder as being
the dirhams left after the division of 5625 dirhams among 243 people, then each
person’s share would be 23 dirhams 8 fulūs, with 216 fulūs left over. Or, we could
think of it as the division of an angle into 243 equal parts, so that each part would be
23°8′ with 216′ left over. This operation of multiplying a fraction, as 36/243, by 60
the Islamic authors called “raising,” and it was used to obtain base-60 expansions of
the fractional parts in a division. It is the analog of what we do to convert a fraction
to percent.

3 The Arithmetic of Common Fractions

Common fractions appear in Kūshyār’s work, as we have seen, in the context of


division, where the division of 5,625 by 243 leads to a quotient of 23 and a
remainder, 36, being written above the divisor. And Kūshyār refers to the quotient
as 23 and 36 parts of 243 [parts] making up the unit. However, common fractions
arise in practical problems of science, commerce, finance, law, etc. and practitioners
(and their teachers) developed a number of ways of relating such fractions to
simpler common fractions, and such fractions were discussed in numerous works of
arithmetic aimed at a wide audience. One such book, which was much studied in
the Maghrib, was the Book of Demonstration and Reminder, written by the twelfth
century the mathematician Abū Bakr Muḥammad al-Ḥaṣṣār.1 So far as is known it
was he who introduced the (now usual) notation of a horizontal bar to separate the
numerator of a common fraction from its denominator, a notation adopted by all
subsequent authors of arithmetic texts in the Maghrib. In his writings we also meet
for the first time a fivefold classification of fractions, which we shall discuss below.
The works on arithmetic were teaching texts, and teachers have always found it
useful to systematize knowledge in such a way that students can gain a view of the
whole subject. Thus, al-Ḥaṣṣār describes five kinds of fractions, the most basic
being simple fractions, i.e., the nine unit fractions 1 2, 1 3 , …, 1 10. Then there were
fractions related to others, an example being fractions of the form dc þ ab  d1, which
were written
ac
:
bd

1
His other work, also on arithmetic, was titled Al-Kāmil (The Complete/Perfect).
3 The Arithmetic of Common Fractions 39

For example, the related fraction

54
89

would be read, “four ninths and five-eighths of a ninth.”


Writers could—and did—extend that notation to relate three or more fractions in
the same manner, and the related fraction

154
289

would be read four ninths and five times an eighth of a ninth and one half of an
eighth of a ninth.
The exact dates of al-Ḥaṣṣār are not known, but it was likely about a century
after he was born that the mathematician Aḥmad Ibn al-Bannā’ was born in
Marrakesh in 1256. During his lifetime he achieved sufficient fame as a skilled
mathematician and astronomer/astrologer to be invited to the capital, Fās (Fez), a
number of times by the Merinid sultans. (He was also famous as a mystic and
magician, who could work wonders!) He died in 1321, leaving fourteen books on
mathematics alone and a number of students who continued his work.
The two mathematical works for which Ibn al-Bannā’ is best known today are A
Summary Account of the Operations of Computation and Raising the Veil on the
Various Procedures of Calculation. The first was evidently too much of a summary
for some readers, since the second is a commentary on the first, expanding on the
material in it. The great Tunisian historian, Ibn Khaldūn, who was born very shortly
after Ibn al-Bannā’’s death and was well acquainted with mathematics, said of Ibn
al-Bannā’’s commentary, “It is an important work. We have heard our teachers
praise it, and it deserves that.”
In his Summary Ibn al-Bannā’ follows the classification we find in al-Ḥaṣṣār and
explains that to find the numerator of a related fraction
ac
bd

one multiplies the number written above the first denominator2 by the second
denominator and adds the product to the number above that denominator.3 Thus, in
the case above, the numerator would be cb + a. He then states the general rule:
“One multiplies what is above the first denominator by the following denominators,
that which is above the second by the denominators following it, and so on to the
end of the line. Then one adds these products.” So, in the numerical example above,

Since Arabic is read from right to left, ‘first’ denominator means the rightmost.
2

3
We have relied on Souissi (1969).
40 2 Arithmetic in the Islamic World

154
289

the numerator would be 4  8  2 + 5  2 + 1, i.e., 75. The denominator, of course,


would be 9  8  2, i.e., 144.
He then proceeds to give rules for dealing with fractions mixed with whole
numbers. For example, in the case of an expression such as
ac
n
bd

where n is a whole number, his rule amounts to calculating the numerator as


n  (b  d) + (a + c  b), that is, to say one interprets the expression as meaning the
sum of n and the fraction
ac
:
bd

But, if the whole number, n, is found on the other side of the related fraction the
numerator is the product of n and that of the related fraction.
Then there were different fractions, which were simply sums of fractions, ab þ dc ,
partitioned fractions, which was the term for products, ab  dc , and, finally, fractions
separated by a sign of subtraction. (That sign was the Arabic word illā (“except”)
prefixed to the fraction being subtracted.
In the case of different fractions he obtains the numerator as we do, by multi-
plying each numerator by the other denominator and adding the two results. For
partitioned fractions Ibn al-Bannā’ writes, ‘One multiplies the numbers written
above the line by each other,’ a clear reference to the horizontal fraction bar. For the
case of subtraction he says that one proceeds as for addition but then subtracts the
smaller product from the larger.
The exposition of the theory and practice of calculating with fractions was, as the
above exposition hints, one that received considerable attention. Thus, Aḥmad b.
Mun‘im, who died in Marrakesh in 1228 (and of whom the reader will learn much
more later), devoted nearly half of his large work, The Laws of Calculation, to the
topic of fractions.
Some one hundred fifty years after the death of Ibn al-Bannā’, an Andalusian
mathematician, ‘Alī b. Muḥammad al-Qalaṣādī, who died in Tunis in 1486 (only 6
years before the voyage of Columbus), wrote a work whose title is an obvious
reference to the commentary of Ibn al-Bannā’, namely Removing the Veil from the
Science of Calculation. As did Ibn al-Bannā’, al-Qalaṣādī also wrote on religious
topics and literature, but it is his commentary on Ibn al-Bannā’’s Summary Account
that concerns us here.
In Part I of his book, in his discussion of multiplication, he gives some rules for
multiplication which make one realize why, in Kūshyār’s Hindu Reckoning, for
3 The Arithmetic of Common Fractions 41

example, one finds halving and doubling treated as separate topics.4 Thus,
al-Qalaṣādī says
“To multiply a number by three, add it to its double,”
and
“To multiply a number by six, add it to half of its product by ten,”
“To multiply a number by seven, put a zero to its right and subtract its triple from its
product by ten,” i.e. 7a = 10a − 3a, the product 3a being calculated as above.

But he goes on with more complicated rules, such as multiplying a number by


twelve by placing the number directly below itself and then placing the number
again below the two first, but so that the units place of the lowest line is lined up
with the tens places of the two lines above. Add these three numbers and the result
will be the answer. The calculation of 147  12 would look like this:

147
147
147
1764

Al-Qalaṣādī explains division as decomposing the dividend into parts equal to


the divisor, and then, in Part I, Chapter 5, he applies it to the problem of factoring
numbers. He first gives the usual test for seeing if a number is divisible by 9, a test
he calls ‘reducing the number by 9.’ He adds that if an even number is divisible by
9 then it is also divisible by 6 and 3. But if 3 or 6 remain when one reduces a
number by 9, as with such numbers as 48 and 78, then it is only divisible by 3 and
6.
If none of this works, reduce it by 8, and his procedure makes it clear that he
knows multiples of 1,000 are divisible by 8. So he only needs to reduce a three-digit
number by 8. For example, since 174 leaves a remainder of 6 when divided by 8 so
does is 3174.
For reduction by 7 his process mirrors long division, although his description of
it is interesting. He says, “Think of the leftmost digit as tens and add it to the digit to
its right, considered as being units. Reduce the sum by seven. Then add the
remainder [after this reduction], thought of, again, as tens, to the next digit to the
right, and continue the reduction in this way.”
He applies this procedure to 5236, where the calculations go
5236 ! 336 ! 56, where we have underlined the remainders after the successive
division by 7. One concludes 5236 is divisible by 7.

4
Interestingly, however, according to Djebbar (1992) the treatment of doubling and halving as
separate topics in Arabic arithmetic was, after al-Ḥaṣṣār, dropped in the Maghrib and the topics
were dealt with as special cases of multiplication and division.
42 2 Arithmetic in the Islamic World

All of this comes together in a problem in which he shows the reader how to
express a fraction as a related fraction. His rule is to express the denominator as a
product of factors and write those factors in a row (in descending order from right to
left) and place a line over the factors. The divide the numerator by these factors, one
after another. “You will obtain the result sought.” As with so much mathematical
instruction, this somewhat cryptic rule becomes clear with an example in which
al-Qalaṣādī shows how to express 75/144 as a related fraction.
“And if someone says to you, ‘Denominate seventy-five according to one
hundred forty-four,’ you decompose the denominator into nine, eight, and two and
divide the numerator first by two, obtaining 37 with a remainder of one, which you
put above the two. Then divide the quotient by eight to get four [with a reminder of
five, which you put above the eight].5 Put the four above the nine. The result will be
four-ninths and five-eighths of a ninth and a half of an eighth of a ninth. Write it as
follows:

1 5 4 00
:
289

And, in following through his reasoning, one can see how this notation for
fractions, so different from ours, might have arisen.
In Part II, Chapter 4 al-Qalaṣādī treats the division of one expression involving
connected fractions by another of the same type. His rule is to form the product of
the numerator of each of the two fractions by the factors [of the denominator] of the
other, and then divide the product of the dividend by that of the divisor, after having
decomposed this (latter product) into its factors. He then gives the following
example: Divide 3 4 and 5 7 of 1 4 by 2 5 and 6 7 of 1 5.
Thus

53
74

is to be divided by

62
:
75

The numerator of the dividend is 26 and the factors of its denominator are 7 and
4. The numerator of the divisor is 20 and the factors of its divisor are 7 and 5. One
must then form the products 26  7  5 (= 910) and divide it by the products
20  7  4 (= 560). Decompose the latter number into its factors, which are 10, 8,
and 7 and divide 910 by these factors. He expresses the answer as

5
At some point in the history of this text, the bracketed phrase I have inserted, clearly essential,
was left out.
3 The Arithmetic of Common Fractions 43

02 6
1
7 8 10

in other words, one and six-tenths and two-eighths of a tenth.


One is reminded of Kūshyār’s procedure in the example we gave earlier of
calculating in the sexagesimal system. In that case, it was converting the sexa-
gesimal numbers to whole numbers expressed in the decimal, positional system
then doing the calculation in that system, and finally converting the answer back to
sexagesimals. Here, one converts the related fractions to ordinary common frac-
tions, then calculates the quotient with them, and finally converts the result back to
related fractions.

4 The Discovery of Decimal Fractions

Today we use not sexagesimal but decimal fractions to represent the fraction
remaining after a division, and it now appears these were a contribution of the
Islamic world. Evidence for this claim is contained in The Book of Chapters on
Hindu Arithmetic, written in Damascus in the years A.D. 952–953 by Abū al-Ḥasan
al-Uqlīdisī. The name “al-Uqlīdisī” indicates that the author earned his living
copying manuscripts of Euclid (“Uqlīdis” in Arabic), but beyond this we know
nothing of the life of a man who seems to have been the first to use decimal
fractions, complete with the decimal point, and therefore the first to write numbers
as we do. Since al-Uqlīdisī specifically states in the preface to his book that he has
taken great pains to include the best methods of all previous writers on the subject,
it is hard to be sure that the decimal fractions were his own discovery, but their
complete absence in Indian sources makes it fairly certain that they were a dis-
covery of Islamic scientists.
Al-Uqlīdisī is also proud of the fact that he has collected ways of performing on
paper, with ink, algorithms usually performed by arithmeticians on the dust board,
and in his Book of Chapters he gives the following reasons for abandoning the dust
board in favor of pen and paper.
Many a man hates to show the dust board in his hands when he needs to use this art of
calculation (Hindu arithmetic) for fear of misunderstanding from those present who see it in
his hands. It is unbecoming him since it is seen in the hands of the good-for-nothings
earning their living by astrology in the streets.

It seems that the street astrologers could be recognized by their use of the dust
board, and al-Uqlīdisī urges the use of pen and paper to escape being taken for a
mendicant fortune teller.
Al-Uqlīdisī’s text contains four parts, of which the first two deal with the ele-
mentary and advanced parts of Hindu arithmetic, and it is in the second part where
decimal fractions first appear. This is in the section on doubling and halving
numbers, where he introduces them as one of the three ways of halving an odd
44 2 Arithmetic in the Islamic World

number. The first way is the one described by Kūshyār who, to halve 5625, con-
sidered it as degrees or dirhams and wrote the result as in Fig. 3, where the lower
30 could be interpreted as fulūs or minutes. The second way is one al-Uqlīdisī calls
numerical, and describes as follows:
… halving one in any place is five (in the place) before it, and this necessitates that when
we halve an odd number we make half of the unit five before it and we put over the units’
place a mark by which we distinguish the place. So the value of the unit’s place is tens to
that before it. Now the five may be halved just as whole numbers are halved, and the value
of the units’ place in the second halving becomes hundreds and this may continue
indefinitely.

When al-Uqlīdisī writes of places in a numeral being “before” other places he is


referring to the direction of Arabic writing, which is from right to left. Thus in 175
the 5 would be before the 7. As an example of what he has explained al-Uqlīdisī
gives the results of halving 19 five times as 059375, where, he says, “the place of
the units is hundred-thousands to what is in front of it.” Figure 6 shows the Arabic
text of al-Uqlīdisī’s work and the use of the decimal point in the form of the short
vertical mark pointing out the unit’s place. The decimal mark is clearly visible
above the ‘2’ in the middle of line 10 of the text in Fig. 6 and over the ‘9’ at the left
end of that line. Using the forms of the numerals given earlier, the reader will have
no trouble identifying the various numerals in that figure.
From a purely mathematical point of view it is especially satisfying to see
decimal fractions, complete with decimal point, explained by reasoning by analogy
from established procedures. The usual procedure for halving an even number, such
as 34, was to begin by halving the units, so 34 ! 32. Then, as a writer like
Kushyar would have put it, “The three is tens of the two to the right of it, so its half
is fifteen. We add the five to the two, which is its units, and it becomes seven. So
the result is 17.” The principle used here was that half a unit in one place (tens,
hundreds, etc.) was five in the place to the right. Al-Uqlīdisī observed that the same
principle could be applied to halving a number with an odd digit in the unit’s place,
and out of such a simple observation came a very useful mathematical tool.
A little later, al-Uqlīdisī again uses decimal fractions, this time to increase 135
by its tenth, then the result by its tenth, etc. five times. Thus he sets out as in Fig. 7
 
1 5
to calculate 135 1 þ 10 . He writes 135 and below it 135 again, but moved one
1
place to the right. This will be 10  135, so he adds it to 135. In the sum
1
135 + 10  135, he marks the unit’s place with a short vertical line above it. When
he has shifted and added four more times, the result will be the desired quantity. (He
mentions that the value of the lowest place is hundred-thousandths.)
He gives an alternative to this method as follows (where we use the decimal
point for al-Uqlīdisī’s vertical line):
 
1 ð135  11Þ
135  1 þ ¼ ¼ 148:5
10 10
4 The Discovery of Decimal Fractions 45

Fig. 6

Fig. 7
46 2 Arithmetic in the Islamic World

and
     
1 ð148:5  11Þ 11 11
148:5  1 þ ¼ ¼ 148:5 þ 0:5 
10 10 10 10
¼ 162:8 þ 0:55 ¼ 163:35;

which shows that al-Uqlīdisī not only added decimal fractions but multiplied them
by whole numbers as well, even though his method of multiplication unnecessarily
splits the number into its whole and fractional parts.
Less than half a century later another Muslim author, Abū Manṣūr al-Baghdādī,
used decimal fractions—also in a problem on computing tenths. He represented
0
what al-Uqlīdisī would write as 17 28 by 08 02 17, but each pair written above the
previous one, in strict analogy to Kūshyār’s notation for sexagesimal fractions.
Al-Uqlīdisī’s use of decimal fractions is something of an ad hoc device,
unsystematized and unnamed. Two centuries later, however, one finds in the
writings of al-Samaw’al, whose work we discuss in the chapter on algebra, the use
of decimal fractions in the context of division and root extraction. In a treatise of
1172, al-Samaw’al introduces them carefully, as part of a general method of
approximating numbers as closely as one likes. Thus al-Samaw’al uses decimal
fractions within a theory rather than as an ad hoc device, although he still has no
name for them and his notation is inferior to that of al-Uqlīdisī. The reader will find
the details in Rashed.
It is in the early fifteenth century that decimal fractions receive both a name and a
systematic exposition. By then Jamshīd al-Kāshī displays a thorough command of
the arithmetic of decimal fractions, for example, multiplying them just as we do
today. It is also in the fifteenth century that a Byzantine arithmetic textbook describes
as “Turkish” , i.e., from the Islamic world, the method of representing 153 12 and 16 14
by 153|5 and 16|25 and their product by 2494|375. (See Hunger and Vogel.)
It was not until over a century later that the European writers began using
decimal fractions. An able publicist for the idea was the Flemish engineer Simon
Stevin, whose book The Tenth was published in 1585. However, his awkward
notation was nowhere near so good as al-Uqlīdisī’s, and it was left to the Scot, John
Napier, to reinvent the decimal point and use decimal fractions in his table of
logarithms, another invention of his.

5 Muslim Sexagesimal Arithmetic

5.1 History of Sexagesimals

Although the student may think it strange that it took almost 500 years (from the
tenth to the fifteenth centuries) for decimal fractions to develop, it must be
remembered that Muslim scientists, from the ninth century onwards, already
5 Muslim Sexagesimal Arithmetic 47

possessed a completely satisfactory place-value system to express both whole


numbers and fractions. It was not decimal, however, but the sexagesimal system we
have already referred to, in which the base is 60, and it arose out of the fusion of
two ancient numeration systems.
The first of these is one used by the Babylonians around 2000 B.C in
Mesopotamia. As we know it from the many surviving cuneiform texts, it was a
positional system, in which the successive places of a numeral represent the suc-
cessive powers (positive and negative) of the base, 60, so it treated whole numbers
and fractions in a unified manner. However, the Babylonians did not use single
ciphers for the fifty-nine digits from 1 to 59, but formed them by repeating the
wedges for 1 and 10 . Thus, the Babylonians would represent the integers 3,
25, 133 and 3753 as

In addition, they extended the system to include fractions. Thus since 1


2 ¼ 30
60
they, would write 12 as 30 also and since

7 70 60 10 1 10
¼ ¼ þ ¼ þ
360 3600 3600 3600 60 602

it would be written as .
There was always a possibility of misunderstanding in using this system, for
there was no special mark to indicate the units (that is, there was no sign like our
decimal point and no custom of writing final zeros in an integer), and therefore, the
magnitude of the number was determined only up to a factor of some power of 60.
Thus, although could represent 1 13, it could also represent 80. One step towards
clarification was taken late in the fourth century B.C., at the time when Babylon
was ruled by the successors of Alexander. At that time, scribes in Babylon began to
write numbers more frequently with a special symbol to indicate zeros within the
numeral, so it was possible to write 71 in such a way as to distinguish it unam-
biguously from 3611 (71 being written as and 3611 as ).
These imperfections, however, are relatively minor and seemed not to have
caused much difficulty for the Babylonians. Much more important is the existence,
two millennia before our era, of a numeration system so well suited for complex
calculations that Greek astronomers, at some time during or after the second century
B.C., adopted it for their calculations. Thus, the astronomer Ptolemy used it in the
mid-second century A.D. in his Greek astronomical handbook The Almagest.
The Hellenistic Greeks’ adoption of the system was, however, rather an instance
of grafting than of transplanting; for, while they used it with a different notation to
represent the fractional part of a number, they retained their own method of rep-
resenting the integral part. This method is an example of the second ancient system
we referred to earlier, in which 27 letters of an alphabet are used to represent the
numbers 1,…, 9; 10, 20,…, 90; 100, 200,…, 900. In the case of the Greeks, they
48 2 Arithmetic in the Islamic World

used 27 letters of an archaic form of the Greek alphabet, according to the scheme
below:

This ancient alphabet stems from that of the Phoenecians, a Semitic people to
whom we owe the inventions of the alphabet and of money. The alphabetic system
of numeration seems to have been common to many of the peoples of the
Mediterranean. Thus, it was used not only by the Greeks and Arabs, but also by the
Hebrews and others.
In this system the numbers up to 999 would be represented by a string of letters,
so that, in the case of the Greeks 48 and 377 would be written MH and TOZ. We
need not go into the special devices that were used to represent numbers larger than
999, for it is the fractions that interest us now. A Greek astronomer, knowing the
Babylonian system, evidently saw the possibility of substituting letters of the
alphabet for the groups of wedges the Babylonians used for digits. Thus 12 13 would
be written IB K, to signify (10 + 2) + 20 60.
The Greeks, however, adapted the Babylonian place-value system only for
fractions, so they wrote PMB IB for 142 15 rather than the more consistent B KB IB
(i.e. 2  60 + 22 + 12 60). The only improvement the Greek system displayed was a
slight cipherization for the digits, so that whereas the Babylonian would have to
write for 13, the Greek could simply write K.
The real transplant of the Babylonian system was done by Islamic mathemati-
cians, in a system that was so widely used by astronomers that it simply became
known as “the astronomers’ arithmetic.” In it, the 28 letters of the Arabic alphabet
were used in an order quite different from their order in the alphabet as it was (and
is) written. If we transcribe these letters according to the system in Haywood and
Nahmad then the correspondence between letters of the Arabic alphabet and
numerals is that shown in Fig. 8. (Although the system extends to 1000—for the
28th letter—there is no need for letters beyond the nūn (50) in the astronomers’
arithmetic, since no digit can be greater than 59. The only fact we need to add is
that, as with the Greeks, “zero” was represented by or , which are two versions
of the same cipher.)
5 Muslim Sexagesimal Arithmetic 49

Fig. 8

Thus, if we represent Arabic letters by the corresponding Latin ones in Fig. 8,


the numeral 84 would be written a kd (i.e., 1  60 + 24), and lb n would represent
32 50
60. These two examples illustrate how the Muslims made a consistent adaptation
of the Babylonian system to their own mode of writing, in the process of which they
introduced a significant amount of cipherization. Of course, there remained the
ambiguity of the value of any given numeral. Although b mh could represent 165
(=2  60 + 45), it could equally well represent 2 45 60, and, in the absence of a
sexagesimal point, some other device was necessary to eliminate this ambiguity.
There were two solutions to this problem. The one was to name each place, so
that the nonnegative powers of 60 (1, 60, 602,…) were called “degrees,” “first
elevates,” “second elevates,”…, while the negative powers of 60 (1/60, 1/602,
1/603,…) were called “minutes,” “seconds,” “thirds,”…. The origin of the name
“degrees” is in astronomy, where the term referred to the 360 equal parts into which
the zodiac circle is divided. The term “minutes” is a translation of the daqā’iq,
which means “fine,” just as the English word “minute” does. The succeeding fine
50 2 Arithmetic in the Islamic World

Fig. 9

parts were, naturally, “the second, third, etc. fine parts.” The other solution was to
name the last place only, so that “b mh minutes” would make it clear that the value
2 45
60 was intended.
In the following survey of Muslim sexagesimal arithmetic we shall follow the
second section of Kūshyār’s Principles of Hindu Reckoning, and it is typical of the
variety of approaches used by Muslim scientists that, although Kūshyār explains a
consistent sexagesimal arithmetic, he does not use letters of the alphabet at all, but
rather the form of the Hindu ciphers used in the Eastern caliphate. Thus, what some
writers would express as ka h mb, Kūshyār writes as in Fig. 9, where the places of
the numeral are written vertically in order to prevent confusion with the Hindu
numeral 210,542. However, here as earlier he uses the ciphers only when he is
actually showing the work. Elsewhere he writes out all the numerals longhand, and,
to give some of the flavor of the work, we shall follow the same practice.

5.2 Sexagesimal Addition and Subtraction

To illustrate addition, Kūshyār gives the following example: “We wish to add
twenty-five degrees, thirty-three minutes and twenty-four seconds to forty-eight
degrees, thirty-five minutes and fifteen seconds.” He sets these two numbers down
in two columns, separated by an empty column, with degrees facing degrees,
minutes facing minutes, and seconds facing seconds (Fig. 10). He then adds
twenty-five to forty-eight, tens to tens, and units to units, and then he adds
thirty-three to thirty-five and twenty-four to fifteen. Whenever a sum exceeds sixty
he subtracts sixty from it, enters the result, and adds one to the place above it. This
is the reason for the upper “one” shown in the second figure. The dust board where
Kūshyār imagined these calculations being carried out would show the successive
parts of Fig. 10, with only the last set of figures showing at the end.
Subtraction, too, is straightforward, and it proceeds from the highest place
downwards, with borrowing when necessary. Figure 11 shows the process of
subtracting rather than adding in the above example, and it clearly offers no serious
difficulties.
5 Muslim Sexagesimal Arithmetic 51

Fig. 10

Fig. 11

5.3 Sexagesimal Multiplication

5.3.1 Multiplication by Leveling

Multiplication and division were, however, another matter. Even such able math-
ematicians as al-Bīrūnī found it most convenient to convert the sexagesimal
numerals to decimal form, perform the computations on the decimal forms by the
rules of Hindu arithmetic, and then convert the answer back to sexagesimals, and
the procedure was so common it was given a special name, “leveling”.
A contemporary of al-Bīrūnī, al-Nasawī, solves the problem of multiplying the two
sexagesimal numbers 4°15′20′′ and 6°20′13′′ in the following way. First, he
expresses both factors in terms of their lowest orders, Thus:

4 150 ¼ ð4  60Þ0 þ 150 ¼ 2550 ; and 2550 2000 ¼ ð255  60Þ00 þ 2000 ¼ 15; 32000 :

Similarly, he calculates the other factor to be 22,813.′′ Since the books that
discuss this method explain how to calculate the products of various orders,
al-Nasawī knows that the product of “seconds” by “seconds” will be on the order of
“fourths” and, calculating in pure decimal numbers, he finds the product to be
349,495,160 fourths. Now it is necessary to perform the inverse operation of
leveling, namely to “raise” this number to a sexagesimal expression, by dividing by
60. (We saw an example of this at the end of the treatment of division in the section
on decimal arithmetic.) Thus, in this case,

349; 495; 1600000 ¼ ð5; 824; 919  60 þ 20Þ0000 ¼ 5; 824; 9190000 þ 200000 :

Finally, proceeding as above, but now with the thirds, then the seconds, and
finally the minutes, al-Nasawī obtains the answer 26°58′1′′59′′′20′′′′.
52 2 Arithmetic in the Islamic World

The foregoing, inelegant procedure was widespread but by no means universal.


Kūshyār, although he mentions it in his treatise as one method, explains how to
multiply two sexagesimal numbers without any such conversion.

5.3.2 Multiplication Tables

At the beginning of his section on sexagesimal arithmetic, Kūshyār describes a


sexagesimal multiplication table, which consists of 59 columns, each headed by one
of the integers from 1 to 59, and each containing 60 rows. The column headed with
the integer 39, for example, contains in its rows the multiples of 39, from 1  39 to
60  39. Although Kūshyār’s book has no such table, examples of these tables have
survived in other treatises (See King and Plate 1.) The rightmost column of each
page in such a table is headed “the number” and contains the alphabetic numerals,
those from 1–30 usually appearing on the right-hand page and those from 31 to 60
on the left. The succeeding columns (going from right to left, as in Arabic writing)
are headed by the alphabetic numerals between 1 and 60. (Of course only a certain
number of them appear on each page, for reasons of space.) Each column gives, as
we mentioned above, the first sixty multiples of the integer that stands at the top,
and in general these multiples will need two sexagesimal digits to express them. For
example, the product of 13 (ig) by 8 (ḥ) would be written with the two-digit
numeral we transliterate as a md. The first twelve rows of the three rightmost
columns in Plate 1, transliterated and then translated, are shown in Fig. 12.
The eighth row below the heading, for example means that 8 13 = 1 44 (104) and
8 14 = 1 52 (112). (We use the convention that n m; r s means n  60 + m +
r/60 + s/602. Another common convention separates the sexagesimal digits by
commas, as n, m; r, s.)
A remarkable example of a multiplication table was compiled, probably by a
Turkish astronomer, around the year 1600 and gives the first 60 multiples of each
two-place sexagesimal number from 00 01 to 59 59, so one can find directly from
the table such products as 14 34  19 = 4 36 46. The table has 212,400 entries and
fills a ninety-page booklet. Other astronomers, doubtless, found it more convenient
to use the more limited tables and compute other products as needed by one of the
algorithms we shall now describe.
The first of these differs only slightly from the method Kūshyār gives for the
multiplication of two decimal numbers. In the sexagesimal case, the numbers are
written vertically rather than horizontally, with an empty column left between them
to contain the product, and Kūshyār’s procedure for the product of 25°42′ by 18°36′
is shown in Fig. 13.
5 Muslim Sexagesimal Arithmetic 53

Plate 1 Part of a sexagesimal multiplication table. The rightmost column is headed “the number”
and shows the alphabetic numerals from 1 to 12. The succeeding columns (from right to left, as in
Arabic handwriting) are headed by the numerals 13, 14, …, 18 and the entries underneath them
give their multiples expressed as two-place sexagesimals. (See Fig. 12 for a transliteration and
translation of the rightmost three columns of this table.) (Photo courtesy of the Egyptian National
Library.)
54 2 Arithmetic in the Islamic World

* An error for NB on the part of the scribe.

Fig. 12

Fig. 13

5.3.3 Methods of Sexagesimal Multiplication

The first two steps, Kūshyār specifies, are done with the aid of the multiplication
table for 18, and since the 30 in the first step and the 12 arising in the second are of
the same order they must be added in the product, so 30 is replaced by 42.
Since the product of minutes by minutes is seconds, the answer is 7158°1′12′′
(that is, “7 first elevates, 58 degrees, 1 minute and 12 seconds”). In the last two
steps, 36 is one place lower than 18 so its products with 25 and 42 must be added to
the column where the answer is taking shape, but one place lower than the corre-
sponding products for 18.
5 Muslim Sexagesimal Arithmetic 55

Fig. 14 The gelosia method of multiplication as described by al-Kāshī, Used for both decimal and
sexagesimal systems it is shown here in the sexagesimal system. The answer is obtained by adding
the entries in each of the eight columns within the square (base 60)

A method, that was popular both in Islam and the West for multiplication in a
positional system, is illustrated in Fig. 14 with an example from Jamshīd al-Kāshī’s
Calculators’ Key. The problem is to multiply 13 09 51 20 minutes by 38 40 15 24
thirds, and since the largest number of places is four a square is subdivided to form
a lattice of 16 subsquares, each of which is divided as in Fig. 14 into two equal
triangles. On the edges of the square that intersect at the top corner, the two factors
are written so that the term of lowest order in one factor and that of highest order in
the other factor are put at the top, each term of both factors being labeled by its
orders. Then each square is filled in with the product of the two numbers on the
outer edges opposite its sides, so that when this product has two ciphers the cipher
of the highest order is put on the left of the square. For example, since 38  13 = 8
14 the 8 will be put in the left part of the left square and the 14 will be put in the
right. When all 16 products are computed, the answer is obtained by adding up the
ciphers in each of the eight vertical columns of the square, and the sums are written
underneath. Since “minutes times thirds” is on the order of fourths, the lowest order
of the product is fourths.
Although a certain amount of work is necessary to prepare the grid, it is then
easy to fill in the lattice by means of a multiplication table, and the only compu-
tation involved is adding up the entries in the columns. Also, the squares can be
filled in any convenient order, since the lattice-work keeps everything arranged.
The source from which we have taken this method, al-Kāshī’s The Calculators’
Key, gives no proof of the validity of the method; however, the proof is easy when
one notices that what is put in the left-hand side of each square is precisely what
would be carried and added to the next product in the method we are used to. The
lattice does this carrying automatically, but what is carried is added to the product
after all the multiplications have been done, rather than during the process as we are
used to.
56 2 Arithmetic in the Islamic World

5.4 Sexagesimal Division

Finally, the method Kūshyār uses for division in the sexagesimal system parallels
that used in the case of multiplication. Hence, to divide 49°36′ by 12°25′, Kūshyār
arranges three columns and proceeds as in Fig. 15. Here

49  3  12 ¼ 13 and 13 36 - 3  25 ¼ 13 36  1 15 ¼ 12 21

so each digit of the divisor is multiplied by the digit of the quotient (obtained by
trial, as we do), and the result is subtracted from that part of the dividend that is
above it (or level with it) and to the right. Finally, the divisor is shifted down one
place, and this is done so that the next digit of the quotient, when it is placed level
with the highest entry of the divisor, will be of the correct order (in this case
“minutes”). After the third step, the question now becomes “12°25′ times how
many minutes produces something not exceeding 12°21′?”, and the answer is
“59 minutes.” The last two steps of Fig. 15 show the final working-out. Again the
general rule is that the product of a digit of the quotient by a digit of the divisor is
subtracted from all of the dividend to the right and above (or level with) the digit of
the quotient.
Thus Kūshyār gives the result as 3°59′ and says, “If we wish precision we copy
the divisor one place lower.” Hence, the result could be continued to as many
sexagesimal places as necessary. Also, Kūshyār remarks that he has attached to his
book a table giving “the results of the division,” that is, the order that results when a
number of one order (say “first elevates”) is divided by that of another (say
“thirds”). Finally, Kūshyār concludes his chapter with a discussion of how to
calculate square roots in the sexagesimal system, an operation of some importance
to astronomers.
Thus, there was widespread in the Muslim world a consistent system of sexa-
gesimal arithmetic that permitted a unified treatment of both whole numbers and
fractions. This system was supported by special tables, and it provided an approach
to all the operations of arithmetic which was every bit as satisfactory as that of the
(initially) less-developed system of decimal fractions.

Fig. 15
6 Square Roots 57

6 Square Roots

6.1 Introduction

Instead of following Kūshyār’s presentation of the extraction of square roots, we


shall follow that of Jamshīd al-Kāshī in his book The Calculators’ Key, which we
have already referred to as the work he wrote in Samarqand two years before his
death in 1429. It is a compendium of elementary mathematics, including arithmetic,
algebra, and the geometry of measurement, which contains a thorough treatment of
decimal fractions, a table of binomial coefficients, and algorithms for extracting
higher roots of numbers. For example, we shall see later how he works out the fifth
root of a number on the order of trillions, namely 44,240,899,506,197.
The following list of titles of the five main chapters of The Calculators’ Key
shows its differences from the work of Kūshyār: (1) On the arithmetic of whole
numbers. (2) On the arithmetic of fractions (including decimal fractions). (3) On the
arithmetic of astronomers (sexagesimal). (4) On the measurement of plane and solid
figures. (5) On the solution of problems by algebra.

6.2 Obtaining Approximate Square Roots

We shall first see how al-Kāshī extracts the square root of 331,781. His method for
the square root is the same as Kūshyār’s, but, unlike Kūshyār, al-Kāshī was writing
for people who would use pen and paper. (It was in Samarqand where the Arabs first
learned of papermaking from Chinese prisoners of war near the end of the eighth
century and, because of its abundant supply of fresh water, Samarqand remained a
center of paper manufacturing for several centuries.) Thus, in the method as al-Kāshī
explains it, none of the intermediate steps are erased. Al-Kāshī organizes his work by
dividing the digits of the radicand, 331,781, into groups of two called “cycles,”
starting from the right. (Thus 331,781 is divided as 33 17 81.) As al-Kāshī explains
it, since the numbers 1, 100, 10,000 … have integer square roots (unlike 10, 1000,
…), the cycles are relevant, for the first (81) counts the number of units, the second
(17) the number of hundreds, the third (33) the number of 10,000’s, etc. He then
draws a line across the top of the radicand and lines down the paper separating the
cycles. At the beginning, therefore, his paper looks like Fig. 16a.
To get the first of the three digits of his root he finds the largest digit n so that n2
does not exceed 33. Since 52 = 25 and 62 = 36 he takes n = 5, which is written
both above and some distance below the 33 (below the last 3) to obtain Fig. 16b.
Now he subtracts 25 from 33 to obtain 8, which he writes below 33, and draws a
line under 33 to show he is done with it. (On the dust board the 33 would be erased
and the 8 would replace it.) Now he doubles the part of the root he has obtained, 5,
and writes the result (10) above the bottom 5, but shifted one place to the right to
obtain Fig. 16c. (The dust board would only show the top 5, the middle 8 17 81,
58 2 Arithmetic in the Islamic World

Fig. 16

Fig. 17

and the bottom 10 of Fig. 16c.) At this stage al-Kāshī has a current answer (5) at the
top and double the current answer (10) at the bottom.
What al-Kāshī next asks is to find the largest digit x so that (100 + x)  x  817.
Experiment shows x = 7, and he writes 7 above the 7 of 17, and next to the 10 on
the bottom, and then he performs the computation of (100 + 7)  7 = 749 and
subtracts the result from 817 to get Fig. 17a. He now begins the process again,
doubling the last digit in 107 to get 114 and writing this above the 107, but shifted
one place to the right, as shown in Fig. 17b. Once again he has the current answer
(57) at the top and double that (114) at the bottom, and as before the question is
this: What is the largest digit x so that (1140 + x)  x  6881? A trial division of
6 Square Roots 59

688 by 114 suggests trying x = 6. This works, and after 1146  6 has been sub-
tracted from 6881, the last digit of 1146 is doubled to make the 1146 into 1152
(=1146 + 6) as in Fig. 17c. (The dust board would show the top 576, the middle 5
and the bottom 1152.)
Thus al-Kāshī has obtained an approximate square root (576) as the current
answer and double that (1152) at the bottom. Finally, he increases 1152 by 1 to get
1153 and divides it into the remainder, the middle 5, to obtain, as the approximate
5
square root of 331,781, the number 576 1153 (=576.00434). Calculation shows the
square of the latter number is 331,780.996, so al-Kāshī’s result is quite close.

6.3 Justifying the Approximation

Two questions arise: (1) What is the justification for al-Kāshī’s procedure for
obtaining the integral part of the root; and (2) What is the justification for the
fractional part? We will begin with the second question.

6.3.1 Justifying the Fractional Part

In fact, the numerator of the fractional part, 5, is equal to 331,781 − (576)2, and the
denominator, 1153, is 5772 − 5762. This is because 1153 is one more than “twice
the current answer”, i.e.

1153  1 þ 2  576 ¼ ð1 þ 576Þ2  5762 :

5
Thus the fractional part of the answer, 1153 , is just that obtained by linear
interpolation, i.e., (331,781 − 576 )/(577 − 5762), a technique that was ancient
2 2

when Ptolemy used it in his Almagest in the first half of the second century A.D.
To understand this technique as a medieval astronomer might have justified it,
imagine a table of square roots obtained by listing in one column the successive
squares from 12 to 1,0002, and, next to these in a second column, the first thousand
whole numbers, as in Fig. 18.

Fig. 18
60 2 Arithmetic in the Islamic World

Fig. 19

pffiffiffi
Then, to find 6, the simplest procedure would be to observe that 4 < 6 < 9
pffiffiffi
implies 2; \ 6\3. Moreover, since 6 − 4 = 2 and 9 − 4 = 5, 6 is 25 of the way
pffiffiffi pffiffiffi pffiffiffi
between 9 and 4. Thus 6 is about 25 of the way between 4 ¼ 2 and 9 ¼ 2, i.e.,
pffiffiffi
6 ¼ 2 25, approximately.
pffiffiffi
The reader will recognize that this reasoning is based on the assumption that x
is proportional to x, which is the same as the assumption that, if we may express
pffiffiffi
ourselves in modern language, the function f ðxÞ ¼ x is linear, i.e. its graph is a
straight line. Although this is not true a glance at the graph of f(x) in Fig. 19 reveals
that it is nearly linear for x > 1 and over not too big an interval [a, b]. Thus, for
example, the straight line joining the two points (16, 4) and (25, 5) is hardly
distinguishable from the graph between these two points, and this is why the
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
technique gives such a good approximation to the fractional part of 331; 781. The
table shows 5762 = 331,776 < 331,781 < 332,929 = 5772 and, since
331,781 − 576 = 5 while 5772 − 5762 = 1153 we conclude that since
2

N = 331,781 is 1153 5
of the way between 5762 and 5772 its square root is about 1153 5
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
of the way between 576 and 577. Thus 331; 781 ¼ 576 1153. 5

Linear interpolation was one of the standard ancient and medieval methods of
what has been called “reading between the lines” (of tables), but we emphasize that
the name “linear interpolation” reflects modern ideas, and those who discovered this
method of approximation had no conception of a straight line as the graph of an
equation. The ancient and medieval concept was simply that in a table of values
pairing x to y one assumed that the change from y to y′ was distributed equally over
the units from x to x′.

6.3.2 Justifying the Integral Part


pffiffiffiffi
As for the extraction of the integral part of N , al-Kāshī knows that if N = abcdef
then the largest integer r with the property that r2  N has half as many digits as N,
6 Square Roots 61

in this case 3. (By taking a = 0 if necessary we may assume that N has an even
number of digits.) Therefore he divides N into what he calls cycles, as N = ab cd ef
and thereby considers N = ab  1002 + cd  102 + ef.
Now, al-Kāshī’s first step is to find the largest number A so that A2  ab. A will
be a one-digit number, since ab has two digits and 102 has three digits. Such an
A will be the first digit of the root, as the reader may verify.
His next step is to calculate the difference

D1 ¼ NðA  100Þ2 ¼ ðab  A2 Þ  1002 þ cd  102 þ ef

and then to find the next place, i.e., the largest B so that

D2 ¼ N  ðA  100 þ B  10Þ2  0

He uses the basic identity (X + Y)2 = X2 + (2X + Y)Y to expand

D2 as N  ðA  100Þ2  (2A  100 þ B  10ÞB  10 ¼ D1  ð2A  10 þ BÞB  100;

and the expression 2A  10 + B is the formal equivalent of al-Kāshī’s instruction to


double A, the previous digit of the root, and then put the digit (B) next to it. As
al-Kāshī says, this next digit is chosen to be the largest, so that the product
(2A  10 + B)B  100 does not exceed the previous difference Δ1. The multiplication
of 2A by 1000 instead of by 10,000 is reflected in its shift one place to the right. Of
course, al-Kāshī never mentions the powers of 10 since they are automatically taken
into account by the positioning.
The procedure should by now be clear. When we have determined B to be as
large as possible so that (A  100 + B  10)2  N we choose C to be as large as
possible so that 0  N − (A  100 + B  10 + C)2, and where, with
X = (A  100 + B  10) and Y = C, (X + Y)2 is again expanded according to the rule
(X + Y)2 = X2 + (2X + Y)Y. This identity, or its alternate form (X + Y)2 −
X2 = (2X + Y)Y is the basis for the algorithm for the extraction of the square root.
Al-Kāshī’s procedure also takes advantage of the fact that in evaluating
N − (X + Y)2 the part N − X2 has been evaluated at the previous step.

7 Al-Kāshī’s Extraction of a Fifth Root

7.1 Introduction

We now follow the beginning of al-Kāshī’s extraction of the fifth root of


44,240,899,506,197—a number on the order of trillions. The extraction of higher
roots of numbers was, according to the testimony of ‛Umar Khayyām, an
achievement of Muslim scholars, for he wrote in his Algebra,
62 2 Arithmetic in the Islamic World

From the Indians one has methods for obtaining square and cube roots, methods which are
based on knowledge of individual cases, namely the knowledge of the squares of the nine
digits 12, 22, 32 (etc.) and their respective products, i.e. 2  3 etc. We have written a treatise
on the proof of the validity of those methods and that they satisfy the conditions. In addition
we have increased their types, namely in the form of the determination of the fourth, fifth,
sixth roots up to any desired degree. No one preceded us in this and those proofs are purely
arithmetic, founded on the arithmetic of The Elements.

‘Umar was neither the first mathematician nor the last who believed falsely
that he was the originator of a method. In this case, we know that Abū al-Wafā’,
who flourished over 100 years before ‛Umar, in the late tenth century, wrote a work
entitled On Obtaining Cube and Fourth Roots and Roots Composed of These Two.
Of course, ‛Umar may not have known of Abū al-Wafā’s treatise, or it may be that
pffiffiffiffi pp ffiffiffiffiffiffiffi
ffiffiffiffiffi pffiffiffiffi
Abū a1-Wafā’ simply pointed out that 4 N ¼ ð N Þ and, since 3 N was already
pffiffiffiffi p ffiffiffiffiffiffiffi
p ffiffiffiffiffi
known from the Indians, roots such as the twelfth, for example, 12 N ¼
4 3
N could
be calculated by known methods. Thus Abū al-Wafā”s work may have been less
innovative than that of ‛Umar.

7.2 Laying Out the Work

However that may be, neither ‛Umar’s treatise nor that of Abū a1-Wafā’ is extant, so
we shall study al-Kāshī’s method from Book III of his Calculators’ Key. He begins
by instructing the reader to write the number across the top of the page and to divide
the number into cycles, which are, this time, successive groups of five digits
beginning from the right. This is because the powers of 10 with perfect fifth roots are
1, 105, 1010, etc. Next, al-Kāshī puts between the cycles, double lines and between
the individual digits single lines, all running down the length of the page, and then he
puts a line above the number, on which he will enter the digits of the root.
Next, he divides the space below the number into five broad bands by means of
horizontal lines. The top band contains the number, and the words “Row of the
number” are written on the edge of this band. The band below it is called “Row of
the square square” (the fourth power) number.” When this process is finished, the
sheet will look as in Fig. 20, and everything is ready. It seems not too far from the
algorithmic spirit of this procedure to look on the cells in Fig. 20 as locations in a
computer’s memory, and in keeping with this Fig. 21 shows a flow-chart for the
root extraction which the reader may find useful to get an overview of the process.

7.3 The Procedure for the First Two Digits

Al-Kāshī’ now proceeds as follows (Fig. 22). The largest integer, a, whose fifth
power does not exceed 4424 is 5, so he puts 5 in “Row of Result” (above the first
cycle) and at the bottom of “Row of Root.” Next, he puts 52 (25) at the bottom of
7 Al-Kāshī’s Extraction of a Fifth Root 63

Fig. 20

“Row of Square,” 53 (125) at the bottom of “Row of Cube,” and 54 (625) at the
bottom of the “Row of the Square Square”. Finally, 4424 − 55 = 1299 is placed in
the “Row of Number” (This number, in virtue of its position, represents 1299 1010.)
Next he begins the process called “once up to the row of the square square,” by
adding the latest entry in “Row of Root” (5) to the most recently obtained digit of
the root (5) and writing the sum (10) in “Row of Root” above 5. Now, he multiplies
the sum by 5 and puts the product, 10  5, above 52 in “Row of Square” and then
adds the two to get 75 = 52 + 50. The sum he multiplies by 5 and puts the product
(755) above 53 in “Row of Cube.” Again, he adds these to get 500 = 53 + 75  5.
Then he multiplies the sum by 5 and puts 500  5 above 54 in “Row of Square
Square.” Finally, he adds these to get 3125 = 54 + 500  5. (The lines within the
bands mean, in the case of the bottom four bands, that all numbers below them
would be erased on a dust board, and, in the case of the top band, the numbers
above would be erased.)
Now, beginning with the 10 in “Row of Root,” he repeats the above as far as
“Row of Cube” (10 + 5 = 15, etc.), then, with 15, to “Row of Square”
(15 + 5 = 20, etc.) and finally 20 + 5 = 25 is put in “Row of Root.”
Thus the numbers lying entirely in the first column are obtained. Now 3125 (in
“Row of Square Square”) is shifted one place right, 1250 (in “Row of Cube”) two
places, 250 (in “Row of Square”) three places, and finally 25 (in “Row of Root”)
four places to the right, and he puts this “25” at the bottom of the next column
(below the cycle 08995), as in Fig. 22.
At this point he seeks b, the largest single digit so that f(b)  129,908,995 = D,
where

f ðbÞ ¼ bðððb  25b þ 250  102 Þb þ 1250  103 Þb þ 3125  104 Þ

where “25b” means 250+b.


It turns out that f(4) = 146,665,024 is too big, and since
f(3) = 105,695,493 < D), al-Kāshī concludes that 3 is the desired value for b. (This
method of evaluating a polynomial is standard in numerical analysis and is called
Horner’s method in many texts on the subject.)
64 2 Arithmetic in the Islamic World

Start

Ri, Ci, i, Xi

DO J = 1 TO 4

RJ = RJ + XI · RJ-1

N Y
XI · R4 ≤ R5

DO J = 1 TO 4 R5 = R5 – XI · R4

DO K = 1 TO 4
R5–J = R5–J – XI · R4–J

DO J = 1 TO 5 – K

RJ = RJ + XI · RJ–1

XI = XI – 1

I = I –1

N Y
I=0

DO J = 1 TO 5 D = 1+R1+R2+R3+R4

RJ = RJ · 10J PRINT X1, ... , Xr


PRINT R5/D

R5 = R5 + CI STOP

XI = [R5/R4]

Fig. 21
7 Al-Kāshī’s Extraction of a Fifth Root 65

Fig. 22

7.4 Justification for the Procedure

The reader may easily verify with a pocket calculator that 5305 < N, while
5405 > N, and thus al-Kāshī has found the next digit of the fifth root. The question
is: “How?,” and the answer lies in the analog of the identity that underlies
extraction of square roots. If C(n, k) denotes the binomial coefficient “n choose k,”
66 2 Arithmetic in the Islamic World

which counts the number of ways of choosing k objects from a set of n objects, then
the binomial theorem applied to squares may be written

ðA þ BÞ2 A2 ¼ ðC ð2; 2ÞB þ C ð2; 1ÞAÞB

and, applied to higher powers, yields the identities:

ðA þ BÞ3 A3 ¼ ððC ð3; 3ÞB þ C ð3; 2ÞAÞB þ C ð3; 1ÞA2 ÞB;
ðA þ BÞ4 A4 ¼ ðððC ð4; 4ÞB þ C ð4; 3ÞAÞB þ C ð4; 2ÞA2 ÞB
þ C ð4; 1ÞA3 ÞB;
ðA þ BÞ5 A5 ¼ ððððCð5; 5ÞB þ C ð5; 4ÞAÞB þ C ð5; 3ÞA2 ÞB
þ C ð5; 2ÞA3 ÞB þ Cð5; 1ÞA4 ÞB:

The numbers C(n, k) are arranged in a triangular array in Fig. 23. Notice that
each row of this triangle begins and ends with a I and that a number greater than 1
in any row is just the sum of the two numbers to the right and left of it in the row
above it. (Thus in the fourth row the “3” is the sum of 1 and 2 in the row above it.)
If we begin numbering the rows with 0 and use the convention C(0, 0) = 1, then for
all 0  k  n, C(n, k) is the kth entry in Row n, and the rule for generating the
triangle corresponds to the fundamental relationship

Cðn; kÞ ¼ ðn  1; kÞ þ C ðn  1; k  1Þ

This triangle is called “Pascal’s Triangle,” after the French mathematician of the
early seventeenth century, Blaise Pascal, whose Traité du Triangle Arithmétique,
published in 1665, drew the attention of mathematicians to its properties. However,
it might with more justice be called al-Karajī’s triangle, for it was al-Karajī who,
around the year A.D. 1000, drew the attention of mathematicians in the Islamic
world to the remarkable properties of the triangular array of numbers.
If we substitute the values of C(5, k) into the expression for (A + B)5 − A5 we
obtain the equality

ðA þ BÞ5 A5 ¼ ððððB þ 5AÞB þ 10A2 ÞB þ 10A3 ÞB þ 5A4 ÞB:

Fig. 23
7 Al-Kāshī’s Extraction of a Fifth Root 67

In the present case, A = 5  102 and B = b  10, and if we substitute these values
for A and B the right-hand side of the expression now becomes

105 ððððb þ 25  10Þb þ 250  102 Þb þ 1250  103 Þb þ 3125  104 Þb;

and the numbers in boldface are those appearing in the function f given earlier.
To see how al-Kāshī’s technique generates the binomial coefficients, begin with
a page divided into four horizontal bands, and, instead of writing the entries within
a band one above the other, write them in a row, towards the right. Now, fill in the
page as follows:
1. Put the first four powers of 1 up the leftmost column, one in each band.
2. If any column has been filled in, start the next at the bottom by adding 1 to the
entry to the left of it.
3. If any column has been filled in up to a given row, fill in the next row of that
column by adding 1 times the entry in the given row to that in the previous
column of the next row.
4. Each column after the second contains one less row than the column to its left.
These rules will generate Fig. 24 in which the columns are just the diagonals
descending to the right in Pascal’s triangle, apart from the initial “1”s in these diagonals.
Of course, al-Kāshī wants not just the binomial coefficients C(5, k), but the
values 5C(5, 4) = 25, 52C(5, 3) = 250, 53C(5, 2) = 1250, and 54C(5, 1) = 3125.
Thus, we construct a figure, on the model of Fig. 24, but this time:
1. Put ascending powers of 5 up the first column.
2. Add 5 instead of 1 as we move across the bottom row.
3. Whenever we move a number up to add it, first multiply it by 5.
Then we obtain Fig. 25 in which the last entries of the rows are the coefficients
given in boldface in the expansion of f earlier.
One point remains, however. The numbers al-Kāshī must calculate with are not
quite the above but 25  106, 250  107, 1250  108, and 3125  109. To represent
these numbers on the array, al-Kāshī must move the 25 to the right four spaces, the

Fig. 24

Fig. 25
68 2 Arithmetic in the Islamic World

250 three spaces, the 1250 two, and the 3125 one. This is because, where they are,
they are being treated as if they were multiplied by 1010, while, in f(b), 25 is
multiplied only by 106. Since 10 − 6 = 4, al-Kāshī must shift the 25 four places to
the right to make it represent 25  106, etc. This is sufficient explanation of why,
after al-Kāshī has found b according to the procedure outlined and has subtracted
f(b) from D, there remains D′ = N − (A + B)5.

7.5 The Remaining Procedure

Figure 26 shows the next part of the algorithm after 3 has been placed both in “Row
of Result” and next to 25 in “Row of Root” (to form the number 253). The numbers
in the parentheses on the right show how the algorithm calculates f(b) in stages.
Thus, 253 is multiplied by 3 to obtain 759, which is then put directly above the
25000 (the last two zeros not being shown) and added to it to obtain 25,759. This is
then multiplied by 3, written above the 1,250,000, and added to obtain 1,327,277.
Finally, this is multiplied by 3 and the product added to the 31,250,000 in the row
above it. This sum is finally multiplied by three and the product, which is f(3),
is subtracted from 129,908,995 in the “Row of the number.” The difference, D′, is
D − f(3).
Next, al-Kāshī begins the process “once up to the row of the square square”
(with 253 + 3 = 256, etc.), then to the row of the cube (with 256 + 3 = 259), then
to the row of the square (259 + 3 = 262). Finally, in the row of the number he puts
265 = 262 + 3. The multiplications of course are all by 3 instead of by 5. The top
numbers in the bands are then shifted so the numbers obtained will represent the
constants in the polynomial:

gðcÞ ¼ ðððc  265c þ 28; 090  102 Þc þ 1; 488; 770  103 Þc þ 39; 452; 405  104 Þc;

The next digit, c, must satisfy a condition entirely analogous to the one b satisfied,
i.e., it must be the largest single digit so that g(c) does not exceed 24,213,502  105.
Al-Kāshī finds c = 6.
The bracketed lines in Fig. 27 denote the computation of the terms of g(6), and
D″ the final difference. Finally, al-Kāshī performs the procedure of going up to the
“Row of Square Square,” etc. The reader should now be able to follow without
difficulty the steps as shown in Fig. 27.

7.6 The Fractional Part of the Root

At this point al-Kāshī has finished the calculation of the integer part of the fifth root
of the given 14-place number. He had perfect control of decimal fractions and there
is no doubt he knew that he could now shift again and continues the procedure to
7 Al-Kāshī’s Extraction of a Fifth Root 69

Fig. 26

extract successive decimal places of the fifth root. Also, al-Khwārizmī, in a part of
his treatise on arithmetic reported by the Latin writer John of Seville (fl. ca. 1140),
pffiffiffi
gives an example of calculating 2 by calculating:
70 2 Arithmetic in the Islamic World

Fig. 27

pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffi 2; 000; 000 : 1414
2¼ ¼
1000 1000

a procedure which al-Kāshī also recommends. So, even without decimal fractions,
one can obtain any desired degree of accuracy.
What al-Kāshī does here, however, is to add up the top numbers in each of the
four bottom rows and increase the sum by one, i.e., he forms
7 Al-Kāshī’s Extraction of a Fifth Root 71

412694958080
1539906560
2872960
2680
þ 1
414237740281

and states that the fifth root of the given number is 536 + 21/414,237,740,281.
Al-Kāshī’s rule for finding the fractional part is based on the approximation
 1=k r
nk þ r ¼ nþ
ðn þ 1Þk nk

where he explicitly calculates

ðn þ r Þk nk ¼ C ðn; 1Þnk1 þ Cðn; 2Þnk2 þ . . . þ 1:

This is, of course, just the method of linear interpolation we discussed earlier, for
nk + r is r units of the total (n + l)k − nk units between two successive kth powers, so
linear interpolation would place its kth root, (nk + r)1/k, the fraction r/[(n +l)k − nk] of
the way between the two kth roots n and n + l.
Figure 28 reproduces a page from the printed version of al-Kāshī’s The
Calculators’ Key, which shows the entire calculation we have just explained. The
reader will benefit from identifying the numerals and following the procedure
through the first “Once up to the row of square square.”

8 The Islamic Dimension: Problems of Inheritance

Al-Khwārizmī devotes the first half of his book on algebra to solutions of the
various types of equations and demonstrations of the validity of his methods, but
the latter half contains examples of how the sciences of arithmetic and algebra could
be applied to the problems posed by the requirements of the Muslim laws of
inheritance.
When a person dies who leaves no legacy to a stranger the calculation of the
legal shares of the natural heirs could be solved by the arithmetic of fractions. The
calculation of these shares was known as ‘ilm al-farā’iḍ, and two examples from
al-Khwārizmī’s work illustrate the applications of arithmetic here.
72 2 Arithmetic in the Islamic World

Fig. 28
8 The Islamic Dimension: Problems of Inheritance 73

8.1 The First Problem of Inheritance

This problem is a simple one, namely,


Example 1 “A woman dies, leaving her husband, a son and three daughters,” and
the object is to calculate the fraction of her estate that each heir will receive.
The law in this case is that the husband receives 14 of the estate and that a son
receives twice as much as a daughter. (It should be said, however, that from the
woman’s point of view Islamic inheritance law was a considerable improvement
over what the pre-Islamic requirements in the Arabian Peninsula had been.)
Al-Khwārizmī then divides the remainder of the estate after the husband’s share
has been deducted, namely 34, into five parts, two for the son and three for the
daughters. Since the least common multiple of five and four is twenty, the estate
should be divided into twenty equal parts. Of these, the husband gets five, the son
six, and each daughter three.

8.2 The Second Problem of Inheritance

This problem is a little more complicated and illustrates how unit fractions were
employed to describe more complex fractions.
Example 2 A woman dies, leaving her husband, son and three daughters, but she
also bequeaths to a stranger 18 þ 17 of her estate. Calculate the shares of each.
One law on legacies is that a legacy cannot exceed 13 of the estate unless the
natural heirs agree to it. (Here complications could enter because of the provision
that if some agree and some do not agree those who do agree must pay, pro-rated,
their share of the excess of the legacy over the third.) In the present case, however,
since 18 þ 17  13 no complications enter, and the second provision on legacies,
namely that a legacy must be paid before the other shares are calculated, now takes
effect.
As in the above problem, the common denominator of the legal shares of her
relatives
1 1 is1520.  Also, the fraction of the estate remaining after the stranger’s legacy
8 þ 7 ¼ 56 has been 
41
paid is 56 Then the ratio of the stranger’s share to the total
 41
shares of the family is 1556 : 56 ¼ 15 : 41. Thus, of the whole estate, the stranger
will receive 15 parts to the 41 parts the natural heirs will receive. Multiplying both
numbers by 20 to facilitate the computation of the shares of the heirs, we find that of
a total of 20 (15 + 41) = 20  56 = 1120 parts the stranger receives 20  15 = 300
and the heirs jointly receive 20  41 = 820. Of these parts, the husband receives 14,
6
namely 205, the son 20 namely 246, and each of the daughters gets 123.
74 2 Arithmetic in the Islamic World

8.3 On the Calculation of Zakāt

Another example of the use of arithmetic in the Islamic faith is in the calculation of
zakāt, the community’s share of private wealth. This is payable each year, at a
certain rate, and the following problem, taken from The Supplement of Arithmetic of
the eleventh-century mathematician, Abū Manṣūr al-Baghdādī, follows the gradual
diminution of a sum of money as the zakāt is paid for 3 years. Its treatment of the
fractional parts of a dirhām reminds one of Kūshyār’s treatment of fractions, and in
presenting it we paraphrase slightly, following the translation in Saidan (1987) “We
want to pay the zakāt on 7586 dirhāms, the amount that Muḥammad ibn Mūsā
al-Khwārizmī mentioned in his work.” (The dirhām was divided into sixty fulūs,
the plural of fils (see Plate 2).)
The rate of zakāt is 1 dirhām in 40, but al-Baghdādī does not divide 7586 by 40
according to the algorithm Kūshyār describes. Rather he calculates the total due on
7586 dirhāms, place-by-place, as follows:
From the first place we remove 1, which we make 40, and then remove 6 from
the 40. This 6 is the zakāt due on 6 dirhāms and it is 6 parts of (the 40 into which
we have divided) a dirhām. Thus, of the 40 there remains 34 parts. This we put
under the five that has remained in the units place, as in Fig. 29a.
1
We must now calculate 40 of the 80 that arises from the 10’s place, to obtain 2,
which we subtract from the five in the unit’s place. This leaves what is shown in
Fig. 29b.
In the 100’s place there is 500, on which the zakāt due is 12 12. Of the 40 parts
into which we have divided the dirhām, 12 is 20, so when we subtract this from 34
there remain 14 parts. Also 12 from 83 leaves 71, so there now remain the figures
shown in Fig. 29c.
1
Finally, 40 of the 7000 we obtain from the 1000’s place is 175, and when we
subtract this from 571 there remains 396, so the answer is that shown in Fig. 29d.
Al-Baghdādī follows this for two more years, after which there remain the
number of dirhāms shown in Fig. 29e, where, e.g. the 14 means 14/(40)3 dirhāms.
(The tax collector is going to get every last fils due!)
Of course, dirhāms are divided into 60 fulūs, not 40, and so, to calculate the
zakāt, the base-40 fractions, which were convenient to use in the intial stage, must
now be converted into sexagesimal fractions, and here al-Baghdādī points out a slip
on the part of al-Khwārizmī, his source for the problem. Evidently, al-Khwārizmī
said that if each of the fractional parts (i.e., 6, 8 and 14) is increased by 12 then they
become sexagesimal parts, i.e., minutes, seconds and thirds. This is of course true
for the 6, because
3
6 9
¼6 2 3¼
40 40  2 60

but it is false for the following parts, and al-Baghdādī gives the correct rule.
8 The Islamic Dimension: Problems of Inheritance 75

Plate 2 Obverse and reverse of two coins from the medieval Islamic world. The one on the right
is a fils of Damascus, minted in 87 A.H. (Anno Hijra). The obverse of the coin on the right names
the Caliph al-Walīd (of the Umayyad Dynasty) and gives the shahāda (Muslim confession of faith:
“There is no god but Allah and Muḥammad is the Messenger of Allāh.”) The one on the left is a
dirham of Medīnat al-Salām (Baghdad) issued in 334 A.H. and names, on the two faces of the
coin, the Būyid rulers Mu‘izz al-Dawla and ‘Imād al-Dawla as well as the Caliph al-Muṭī‛. (Photo
courtesy of the American Numismatic Society, New York.)

Fig. 29
76 2 Arithmetic in the Islamic World

Without question, ‛ilm al-farā’iḍ is an important subject for Muslims, but in


estimating the place of mathematics in that discipline a cautionary note, written by
the great fourteenth century Maghribi historian, Ibn Khaldūn, is worth recording.
Religious scholars in the Muslim cities have paid much attention to it. Some authors are
inclined to exaggerate the mathematical side of the discipline and to pose problems
requiring for their solution various branches of arithmetic, such as algebra, the use of roots,
and similar things. It is of no practical use in inheritance matters because it deals with
unusual and rare cases. (Transl. in Rosenthal, cited in bibliography of Chap. 1.)

Exercises

1. Use Kūshyār’s method to add and subtract 12,431 and 987, showing your steps
as in the text.
2. Develop an algorithm for halving a number that starts with the highest place in
a number. Why do you think the Muslim calculators worked from the lowest
place?
3. Use an operation modeled on raising to obtain a decimal expansion of 243 7 .
4. Use Kūshyār’s method to multiply 46 by 243.
5. Use Kūshyār’s method of division to divide 243 by 7, and then use the method
of raising to find a 3-place sexagesimal approximation to 57.
6. Adapt the method of raising to find a 3-place decimal approximation to 57.
7. Devise a method for converting decimal integers to sexagesimal integers. Do
the same for fractions. Now do the latter, but going from sexagesimal to
decimal.
8. List some possible values for ke mb h, including some fractional ones.
9. Add, subtract and multiply the two sexagesimal numbers 36, 24 and 15, 45.
Divide 2, 6, 15, 0 by 8, 20.
10. Use the lattice method to multiply 2468 by 9753.
11. Use the procedure in Sect. 3 to express 19/35 as a related fraction.
12. With A and N as in the section on square roots show that ((A + 1)  100)2 > N,
while (A  100)2 < N. Conclude that A is the first digit of the root.
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
13. Use al-Kāshī’s method, including linear interpolation, to find 20; 000.
14. If a man dies, leaving no children, then his mother receives 16 and his widow 14 of
the estate. If he has any brothers or sisters, a brother’s share is twice that of a
sister. Find the fractions of the estate due if a man dies, leaving no children but
a wife, a mother, a brother, two sisters, and a legacy of 19 of the estate to a
stranger.
15. Give a rule for converting the remaining base-40 parts in the example from
al-Baghdādī to sexagesimal parts. Generalize this rule to one for converting
fractions from base n to base m.
Exercises 77

16. Show that for any single digit b

f ðbÞ ¼ ð5  102 þ b  10Þ5  ð5  102 Þ5

and conclude that b is the desired second digit of the fifth root, where f is the
function in our discussion of al-Kāshī’s extraction of the fifth root.
17. Al-Kāshī’s method of evaluating f(b) suggests evaluating an arbitrary
polynomial

gðxÞ ¼ an xn þ an1 xn1 þ    þ a1 x þ a0

as

gðxÞ ¼ ð. . .ðan x þ an1 Þx þ an2 Þx þ    þ a1 Þx þ a0 ;

where the initial dots denote an appropriate number of parentheses and those in
the middle denote intermediate terms.
(a) Evaluate g(2), where g(x) = 5x3 − 3x2 + 7x + 6 by this method.
(b) If addition and multiplication are each counted as one operation how many
operations are necessary to evaluate g(x) by this formula? How many are
necessary according to the usual method?
18. Show that the sum 412,694,958,080 +  + 1 calculated in al-Kāshī’s extrac-
tion of the fifth root is equal to 5375 − 5365.
19. Use al-Baghdādī’s method and format (as in Fig. 29) to supply the details of the
computation of the zakāt for year two.

Bibliography

Al-Kāshī. 1977. Miftah al-Hisab. (edition, notes and translation by Nabulsi Nader). Damascus:
University of Damascus Press. (This is the work of al-Kashī whose title we translate as The Key
of Arithmetic. Our Fig. 28 is taken from this book, courtesy of the publishers).
Al-Uqlīdisī, Abū al-Ḥasan. The Arithmetic of al-Uqlīdisī (transl. and comm. A. S. Saidan).
Dordrecht/Boston: Reidel, 1978.
Crossley, J.N., and A.S. Henry. 1990. “Thus Spake al-Khwārizmī: A translation of the Text of
Cambridge University Library Ms. Ii.vi.5.” Historia Mathematica 17: 103–131. (A translation
of a medieval Latin translation of his book on Hindu reckoning.).
Dakhel, Abdul-Kader. 1960. Al-Kāshī on Root Extraction. Vol. 2 Sources and Studies in the
History of the Exact Sciences, ed. W.A. Hijab and E.S. Kennedy. Beirut: American University
of Beirut.
Djebbar, A. 1992. Le traitement des fractions dans la tradition mathématique arabe du Maghreb. In
Histoire des fractions, fractions d’histoire, ed. P. Benoit et al., 223–245. Bâle-Boston-Berlin:
Birkhäuser Verlag.
Ibn Labbān, Kūshyār. 1965. Principles of Hindu Reckoning (transl. and comm. M. Levey and
M. Petruck). Madison and Milwaukee, WI: University of Wisconsin Press.
78 2 Arithmetic in the Islamic World

King, D.A. 1974. “On medieval multiplication tables.” Historia Mathematica 1: 317–323.
King, D.A. 1979. “Supplementary notes on medieval Islamic multiplication tables.” Historia
Mathematica 6: 405–417.
Rashed, R. 1994. “Numerical analysis: The extraction of the nth root and the invention of decimal
fractions (eleventh to twelfth centuries).” In The Development of Arabic Mathematics: Between
arithmetic and algebra, ed. R. Rashed (translated by A.F.W. Armstrong). Dordrecht; Boston:
Kluwer Academic.
Saidan, A.S. 1974. “Arithmetic of Abū 1-Wafā’’. Isis 65: 367–375. (This summarizes an important
work on finger arithmetic in medieval Islam).
Saidan, A. S. “The Takmila fi’l-Ḥisāb by al-Baghdādī”. In: From Deferent to Equant: A Volume of
Studies of the History of Science in the Ancient and Medieval Near East in Honor
of E. S. Kennedy. (D. A. King and G. A. Saliba, eds.). New York: New York Academy of
Sciences. 1987, pp. 437 – 443.
Souissi, M. (ed. and French translation of Ibn al-Bannā’, Summary of of Arithmetic Operations)
Al-talkhīs a‘mal al-ḥisāb.Tunis, 1969.
Van der Waerden, B. L. and M. Folkerts. Written Numerals. Walton Hall, UK: The Open
University Press, 1976.
Woepcke, F. 1986. Traduction du traite d’arithmétique d’Aboul Haçan Ali Ben Mohammed
Alkalçādī. Reprinted in Woepcke, F. Études sur les mathématiques arabo-islamiques, Vol II,
ed. F. Sezgin. Hirshberg: Strauss Offsetdruck, GmBH.
Chapter 3
Geometry in the Islamic World

Since Medieval Islam was an heir to Greek mathematics, where geometry devel-
oped to such a great extent, it cannot be surprising that that area of study found avid
students in the lands of medieval Islam. Part of the motivation for such study,
naturally, came from the need to measure and build that all advanced civilizations
have. And part came from the desire to master the intricate geometrical models of
celestial motions found in Ptolemy’s Almagest. Another factor was, surely, the
desire to understand the basis for such apparatus as parabolic mirrors, astrolabes,
planetary equatoria, and sundials. But another important consideration was the
appeal to many thinkers of the kind of certainty that geometrical proofs provided.
One eminent tenth-century geometer, Abū Sahl al-Kūhī, even doubted that
Archimedes’ famous approximation to the ratio of the circumference of a circle to
its diameter was really by Archimedes, since the method used “is an approximation
and his purpose is always to discover knowledge of things exactly, not
approximately…”.1
We cannot treat all these different motivations in detail, but we shall focus on
three: the interest in geometrical constructions, the use of geometry in the mea-
surement of lengths, areas, and volumes and the discovery of geometrical theorems.

1 Geometrical Constructions

That geometrical constructions which were of keen interest to the ancient Greek
geometers is evident from the fact that Euclid devoted two of the thirteen books of
his Elements to an account of some of the constructions that had been done up to his
time. In Book IV, Euclid explains how to construct a regular pentagon, hexagon,
and 15-gon. And in Book XIII, he tells how to construct the regular polyhedra,

1
For details see Berggren 1983, which also supplements this chapter with an episode in the history
of centers of gravity in medieval Islam.

© Springer Science+Business Media New York 2016 79


J.L. Berggren, Episodes in the Mathematics of Medieval Islam,
DOI 10.1007/978-1-4939-3780-6_3
80 3 Geometry in the Islamic World

namely the tetrahedron, cube, octahedron, dodecahedron, and icosahedron—which


have, respectively, 4, 6, 8, 12, and 20 faces.
Much of the Euclidean geometry studied in our schools deals with the geometry of
triangles and circles, where the instruments used for construction are the straightedge
and the compass. The former is used to join two points with a straight line segment or to
extend a straight line segment in either direction, and the latter is used to draw a circle
with an arbitrary point as center and passing through any other given point. The ability
to join points or to extend straight line segments is assumed in Postulates 1 and 2 of The
Elements, and the ability to draw arbitrary circles is postulated in Postulate 3. Perhaps
for these reasons the straightedge and compass are called “the Euclidean tools”.
The student should be careful to distinguish between a straightedge and a ruler, for,
unlike the ruler, the straightedge is not assumed to have parallel edges or any marks
along an edge. Similarly, the compass Euclid assumes for drawing circles is not the rigid
compass we are used to, which stays set at whatever distance we open it to and may be
used for transferring lengths. It is, rather, a compass that, once set on the paper, will draw
around a given point as center, the circle that passes through any other point, but it will
not transfer lengths. For this reason it has been called a collapsible compass, since its
legs fall back together when they are removed from the drawing plane.
The first thing Euclid does in The Elements is to show that his collapsible
compasses are also able to transfer lengths, and to introduce this chapter on geo-
metrical constructions we shall give the essentials of Euclid’s proof. It is the first
systematic discussion of a theory of constructions, in that it shows how construc-
tions with one kind of tool can be done with a (seemingly) weaker tool.
In Proposition 1 of Book I, Euclid solved the following problem: On a given
segment AB construct an equilateral triangle ABG (Fig. 1). By the properties of his
compass Euclid is allowed to construct the circles centered at A and B having radii
AB. If they intersect at G then AG = AB = BG, so that Δ(ABG) is equilateral.
In Prop. 2, Euclid shows how to place at a given point D a straight line segment
DW equal to a given segment AB (Fig. 2). He says to draw the straight line BD
and, by Prop. 1, to construct the equilateral triangle BDG. Draw the circle with
center B and radius BA, and let this circle cut GB, extended, at E. Finally, let the
circle with center G and radius GE cut the side GD extended at W. Then
DW = GW – GD = GE – GB = BE = AB, which was required.
Having proved these two propositions, Euclid finishes the demonstration with
Prop. 3, where he shows how to cut off at the point Z on a given segment DE, a
segment equal to a given segment AB (Fig. 3). By Prop. 2, he can construct a
segment ZF equal to AB, and the circle with center Z and radius ZF will intersect
the line segment DE (extended if necessary) at a point G. The radius ZG is equal to
AB. With this proposition Euclid is able to transfer a given length from any point to
any other point, and therefore has shown that the collapsible compass is able to do
the same operations as ours.
The reader has doubtless observed that this demonstration serves no practical need.
The Greeks knew as well as we how to lock two compass arms in a fixed position, so
why did Euclid bother with some less powerful tool? The answer lies in the fact that one
of Euclid’s concerns was with achieving the greatest possible economy of basic
1 Geometrical Constructions 81

Fig. 1

Fig. 2

assumptions. This concern is a matter of taste rather than of logical necessity, but it has
been characteristic of mathematicians since the time of the Greek geometers. We shall
see that it was also a concern that the geometers of Islam shared.
Other basic constructions in The Elements are those of the perpendicular to a
given line and passing through a given point, the division of a line segment into an
arbitrary number of equal segments, the bisector of an angle and the tangent to a
82 3 Geometry in the Islamic World

Fig. 3

given circle passing through a given point outside the circle, and these are probably
familiar to the reader.

2 Greek Sources for Islamic Geometry

Writers of the Islamic world were familiar with The Elements since the late eighth
century through the translations that were done in Baghdad on commission for the
caliphs Hārūn al-Rashīd and al-Ma’mūn. The numerous Arabic editions and
commentaries which have survived since that time testify to the immense influence
of Euclid’s Elements on Islamic mathematics, since it was one of the basic texts that
any student of mathematics and astronomy would have to read.
The name of Euclid, however, had to share honors in Islamic mathematics with
that of another mathematician, namely Archimedes, whose treatise On the Sphere
and Cylinder excited great admiration among the Muslim mathematicians and
inspired some of their best works. In the preface to his book Archimedes mentions
his discovery of the area of the segment of a parabola and, since his treatise on the
subject was not known in the Muslim world, this reference stimulated Thābit b.
Qurra and his grandson Ibrāhīm b. Sinān to a successful search for proofs of
Archimedes’ result. In addition, the problem posed in the second part of
Archimedes’ work, that of dividing a sphere into two segments by a plane so that
the two segments have to one another a given ratio, acted as a considerable stimulus
to investigations in algebra and the conic sections. Another treatise, attributed to
Archimedes by Arabic sources, but unknown in Greek, is titled The Heptagon in the
Circle. It was put into Arabic by Thābit b. Qurra, who translated or revised
translations of all Archimedean works extant in medieval Arabic. This work takes
up the problem of constructing a regular polygon not discussed by Euclid, namely,
the seven-sided polygon, the first unsolved case after Euclid’s construction of
polygons with 3, 4, 5, and 6 sides. The extensive literature on Archimedes’ works is
indicative of the fact that they formed the second pillar of Islamic geometry.
2 Greek Sources for Islamic Geometry 83

The third Grecian column supporting geometrical research in Islam was


Apollonios of Perga’s The Conics, a work in eight books (chapters) written around
200 B.C. Apollonios’s work is difficult, and in the last few books he treats advanced
topics such as the minimum distance from a point to a conic, so it is no surprise that
only the first four of the original eight books survive in Greek. Evidently, the
remaining books were either too specialized or too difficult for the scholars of late
antiquity. It is a testimony to the ability of the Muslim geometers that seven books
survive in Arabic, and the bibliographer al-Nadīm tells us that in the tenth century
parts of the eighth book were also extant. The work not only formed a base for
much advanced research in geometry and optics—and even algebra, as we shall see
in the work of ‛Umar Khayyām—but also it inspired one of the most able of
Muslim scientists, Ibn al-Haytham, to attempt a restoration of the eighth book.
Although a student meets the basic Euclidean constructions during the high
school years, the elementary properties of the conic sections are often not
encountered until a first university course in the calculus, and then by an approach
that is quite different from that used by the ancient and medieval authors. For these
reasons, the following account of some of the basic ideas in Apollonios’s book will
provide necessary background for understanding the material in the following
sections.

3 Apollonios’s Theory of the Conics

A surface of a double cone is formed by the straight lines that pass through the
points on the circumference of a circle, called a base, and a fixed point not in
the plane of the base (Fig. 4). Any one of the straight lines is called a generator of
the surface, the fixed point is its vertex and the straight line through the vertex
and the center of the base is called the axis. A cone is the solid figure bounded by
the part of the surface of a double cone between the vertex and a base.
Euclid and Archimedes both wrote about conic sections before Apollonios, but
in their treatments of conic sections the cone was the so-called right cone, in which
the axis is perpendicular to the base circle. The right cone was then cut by a plane
perpendicular to a generator, and in this way one obtained a plane section, the kind
of section depending on the angle at the vertex of the cone. Thus in the ancient
world the conic sections were plane figures, whereas we look at the boundaries of
these plane figures and think of the conic sections as curves.
Apollonios generalized this method of generating the conic sections by con-
sidering plane sections of an arbitrary double cone, whose axis may be skew to the
base, and he showed that, apart from the circle, only the three known conic sections
could arise.
At the beginning of his Conics Apollonios used the fact that these figures are
sections of cones only to establish their basic properties, called “symptoms”, and
for the rest of the eight books he proved everything from these symptoms. Since, in
84 3 Geometry in the Islamic World

Fig. 4

the following, we shall be concerned only with parabola and hyperbola we shall
limit our discussion of the symptoms to these two sections.
According to Apollonios a parabola is the common section of a cone and a plane
when the plane is parallel to a generator of the cone. And the hyperbola is either one
of the two common sections (each called a branch) formed when the plane meets
both parts of the double cone. In any conic section a line joining two points on the
boundary is called a chord. Apollonios showed that the midpoints of all chords
parallel to a fixed chord lie on a straight line, and, if this straight line intersects the
boundary in A, then the tangent at A is parallel to all the chords. The straight line is
called a diameter of the section and an intersection of a diameter with the boundary
is called a vertex of the section. The half-chords lying on one side of the diameter
are called the ordinates to the diameter. When the ordinates are perpendicular to the
diameter, such a diameter is unique and is called the axis. We have illustrated these
ideas for a parabola in Fig. 5, where FE is a chord, YZ and UV are chords parallel
to FE, and AB is the diameter passing through the midpoints of these chords. XY is
a typical ordinate to the diameter AB, and the line CD is the axis.
3 Apollonios’s Theory of the Conics 85

Fig. 5

3.1 Symptom of the Parabola

In the case of the parabola, the diameters are all parallel to the axis CD. Let AB be a
given diameter, X be any point on AB, and XY the ordinate at X. Apollonios
showed (Fig. 5) that corresponding to the diameter AB there is a fixed line segment
p so that the rectangle that is equal to the square on XY and has one of its sides
equal to AX will have its other side exactly fit the segment p. The segment p is
called the parameter (or latus rectum) belonging to diameter AB. If we set
AX = x and XY = y, then the Apollonian symptom becomes the modern equation
p  x = y2. The Greek term Apollonios used to describe this is paraballetai and so
he called this section parabolē, from which comes our word “parabola”. (The word
paraballetai literally means “it is put alongside”, referring to the rectangle exactly
fitting along the parameter.)

3.2 Symptom of the Hyperbola

Here the curve has a center, the point on the axis midway between the vertices of the
two sections (Fig. 6). Any line through this center is a diameter and the center bisects
the part of a diameter between the two branches. Let C, C′ be the endpoints of the
part of some diameter between the two branches of the curve and let a = CC′, called
the transverse side (or latus transversum). Apollonios proved that corresponding to
a there is a segment p with the following property: A rectangle that is equal to the
square on an ordinate XY and has one side equal to CX will have its other side
exceed p. Moreover, the rectangle (shaded in Fig. 6) contained by the excess of this
86 3 Geometry in the Islamic World

Fig. 6

side over p and by CX is similar to the rectangle whose sides are a and p. Thus the
other side s satisfies the proportion s:CX = p:a, i.e., s = (p/a)  CX.
To see what the symptom for the hyperbola says geometrically let C′ be the other
end of the diameter and let CP be the parameter. Also let the perpendicular to CX at
X meet the line C′P at E. Then Apollonios’s symptom implies that the rectangle
whose sides are CX and XE is equal to (XY)2. (The side p is called the parameter.)
Since the Greek word for “it exceeds” is “hyperballetai”, Apollonios’s name for this
case is hyperbolē, whence comes our word “hyperbola”. Again, if we let
CX = x and XY = y the symptom becomes

y2 ¼ ðp þ sÞx ¼ px þ ðp=aÞx2 ;

which is a modern equation for the hyperbola. (In the case when p = a, the shaded
rectangle in Fig. 6 is a square, and this is the case Abū Sahl al-Kūhī uses in the next
section.)
The above properties of the sections were hardly Apollonios’s discoveries, since
they were known to Archimedes, but one of Apollonios’s contributions was, having
shown that the symptoms characterize the conic sections, to use the properties
stated in the symptoms to name the sections parabola, hyperbola, and ellipse
(“ellipse” from elleipsis, meaning “falling short”).
4 Abū Sahl al-Kūhī on the Regular Heptagon 87

4 Abū Sahl al-Kūhī on the Regular Heptagon

4.1 Archimedes’ Construction of the Regular Heptagon

It was during the first half of the fourth century B.C. when the Greek, Menaechmos,
invented the conic sections and used them to solve the problem of constructing a
cube whose volume is twice that of a given cube. If the side of the given cube is
pffiffiffi
a and the side of the required cube is b then b3 = 2a3, so b ¼ 3 2a2 , and the
pffiffiffi
problem is one of the scaling up of a line segment of length a to one of length 3 2a.
(Because of the work on the theory of equations by the nineteenth century French
mathematician, E. Galois we know that this cannot be done with straightedge and
compass. However, both the ancient Greek and the Muslim geometers realized that
this construction and many others are possible with conic sections.) And, in both
the Greek and the Islamic worlds the principal uses of the conic sections (other than
the circle) were in geometrical constructions, the design of mirrors that focus light
to cause burning, and the theory of sundials. (It was J. Kepler who, early in the
seventeenth century, first used ellipses in astronomy to model the planetary orbits.)
Prior to the appearance of the Conics Euclid had written a (lost) treatise on the
conic sections and Archimedes had written treatises on finding the area of a seg-
ment of a parabola and the volumes of several solids formed by the rotation of conic
sections about an axis. He had also used the hyperbola in a geometric construction,
but he missed the chance to use it in the construction of the regular seven-sided
polygon, called a heptagon, in a treatise that is attributed to him.
Archimedes begins the treatise with the square ABDG and its diagonal BG
(Fig. 7). He now turns a straightedge around D so that it crosses the diagonal BG,
the side AG and the side BA extended at points T, E, and Z, respectively, and so
that Δ(AEZ) has the same area as Δ(DTG). Lastly, he draws KTL parallel to AG.
He then proves that K and A divide the segment BZ so that the three segments BK,
KA, and AZ can form a triangle and so that BA  BK = ZA2 and KZ  KA = KB2.
Then, form Δ(KHA) so that KH = KB and AH = AZ, and through B, H, and Z

Z
B K A

T
D L G

Fig. 7
88 3 Geometry in the Islamic World

_
draw the circle BHZ. Archimedes proves that BH is one-seventh of the circum-
ference of the circle.
No other construction like this one is known in Greek or Islamic mathematics,
and this uniqueness is a hallmark of Archimedes’ work. However, the construction
of line DZ was not one that could be done with standard tools of Euclidean
geometry, the straightedge, and compass.2 In fact, it was so puzzling that the
tenth-century Muslim mathematician, Abū al-Jūd remarked with some justice that
“perhaps its execution is more difficult and its proof more remote than that for
which it serves as a premise.” No one could deny the elegance of Archimedes’
solution, but the auxiliary construction of line DZ did, as Abū al-Jūd suggested,
raise a serious problem.
Of course, if one thinks of a straightedge rotating around D so its passes between
A and G then, as it moves toward A, Δ(AEZ) can become arbitrarily small while
Δ(DTG) approaches one-quarter of the square. On the other hand, as the straight-
edge approaches G, Δ(AEZ) becomes arbitrarily large and Δ(DTG) becomes
arbitrarily small. Thus, for some intermediate position, the two triangles will be
equal, and so Archimedes’ procedure is, if we like, an existence proof, but it is
hardly a construction. Thus the problem remained as the one not constructively
solved for almost 1200 years.

4.2 Abū Sahl’s Analysis

In the latter half of the tenth century, however, in Baghdad and the surrounding
area, there gathered a group of remarkable scientists from all over the eastern part of
the Islamic world under the patronage of a series of kings belonging to the Būyid
family. Foremost of these kings was ‘Aḍud al-Daula (“Arm of the State”), and one
of the chief scientists at his court was Abū Sahl al-Kūhī, who came from the
mountainous region (“kūh” is the Persian word for “mountain”) south of the
Caspian Sea. According to the biographer, Al-Bayhaqī, who lived about a century
later, Abū Sahl was originally a juggler of glass bottles in the market place of
Baghdad, but then he gave up juggling for study and research in the sciences.
Perhaps it was his experience as a juggler that aroused his interest in centers of
gravity, for his correspondence contains some of the deepest theorems on centers of
gravity since the time of Archimedes. In fact, Abū Sahl knew Archimedes’ work
well, and he wrote a commentary on Book II of On the Sphere and Cylinder, in
which he explained how to solve by conic sections the problem of constructing a
sphere with a segment similar to a segment of one sphere and having surface area
equal to a segment of a second sphere. In addition, he wrote a work on “the
complete compass”, an instrument for drawing conic sections. Consistently with

2
The construction is of a type known as “verging constructions,” on which see the discussion of
the construction of a regular nonagon in the following section.
4 Abū Sahl al-Kūhī on the Regular Heptagon 89

this interest and experience in conic sections, Abū Sahl looked at the problem of
constructing a regular heptagon, and he saw that a solution lay in conic sections.
His method of attack was inspired by Archimedes’ proof, and when he refers to the
construction of the heptagon as a problem that no geometer before him, “not even
Archimedes”, had been able to solve, he was no doubt referring to the problem of
actually doing the construction that Archimedes’ method calls for.
Abū Sahl’s method is to analyze the problem first, that is to suppose that the
heptagon has been constructed and to reason backward, to the givens, by a series of
inferences that can be validly reversed. Analysis is an ancient method, which
Proklos, a biased source, attributes to Plato. According to this method the mathe-
matician assumes what is to be proved and then reasons from this until he reaches
the given. If the chain of reasoning can be reversed then he has found the synthesis,
or proof, of what is required starting from the given. Abū Sahl used analysis to find
a series of constructions equivalent to that of the regular heptagon, until he arrived
at given. Many geometers of the late tenth century felt that a complete solution of a
problem required both the analysis and the synthesis, and Ibrāhīm b. Sinān, whom
we shall speak of later in this chapter, wrote a treatise on these two methods.
We shall, however, present only the analysis as it is found in a treatise written by
Abū Sahl and dedicated to King ‘Aḍud al-Daula, that is we shall trace the series of
constructions by which Abū Sahl reduced the problem of constructing the regular
heptagon to one of constructing two conic sections. When he has done this he has
shown how a peculiar construction, a seemingly ad hoc device, could be fitted into
the theory of conic sections. Such a unification of disparate mathematical methods
is the very stuff of which mathematical progress is made.

4.2.1 First Step of Reduction: From Heptagon to Triangle

Suppose that we have succeeded in constructing the side BG of a regular heptagon


_ _ _ _
(Fig. 8) in circle ABGD and that AB ¼ 2 BG . Then arc ABG ¼ 3 BG and, since
_ _ _
BG is one-seventh of the whole circumference, A DG ¼ 4 BG. According to VI, 33
of Euclid’s Elements angles of Δ(ABG) on the circumference are in the same
proportion as the arcs they subtend, and therefore ^B ¼ 4^A while ^G ¼ 2^A.
Thus, the construction is reduced to the problem of constructing a triangle whose
angles are in the ratio 4:2:1

4.2.2 Second Step of Reduction: From Triangle to Division of Line


Segment

Let ABG be a triangle so that ^B ¼ 2^G ¼ 4^A (Fig. 9) and prolong BG in both
directions to D and E so that DG = GA and EB = BA. Complete Δ(AED). (In the
following ^A; ^B;^G will denote the angles of Δ(ABG) at the corresponding
vertices, and other angles at these vertices will be referred to unambiguously.)
90 3 Geometry in the Islamic World

Fig. 8

The idea of the proof is to show that ^A ¼ ^D so that the two triangles ABG and
DBA are similar, then to show that ^BAE ¼ ^G so that the two triangles AEB and
GEA are similar. When this is done then, by the first similarity, DB/BA = AB/BG
and, by the second, GE/AE = AE/BE. Thus, it follows that

BA2 ¼ DB  BG and EA2 ¼ GE  EB:

However, since AB = BE, ^E ¼ ^BAE ¼ ^G so that EA = AG = GD. Thus,


the second of the preceding equalities becomes GD2 = GE … EB and the
first becomes BE2 = DB  BG since BA = BE. Thus, once we have shown that
^A ¼ ^D and ^BAE ¼ ^G will have shown that the construction of a regular
heptagon implies the division of a straight line ED at two points B, G so that

(1) GE  EB ¼ GD2 and (2) DB  BG ¼ BE2 :

As for the angles, then, notice that ^BGA is an exterior angle of the isosceles
triangle AGD, where AG = GD, so that ^BGA ¼ ^DAG þ ^D ¼ 2^D. But we
are given that ^BGA ¼ 2^A, so that ^A ¼ ^D. In the case of the other angle,
observe that ^B is an exterior angle of the isosceles triangle ABE so that ^B ¼
2^BAE while, at the same time, ^B ¼ 2^G, so that ^BAE ¼ ^G.3

4.2.3 Final Step of Reduction: From the Divided Line Segment


to Conic Sections

Let ED be a line segment divided at B, G so that (1) and (2) above are satisfied
(Fig. 10). Draw ABZ perpendicular to ED with AB = BG and BZ = GD, and then
complete the rectangle BZTE. Then ZA  AB = DB  BG = BE2, and since

3
Notice that B, K, A, Z in Fig. 7 correspond (respectively) to E, B, G, and D in Figs. 9 and 10.
4 Abū Sahl al-Kūhī on the Regular Heptagon 91

Fig. 9

Fig. 10

AB = BG and BE = TZ we may write ZA  BG = TZ2, which says that the point T


lies on a parabola whose vertex is A and whose parameter is BG.
On the other hand, by (1) GE  EB = GD2; but, GD = BZ = ET, so
GE  EB = ET2 so that T lies on a hyperbola with vertex B whose transverse side
and parameter are both equal to the segment BG.
Our analysis has now led us to two conics–a parabola and a hyperbola–both
determined by the division of ED at B, G. The intersection point, T, of these two
conics determines the lengths ET and TZ, and these produce the remaining two
segments, GD = ET and EB = TZ, with the property that the line EBGD is divided
at B and G so that (1) and (2) are satisfied. Thus given BG, the side of the heptagon
we wish to construct, we may construct the segment EBGD, then the Δ(ABG), and
finally the heptagon. Of course, once a heptagon has been constructed in some
circle it may be constructed in any other circle by similarity (Elements IV,2).
Abū Sahl was under no illusions that the conic sections could be constructed by
straightedge and compass. As we have mentioned, he wrote a special treatise to
describe an instrument, the complete compass, that could draw conic sections.
Rather, the point of this treatise is that if one is given the next class of curves
beyond a straight line and circle, namely the conic sections, then one may construct
in any circle a regular heptagon. Many centuries after Abū Sahl’s time mathe-
maticians, following Descartes’ discovery of analytic geometry, would begin to
92 3 Geometry in the Islamic World

classify curves according to the degree of the algebraic expression for the curves—
thus “quadratic”, “cubic”, etc. In the ancient world, however, problems were
described as plane, solid, or curvilinear accordingly as it was possible to solve them
by straight lines or circles, sections of cones, or in more complicated ways. In this
context Abū Sahl’s proof may be seen as showing that the construction of a regular
heptagon belongs to an intermediate class of problems whose solutions demand at
worst solid curves. Thus, he limits both the level of difficulty of the problem and the
means necessary to solve it, and he places the problem within the context of the
known mathematical theory of conic sections.

5 The Construction of the Regular Nonagon

5.1 Verging Constructions

The construction of the regular nonagon, i.e., the regular polygon with nine equal
sides, is a special case of the trisection of the angle, since the central angle of a
nonagon is 360°/9 = 120°/3 (Fig. 11). But 120° is the central angle subtending the
side of an equilateral triangle inscribed in a circle, so the regular nonagon in a circle
can be constructed by trisecting this angle. This was well known to the ancient
Greeks, and Pappus of Alexandria gives three methods of trisecting an angle, all of
which use conic sections. The only ancient method that seems to have been
transmitted to the Muslim scholars may be found in the works of Thābit b. Qurra
and his patron and colleague Aḥmad b. Shākir b. Mūsā. However, the Greek treatise
from which these two worked seems to have been lost, for near the end of the tenth
century the geometer ‛Abd al-Jalīl al-Sijzī, who, as a younger contemporary of Abū
Sahl al-Kūhī, participated in solar observations with him at Shirāz in 969–970,
wrote, “It was not possible for any of the ancients to solve this problem (trisecting
the angle) despite their strong desire … .” What al-Sijzī meant by this becomes
clear later on when he refers to “Another lemma of one of the ancients (for tri-
secting an angle) that uses a ruler and moving geometry (i.e., verging construc-
tions), but which we have to solve by fixed geometry.” Thus al-Sijzī knew of
ancient procedures for trisecting the angle, but they were of a kind he called
“moving geometry”. What al-Sijzī was referring to by the phrase “moving geom-
etry” is a kind of construction that Apollonios and other Greek writers called
vergings. In a verging construction, one is given two curves, usually straight lines
or circular arcs, point P not on the curves and also a straight line segment AB. The
problem is to construct a straight line segment CD = AB so that one endpoint lies
on one of the given curves, the other endpoint on the other, and so that CD verges
toward P, that is, when extended it passes through P (Fig. 12a).
A verging construction was used by Hippocrates of Chios in the early fourth
century B.C. to find the area of crescent shaped figures called lunules, and later
Archimedes used vergings to prove theorems on the spiral. There are as well, in
Arabic manuscripts ascribed to Archimedes, uses of vergings to trisect an angle
5 The Construction of the Regular Nonagon 93

Fig. 11

Fig. 12 (a), (b)

and, as we saw in the previous section, to construct the regular heptagon—and all of
these are done without further comment. This may be because he felt that such
constructions were as legitimate as any others. In any case, geometers well before
his time knew how to perform such constructions by means of the conic sections,
94 3 Geometry in the Islamic World

and Apollonios, a younger contemporary of Archimedes, wrote a two-part book on


vergings that could be done by a straightedge and compass alone.

5.2 Fixed Versus Moving Geometry

By the tenth century, however, some geometers felt that verging constructions were
not acceptable as independent operations and they tried to find other solutions to
problems that their predecessors had solved by vergings. The solution al-Sijzī offers
to the problem of trisecting the angle is one that seems to have originated with the
Muslim geometers. Al-Sijzī refers to its main lemma as “the lemma of Abū Sahl
al-Kūhī”; furthermore, the whole trisection appears in Abū Sahl’s work, so it seems
fairly certain that this trisection is one more discovery of the gifted geometer from
the region south of the Caspian Sea. Since Abū Sahl was at Shirāz with al-Sijzī
969–970 it is possible that al-Sijzī learned of the trisection then.

5.3 Abū Sahl’s Trisection of the Angle

Trisecting the angle according to Abū Sahl’s method depends on solving the fol-
lowing problem: Let the semicircle AZD be given and let AD be its diameter, with H
its center (Fig. 12b). Let the angle ABG also be given, with B on AH. It is required
to find on the diameter a point E so that if EZ is parallel to BG then EZ2 = EH  ED.
Construction. To do this we construct on AH as diameter the hyperbola HZL,
whose parameter and transverse side are both equal to AH so that the angle the
ordinate makes with the diameter is equal to ^ABG. (This construction was
explained by Apollonios in The Conics, Book I, Props. 54 and 55, and we shall later
on see a practical method due to Ibrāhīm ibn Sinān for constructing such a
hyperbola.) Let this hyperbola cut the semicircle at Z, and from Z draw the ordinate,
ZE (which will be parallel to BG). Then, by the properties of the hyperbola,

EH  ED=EZ2 ¼ Transverse sideð¼ AHÞ=Parameterð¼ AHÞ;

so that EZ2 ¼ EH  ED; and the problem is solved.


Now to trisect an arbitrary acute angle ABG according to al-Kūhī’s method
extend the side AB in a straight line to D, where the length BD may be chosen
arbitrarily (Fig. 13). On AD as diameter draw the semicircle AGZD and let H be the
midpoint of AD. Draw EZ parallel to BG, where E is chosen, as above, so that
EH  ED = EZ2. Then draw ZH and ZD, as well as BT parallel to ZH. Then
^ABT¼2^TBG. Then ^ABG¼ ^ABT þ ^TBG¼ 2^TBG þ ^TBG¼ 3^TBG,
so we have trisected the given angle ABG.
To prove the foregoing, notice that the condition EH  ED = EZ2, which the
point E satisfies, implies that ED/EZ = EZ/EH, and since the two triangles HEZ,
5 The Construction of the Regular Nonagon 95

Fig. 13

ZED have the angle E in common and the pair of sides containing this angle
proportional it follows that they are similar, so ^EHZ = ^D . However, the two
sides ZH, HD of Δ(ZHD) are equal, so that ^HZD = ^D, and thus
^EZH = ^HZD. Now ^EHZ is an exterior angle of the Δ(ZHD), so that
^EHZ = ^HZD + ^D = 2^HZD. However, since ^BEZ is exterior to Δ(ZEH),
it follows that ^BEZ = ^EZH + ^EHZ ¼ ^HZD + 2^HZD = 3^HZD. Thus,
^EHZ = (2/3) ^ABG, and, since BT is parallel to ZH, ^ABT = (2/3)^ABG,
i; e: ^TBG = (1/3)^ABG.

6 Construction of the Conic Sections

6.1 Life of Ibrāhīm b. Sinān

The conic sections were not used for theoretical purposes only. Indeed, the Greeks had
realized that as the sun traces its circular path across the sky during the day, the rays
that pass over the tip of a vertical rod set in the earth form a double cone, and, because
the plane of the horizon cuts both parts of this cone, the section of the cone by the
horizon plane must be a hyperbola on a horizontal surface (see Plate 1). It was
therefore of use to instrument makers to know how to construct hyperbolas, since it
would be necessary to engrave or cut them on the surfaces of sundials. No doubt there
were for this purpose “tricks of the trade” that craftsmen used, and, perhaps, few
artisans ever looked into a book explaining how to draw a hyperbola. However,
whatever may have been the relationship of theory to practice, such treatises were
written, one of them by the grandson of Thābit b. Qurra, named Ibrāhīm b. Sinān.
Although his life ended by a liver tumor in A.D. 946 at the age of 37, his surviving
96 3 Geometry in the Islamic World

Plate. 1 Diagrams of a sundial for the latitude of Cairo and a universal sundial in a fifteenth
century Egyptian treatise on sundial theory by the muwaqqit al-Karādīsī. (A muwaqqit is a person
who determines the times of prayer in Islam.) The two hyperbolas represent the paths of the
shadow of the pointer (miqyās) at the two solstices. (Taken from MS Cairo Dār al-kutub riyāḍa
892. Courtesy of the Egyptian National Library.)

works ensure his reputation as an important figure in the history of mathematics. His
treatment of the area of a segment of a parabola is the simplest that has come down to
us from the period prior to the Renaissance. (He tells us he invented the proof in order
to rescue the family’s scientific reputation when he heard accusations that his grand-
father’s method was too long-winded.) One of his works we have already referred to,
namely the one titled On the Method of Analysis and Synthesis in Geometrical
Problems, shows his interest not only in particular problems but in general methods or
theories as well. In addition, in his work on sundials he treats the design of all possible
kinds of dials according to a single, unified procedure, and it represents a fresh,
successful attack on problems that had often defeated his predecessors.
In this section, we shall focus on his another work, On Drawing the Three Conic
Sections. This work contains a careful discussion, with proofs, of how to draw the
parabola and ellipse, as well as three methods for drawing the hyperbola. Perhaps so
many methods were given for the hyperbola because it was the one of the most
interest to the instrument makers, although they often used tables in preference to
6 Construction of the Conic Sections 97

geometric constructions.4 From this work we shall present two selections, the one
dealing with the construction of the parabola, which is necessary for the construction
of burning mirrors, and the other giving one of the three methods for drawing the
hyperbola.

6.2 Ibrāhīm b. Sinān on the Parabola

Ibrāhīm’s method is the following. On the line AG (Fig. 14) mark off a segment AB
and construct BE perpendicular to AB. Now on BG pick as many points H, D, Z,…
as you wish. Starting with the point H, draw the semicircle whose diameter is AH,
and let the perpendicular BE intersect it at T. Through T draw a line parallel to AB
and through H draw a line parallel to BE. Let these lines intersect at K.
Next, draw a semicircle on AD as diameter and let it intersect BE at I. As before,
draw lines through I and D parallel to AG and BE, respectively, and let them meet
at L. Do the same construction for the remaining points Z,… to obtain corre-
sponding points. Then the points B, K, L, M, … lie on the parabola with vertex B,
axis BG, and parameter AB. If K′, L′, M′,… are chosen on the prolongations of KH,
LD, MZ,… so that KH = HK′, LD ¼ DL0 ; MZ ¼ ZM0 ; … then they too lie on the
parabola.

Fig. 14

4
For more on the use of tables for constructing sundials, with a photo of part of a table composed
by al-Khwārizmī, see the article “Mizwala” by D.A.King. in Encyclopæedia of Islam II, Vol. 7,
1993 pp. 210–211.
98 3 Geometry in the Islamic World

Ibrāhīm proves that K is on the parabola he has described as follows: If the


parabola does not pass through K let it pass through another point on KH, say N.
Then NH2 = AB  BH, by the property of the parabola. On the other hand, since TB
is perpendicular to the diameter of the semicircle ATH, it follows from Elements
II,14 that TB2 = AB  BH. Further, by construction, TBHK is a parallelogram so
TB = KH. Thus, KH2 = TB2 = AB  BH = NH2, and so KH = NH, which is a
contradiction. Thus K lies on the parabola we described, and the same proof, with
the names of points changed, applies equally well to L, M,… and so the validity of
the construction has been shown.
The reader who would like to try Ibrāhīm’s method can make it easier by
drawing the successive semicircles through A arbitrarily rather than taking their
diameters as given in advance. This saves bisecting the lines AH, etc.

6.3 Ibrāhīm B. Sinān on the Hyperbola

As we have said, this is only one of the three methods he gives, but it is certainly
the one most easily done. On a fixed segment AB (Fig. 15) draw a semicircle and
prolong its diameter AB in the direction of B. On the half of the semicircle near B
pick points as G, D, H, …and at each of these points construct the tangents to the
semicircle GZ, DT, HI, … Let these tangents meet the diameter extended at Z, T, I,
…, respectively, and through these points draw the parallel straight lines ZK, TL,
IM, … making an arbitrary angle with the line AB. On these lines, on the same side
of AB, cut off ZK = GZ, TL = DT, IM = HI, … Then the points K, L, M, … lie on
a hyperbola.
Indeed, since the lines GZ, DT, HI, … are tangents to a circle it follows from
Elements III, 36 that GZ2 = ZB  ZA, DT2 = TB  TA, HI2 = IB  IA, … and since
KZ = ZG, etc., it follows that KZ2 = ZB  ZA, LT2 = TB  TA, and MI2 = IB  IA.
According to the symptom of the hyperbola, given earlier, these relations say that
the points B, K, L, M, … lie on a hyperbola with the line AB as a diameter, whose
ordinates all make with the diameter angles equal to KZB and whose parameter and
transverse sides are both equal to AB. Here again, the rest of the one branch of the
hyperbola can be constructed simply by prolonging each of DZ, LT, MI, … an
equal length beyond ABG to K′, L′, M′, ….

Fig. 15
7 A Problem in Geometrical Optics 99

7 A Problem in Geometrical Optics

The field of geometric optics was a rich source of problems to intrigue the geometer,
one that provided several problems whose solutions required the application of conic
sections. As early as Euclid’s Optics one finds books on geometrical optics, and a
major ancient optical treatise was Ptolemy’s Optics. Both were known in the medieval
Islamic world. Euclid knew the Law of Reflection and Ptolemy conducted experi-
ments to try to discover the Law of Refraction. And, probably about the same time as
Apollonios wrote his Conics another Greek, Diocles, composed a treatise on the focal
property of parabolic mirrors,5 and such “burning mirrors,” as they were called,
constituted another topic treated in Arabic writings on the subject.
Certainly one of the most difficult construction problems solved by geometers in
Islamic lands was a problem from geometrical optics known today as “Alhazen’s
Problem.” (The name stems from the Latin version of Ibn al-Haytham’s full name,6
and the problem appears in his great work, Optics, commemorated on the Pakistini
stamp shown here.)
In Book V of his Optics Ibn al-Haytham solves the following problem: Given a
convex or concave spherical mirror of known radius, the positions of the viewer and
of the object viewed, find the point on the mirror at which the viewer will see the
reflection of the object. Both the eye of the viewer and the object are taken to be
points, so the problem, stated for a convex mirror in terms of the diagram below, is
as follows:

Given a sphere (Fig. 16) with center A and known radius, as well as the positions
of the viewer V and of an object, O, visible in the mirror, find the point, R, on the
mirror at which the viewer sees the object. By the law of reflection that will be the
point R on the circumference of the great circle in the plane AOV such that the lines
OR and VR make equal angles with the line, t, tangent to the circle at R.
Ibn al-Haytham needed six lemmas to solve this problem for convex and con-
cave mirrors, and—as was the case with al-Kūhī in a different problem—he had to
use conic sections to prove the possibility of some of the constructions needed in
the lemmas. We shall content ourselves here with the proof of one case of a key
lemma and one case of a key construction, and, in doing so, we shall follow the

5
See Toomer, 1976.
Ibn al-Haytham’s ism was Al-Ḥasan, which the Latin authors rendered as “Alhazen.”
6
100 3 Geometry in the Islamic World

Fig. 16

exposition of Al-Mu’taman b. Hūd, who—in his mathematical collection known as


The Book of Completion (Istikmāl)—simplified the proofs of some of the lemmas.7
Ibn Hūd was, near the end of the eleventh century, king of Saragossa, one of the
number of small kingdoms making up Al-Andalus. He intended that The Book of
Completion was to have two parts, but only the first part has been found. However,
its contents show that scholars in the western part of the Muslim world had access
to much the same ancient heritage that scholars in the eastern part had, as well as
some of the great works written by eastern scholars themselves, including Ibn
al-Haytham’s Optics.
Lemma 4 Given a right angle triangle ABG, (Fig. 17) with a right angle at B, and a
point, D, on the line containing one of the two sides, say GB, containing the right
angle. It is required to construct a line DMK, intersecting AB at M and AG at K so
that the ratio of the segment MK to the segment KG is equal to a given ratio.
Caption: In Fig. 17 above, solid lines indicate the parts of the figure that are
given. Subsequent constructions are indicated by dashed lines.
Ibn al-Haytham’s and al-Mu’taman’s proofs of this lemma are basically the
same, although, since the latter is shorter we shall follow it. There are three possible
positions for the point D relative to the side AB: outside of side GB in the direction
of B, or outside of GB in the direction of G, or inside GB. We shall take the case
when D is on line GB beyond B.8

7
We follow the numbering of the lemmas in Hogendijk 1996.
8
Ibn Hūd provides a different figure for each case, but the proof applies equally to all figures. In the
statement and proof of this lemma we have taken considerable liberties with Ibn Hūd’s presen-
tation, but we have followed each step of the proof as given in Hogendijk’s translation.
7 A Problem in Geometrical Optics 101

Fig. 17

Construction for Lemma 4: From D draw line DE parallel to side AB of the


triangle ABG, and let it meet line AG at point E. Construct a circle passing through
the three points, G, D, and E, and draw the line DA. Because ED is parallel to AB
(which is perpendicular to GBD) ∠GDE is right and so GE is a diameter of circle
GDE. At point E on diameter GE construct ∠GEZ = ∠DAB. From point Z draw a
secant to the circle, ZHT, that meets diameter EG at a point T such that the ratio
AD:HT is equal to the assumed ratio.9 Then draw DH and let it meet diameter EG at
K and line AB at M. Then MK:KG = AD:HT, which (latter) is the assumed ratio.
Proof: Draw line GH. Then ∠GHK = ∠GED (= ∠GAM), and
∠MAD = ∠GEZ = ∠GHT. The latter is because the opposite angles GEZ and
GHZ of the quadrilateral GEZH in the circle are supplementary10 and ∠GHT and
∠GHZ are obviously supplementary. Thus ∠KHT = ∠KAD. And the opposite
angles at point K are equal, so triangles KAD and KHT are similar. By the same
reasoning, triangles KAM and KHG are similar. Therefore, the assumed ratio, AD:
HT, is equal to the ratio MK: KG, which is what we wanted to prove.
Lemma 5 Let AB be a given circle (Fig. 18) with center G, and D and E be two
points outside the circle. From D and E, it is required to draw two lines meeting at
the circumference of the circle at a point B such that the tangent to the circle at B
bisects angle DBE.

9
Ibn al-Haytham shows how to do this in the previous lemma, Lemma 3.
10
By Elements III, 22.
102 3 Geometry in the Islamic World

Fig. 18

Proof: Suppose, first, that D and E are at the same distance from the center. First,
draw lines GD, GE, DE and circumscribe triangle GDE by a circle. Let it meet the
given circle AB at point B and join DB, BE I say that point B is what we wanted.
Proof of this: We extend line DG beyond the center, G, to some arbitrary point
Z, to form an exterior angle of Δ(GDE). (Fig. 18) Since ∠GBD is equal to each of
the angles GED, GDE [Elements III-21], it is half of the exterior ∠EGZ. But ∠DBE
is equal to ∠DGE. Thus half of ∠DBE plus ∠GBD is a right angle. If one draws a
perpendicular to line GB from point B, it (the perpendicular) is tangent to the circle
and it bisects ∠DBE.11 And this is what required.
Now, suppose the distances of the two points to the center are not the same, and
let the points be points E, H (Fig. 19). Draw line HG and extend it to meet the circle
at point A. Now draw GE and an arbitrary line, TL. Then divide TL at point K such
that the ratio of TK to KL is equal to the ratio of HG to GE. Also, bisect TL at M,
and draw from M a perpendicular, MN.
With vertex at point L (Fig. 20) construct ∠MLN equal to half of ∠EGA. And
from point K draw line KCO, intersecting NL at O, in such a way that the ratio of
CO to OL is equal to the ratio of HG to GA.12 On line HG at point G make ∠HGB

11
Hogendijk 1996 (p. 83 n. 30) points out that this construction is found in Ptolemy’s Optics,
Book IV.
12
A marginal remark in the manuscript states, correctly, “It (the construction of KCO) has been
shown in the previous proposition [Lemma 4].”
7 A Problem in Geometrical Optics 103

Fig. 19

Fig. 20
104 3 Geometry in the Islamic World

equal to ∠COL. Let line GB meet the circle at point B, and then draw lines HB and
BE. Then the tangent [to circle AB] at point B bisects ∠HBE.13
Proof of this: We join LC, CT, and we draw from point L line LF parallel to line
CT, to meet line CK at point F. We make on line GH, ∠HGD equal to ∠CLF. Let
line GD meet line BH at point D. We join DE. Then, since the ratio of HG to GB,
which is equal to GA, is equal to the ratio of CO to OL, and the angle at G is equal to
the angle at O, triangle GHB is similar to triangle COL. Thus ∠GHD = ∠LCF. But
∠HGD = ∠CLF. Therefore, GH:GE is equal to TK:KL, which is equal to CK:KF,
and to CL:LF, which is equal to HG:GD.14 Thus HG:GD = HG:GE. Therefore,
GE = DG. Now extend line DG toward Z. Then ∠AGZ = 2∠CLK, but the whole
∠AGE was assumed to be 2∠NLM. Therefore, ∠ZGE = 2∠OLC, which (latter) is
equal to ∠GBH. Since ∠ZGE is an exterior angle of the isosceles triangle GED, it is
twice ∠GED. Therefore, ∠GBD = ∠GED, so G, B, D, and E are concyclic.15 Thus
the tangent at point B bisects ∠DBE, as above, which is what we wanted to prove.16

8 Geometry with a Rusty Compass

An aspect of Islamic civilization that has always impressed outsiders has been the
elaborate geometrical designs executed in wood, tile, or mosaic and found in
abundance throughout the Islamic world. For example, the especially fine regular
tilings of the plane found in the Alhambra at Granada in Spain are admired the
world over. A crafts tradition of such sophistication involves a considerable amount
of geometrical knowledge, even if that knowledge was simply passed on from
master to apprentice rather than being written down (see Plates 2, 3, and 4).
Of course, there had been a strong tradition of geometrical design in the Middle
East since the time of ancient Egypt, and this continued both in ancient Greece and
elsewhere. At some time geometers became aware of this tradition and the problems
the artisans solved, and they began to try to justify the procedures and to see how
far various methods could be pushed.
For example, in the Arabic version of the eighth book of Pappus of Alexandria’s,
Mathematical Collection is a very interesting section on geometrical constructions that

13
The construction is essentially that of Ibn al-Haytham in Optics V, but the proofs are not the
same.
14
By definition of K, GH:GE = TK:KL according to the definition of K. And TK:KL = CK:KF
because CT and LF are parallel. Since LM = MT and CM is perpendicular to LT we have angle
CLM = angle CTM. Because CT and LF are parallel, ∠angle CTM = ∠angle MLF. Hence ∠angle
CLK = ∠angle KLF, so CK:KF = CL:LF, by Elements VI, 13. Moreover, CL:LF = HG:GD
because also τριανγλε Δ(CLF) is similar to Δ(HGD) triangle HGD, which is a consequence of
angle since ∠GHD = ∠angle LCF ∠and angle HGD = ∠angle CLF.
15
By Elements III, 21.
16
The proof is much shorter than that of Ibn al-Haytham. Note, however, that there is no figure for
the case when the two points are outside the circle.
8 Geometry with a Rusty Compass 105

Plate. 2 This part of the facade of the Shir Dor madrasah in Samarqand illustrates the variety of
elements to be found in Islamic art. It combines calligraphic elements (on the border and inside
arch) with arabesques, geometric designs and both anthropomorphic and zoomorphic elements

Plate. 3 This detail from the facade of the Friday mosque in Isfahan combines pentagons,
octagons, and stellated decagons with nonconvex pieces to tile the plane. The pattern could
obviously the extended infinitely in all directions
106 3 Geometry in the Islamic World

Plate. 4 An example from Isfahan of the kind of ceiling known as al-muqarnas They were
sufficiently common in Islamic architecture for al-Kāshī to devote a section of his Calculators’ Key
to their theory

are possible using only a straightedge and a compass with a restricted opening,
sometimes called a “rusty compass”.17 Since the rest of the eighth book is devoted to
instruments and machines of interest to people doing various crafts, it seems likely that
this section too was aimed at addressing problems encountered by craftsmen.
The continuing Islamic interest in these problems is witnessed by the fact that the
Arabic text of Book VIII of Pappus’s Collection was copied late in the tenth century
by al-Sijzī, whom we have mentioned earlier in connection with the regular hep-
tagon, from an earlier copy that belonged to the Banū Mūsā, the patrons, and
friends of the ninth-century mathematician, Thābit b. Qurra.
Another treatise on geometrical constructions with various restrictions on the
tools is attributed to Abū Naṣr Al-Farabī, who is best known today for his important
commentaries on Aristotle and his great work on music. He was born in 870, when
the Banū Mūsā were old, and he taught philosophy in both Baghdad and Aleppo, an
important trading city in northern Syria. He lived a long life of active scholarship
and was killed by highway robbers outside Damascus in the year 950, not too long
after Ibrāhīm b. Sinān died. In addition to his other work, he wrote a treatise with
the title, A Book of Spiritual Crafts and Natural Secrets in the Details of

17
Of course, all compasses have some restriction on their opening, even if they are not rusty and,
so, many artisans would face the same kind of problem as one would face with a rusty compass.
8 Geometry with a Rusty Compass 107

Geometrical Figures. Later, Abū al-Wafā’, whom we shall speak of in the chapter
on trigonometry, and who was young when Al-Farabī died in 950, incorporated all
of Al-Farabī’s work in his own, more prosaically titled treatise, On Those Parts of
Geometry Needed by Craftsmen. It is from this treatise that we have chosen the
following extracts, where the numbering of the problems is our own.
Problem 1 To construct at the endpoint A of a segment AB a perpendicular to that
segment, without prolonging the segment beyond A.
Procedure. On AB mark off with the compass segment AC (Fig. 21), and, with the
same opening, draw circles centered at A and C, which meet at D. Extend CD
beyond D to E so that ED = DC. Then ^CAE is a right angle.
Proof. The circle that passes through E, A, C has D as a center since
DC = DA = DE. Thus EC is a diameter of that circle and therefore ^EAC is an
angle in a semicircle and hence is a right angle.
Problem 2 To divide a line segment into any number of equal parts.
Procedure. Let it be required to divide the line segment AB (Fig. 22) into the equal
parts AG = GD = DB. At both endpoints erect perpendiculars AE, BZ in different
directions and on them measure off equal segments AH = HE = BT = TZ. Join H
to Z and E to T by straight lines, which cut AB at G, D, respectively. Then
AG = GD = DB.
Proof. Indeed, AHG and BTD are two right triangles with equal angles at G and
D (and therefore at H and T). In addition HA = BT. Thus the triangles are con-
gruent and so AG = BD. Also the parallelism of HG and ED implies that the two
triangles AHG and AED are similar, and thus DG/GA = EH/HA. But, EH = HA
and so DG = GA.

Fig. 21
108 3 Geometry in the Islamic World

Fig. 22

Fig. 23

Problem 3 To bisect a given angle, BAG.


Procedure. The Euclidean method (explained in Elements I, 9) involves cutting off
equal segments AB, AG on the two sides of the angle, constructing the equilateral
triangle on BG, and then joining A, D to bisect the angle. According to Abū al-
Wafā”s variation of this, shown in Fig. 23, AB and AG are marked of on the sides
of angle A, both equal to the fixed compass opening. Then, with the same opening,
BD and DG are constructed on the other side of line BG from A, so Δ(BGD) is
isosceles, not necessarily equilateral, but AD still bisects angle A.
Next, Abū al-Wafā’ finds the center of a given circle. We shall use this in
explaining the next rusty compass construction, but we leave it for the reader to find
the construction that locates the center of a given circle.
Problem 4 To construct a square in a given circle.
Procedure. Locate the center S and draw a diameter ASG (Fig. 24) With compass
_ _ _ _
opening equal to the radius (Exercise 9) mark off arcs AZ; AE; GT and GH and
draw the lines ZE and TH, which cut the diameter at I and K. Draw ZK and TI,
which intersect at M, and then draw the diameter through S, M. Let it meet the
8 Geometry with a Rusty Compass 109

Fig. 24

_ _
circle at D and B. Then ADGB will be a square. Proof. Since ZA ¼ AE the
_
diameter GA bisects the arc ZE and therefore GA is perpendicular to ZE, the chord
of that arc. Similarly GA is perpendicular to TH, and so ^TKI and ^ZIK are right.
Since TH and ZE are chords of equal arcs they are equal and therefore their halves,
TK and ZI, are equal, and since they are also parallel (both being perpendicular to
GA) the figure TKIZ is a rectangle. Its diagonals ZK and TI therefore are equal and
bisect each other, and so MK = MI, i.e., Δ(MKI) is isosceles. Since the equal
chords ZE and TH are equidistant from the center (Elements III,14) , KS = SI, and
so in the isosceles triangle MKI the line MS bisects the side KI and is therefore
perpendicular to the side. Thus the diameter DB is perpendicular to the diameter
GA and ADGB is a square.
As a final example of geometry with a rusty compass we take the following from
Abū al-Wafā”s treatise.
Problem 5 To construct in a given circle a regular pentagon with a compass
opening equal to the radius of the circle.
Procedure. At the endpoint A of the radius DA erect AE perpendicular to AD
(Fig. 25), and on AE mark off AE = AD, then bisect AD at Z and draw the line ZE.
On this line mark off ZH = AD and bisect ZH at T. Then construct TI perpen-
dicular to EZ and let TI meet DA extended at I. Finally, let the circle with center I
_
and radius AD meet the given circle at points M and L. Then ML is one-fifth of the
circumference of the given circle, and the perpendicular bisector of the chord ML
_
bisects the complement of ML relative to the circle at O. Again, the perpendicular
_ _
bisectors of the chords of MO and LO bisect the arcs themselves at N and P,
respectively, and therefore the circle is divided into five equal arcs, whose chords
will be the sides of the regular pentagon.
Abū al-Wafā’ proves the validity of the construction as follows. Draw the lines
LD, LI, and LA (Fig. 25). First of all, the triangles TIZ and AEZ are congruent
110 3 Geometry in the Islamic World

Fig. 25

since they are right triangles with the common angle Z and the side TZ = AZ. Thus,
EZ = ZI, and so:

ZI2 ¼ EZ2 ¼ EA2 þ AZ2 ¼ DA2 þ AZ2 :

Hence
DA2 ¼ ZI2  AZ2 ¼ ðZI þ AZÞðZI  AZÞ ¼ ID  IA

and so A divides DI into two unequal parts so that the rectangle whose sides are the
smaller part and the whole is equal in area to the square whose side is the larger
part, the division into what the Greeks called “the section”, although we now call it
“the golden section”. Now DA = LI, so LI2 = ID  IA, which may be rewritten as
the proportion ID/LI = LI/IA, and since in the two triangles LIA, DIL the angle I is
common and the sides containing this angle are proportional, it follows that the two
triangles are similar. Thus DL/AL = LI/AI and, since DL = LI, AL = AI.
To finish the proof, we recall that Euclid proved in XIII,9 that if the side of a
hexagon in a circle (AD) is extended in the direction of A by the side of a decagon
in that circle then A divides the whole segment into the golden section, and the side
of the hexagon is the larger of the two segments. Now, Abū al-Wafā’ says, since the
line segment ID has been divided in the golden section at A so that AD is the side of
a hexagon in the circle (i.e., a radius), it follows that AI is the side of a decagon in
the circle and hence the same holds for AL = AI. (Strictly speaking, Abū al-Wafā’
needs to make a slight additional argument to deduce this from the converse of
XIII,9, but it is a straightforward proof by contradiction and he left it out.) Thus AL
is one-tenth of the circumference of the circle.
8 Geometry with a Rusty Compass 111

However, if two circles intersect at L and M then the line joining the two centers
_ _ _
bisects the arcs between L and M. Hence, LA ¼ AM and therefore LM is one-fifth
of the circumference of the given circle.
Abū al-Wafā”s treatise contains a wealth of beautiful constructions for regular n-
gons, including exact constructions for n = 3, 4, 5, 6, 8, 10. It also gives a verging
construction for n = 9 which goes back to Archimedes and the approximation for
n = 7 that gives the side of a regular heptagon in a circle as equal to half the side of
an inscribed equilateral triangle. This approximation, by no means original with
Abū al-Wafā’, was probably ancient even when Heron gave it in his Metrica in the
first century A.D. However, it is a good approximation, and as a practical matter is
much simpler than the exact construction by conics.

9 Al-Mu’taman b. Hūd’s Book of Completion

Among the mathematical gems in Ibn Hūd’s Book of Completion are a proof of
Heron’s theorem not known from other sources, a proof of the invariance of the
cross-ratio under projection, and simplified proofs of some of the lemmas that Ibn
al-Haytham used to solve Alhazen’s Problem, which we discussed above.18 As one
example of these gems we shall give his statement and proof of what is known as
Ceva’s Theorem, once thought to have been first stated by the Italian geometer,
G. Ceva, late in the seventeenth century. The theorem states that if three lines drawn
from the vertices of a triangle to the opposite sides all pass through a given point Z
then there is a surprising relation between the ratios in which these lines divide the
opposite sides. The theorem is true whether the lines from A, B, and G meet inside
or outside the triangle, although Ibn Hūd states and proves it only for the case when
the three lines drawn from the vertices to the opposite sides meet inside the triangle.
Both the statement and proof of the theorem employ the notion of compound
ratio. In modern terms one would think of it as the product of the numbers rep-
resented by the two ratios. But, although the notion was freely used in advanced
mathematics in the ancient and medieval world it was never satisfactorily defined.
An interpolated definition of the term in the Elements says that to compound two
ratios is to multiply the size of one by the size of the other to get another ratio.
Euclid himself gives a hint of what he understood by the notion in the proof of VI,
23 when he says, of three magnitudes K, L, and M, “But the ratio of K to M is
compounded of the ratio of K to L and the ratio of L to M.” In the following, we
shall use the notation “ðWX=YZÞ  ðPQ=RSÞ” for the ratio compounded of the two
ratios WX/YZ and PQ/RS.
Ibn Hūd’s Theorem: From point A in triangle ABG (Fig. 26) line AT has been
drawn to meet side BG at point T, and from point B line BE has been drawn to meet

18
For a translation of these into English see our section on the Magnrib and Al-Andalus in Katz
2016.
112 3 Geometry in the Islamic World

Fig. 26

Fig. 27

line AT at point Z and side AG at point E. The straight line GZ is drawn and is
extended to meet side AB at point D. Then the ratio of BD to DA composed with
the ratio of AE to EG is as the ratio of BT to TG.
Ibn Hūd’s proof uses a result in plane geometry that is often called the “plane
version” of Menelaus’s Theorem. (For the spherical version see our Chap. 6 below.)
The plane version says that, as Ptolemy states it in Almagest I 13, if two straight
lines, BE and GD, are drawn to meet two straight lines, AG and AB, and they
GA GD ZB
intersect at point Z, then ¼  .
AE DZ BE
GE GZ DB
And, also ¼  .
EA DZ BA
Ibn Hūd was well acquainted with the Almagest, but it is also possible that he
learned of this result from the Arabic version of Menelaus’s Spherics. In any case,
he knew the theorem, and, since he uses it without comment in his proof, he was
clearly writing for persons who would recognize the application of a fairly
advanced theorem.
9 Al-Mu’taman b. Hūd’s Book of Completion 113

Ibn Hūd begins his proof with a tacit reference to Menelaus’s theorem by stating
that BD/DA = ðBZ=ZEÞ  ðEG=GAÞ. But ðAE=EGÞ  ðEG=GAÞ ¼ AE=GA. Thus
ðBD=DAÞ  ðAE=EGÞ = ðBZ=ZEÞ  ðAE=GAÞ. And, (here, too, he applies
Menelaus’s Theorem) this ratio is equal to BT/TG. Thus, ðBD=DAÞ  ðAE=EGÞ ¼
BT=TG. That is, what we wanted to demonstrate.

10 Practical Geometry of Measurement

Up to now the material presented in this chapter represents what one might call the
“high end” of geometry in medieval Islam. Our next author, however, is on a less
sophisticated level, and his treatise would, almost certainly, be of most interest to
people whose work is related to surveying or who taught such persons.
The author in question is Muḥammad ibn ‘Abdūn, who lived in the middle half
of the tenth century and wrote a work titled On Measurement. Muḥammad taught
mathematics in Cordova, where he was born, and, after a study of medicine, became
personal physician to Al-Hakam II, one of the Umayyad rulers of Spain.19
Ibn ‘Abdūn’s work is not confined to calculating lengths, areas, or volumes,
however, as the reader will see our selections below. Rather it gives a systematic
treatment of geometrical measurement, and one of the features of his work is, for a
specific type of figure (say a triangle), how to deduce each of the number of
quantities associated with it from knowledge of other associated quantities. He
begins his work by discussing the measurement of rectangles, triangles, parallelo-
grams, and other plane figures. He then writes about a number of solid figures and
circles, and then turns to the problem of measuring segments of circles, both greater
than a semicircle and less than a semicircle.
We begin with Problem 17,20 from his discussion of rectangles.
In this problem he assumes that the diagonal of a rectangle is 10 and the
difference of the sides is 2. He then shows how to find the area (48), and then refers
to an earlier problem for how to find the two sides.
In Problem 18, he assumes a rectangle in which the diagonal is 10 and the area
48. The problem is to find the sides, say x, y, of the rectangle. His procedure is to
square the diagonal (d) and to double the area (xy). He then calculates
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
d 2 þ 2ðx  yÞ, which is 14. He knows this is the sum of the two sides, since
d ¼ x2 þ y2 , so (x + y)/2 = 7. But, then, 1 = 49 – 48 = ðx þ2 yÞ2  x  y ¼ ðxy
2 2
2 Þ .
Hence x − y = 2. Since x + y = 14, x = 8, and y = 6. And Ibn ‘Abdūn observes that
if 2x  y [ d 2 the problem is impossible.
Notice that the rectangles in Problems 7 and 8 have the same dimensions and
their diagonals divide them into 3-4-5 right triangles.

19
In preparing this material we have used the Arabic text of the work established in Djebbar 2005
and 2006.
20
We are using the numbers assigned to the problems in Djebbar’s edition of the Arabic text.
114 3 Geometry in the Islamic World

Fig. 28

Beginning with problem 59 he discusses triangles,21 which, he says, are of three


types: the acute, the obtuse, and the right. He continues that one can tell whether the
angle contained by two sides of a triangle is acute by squaring the two sides
containing the angle. If their sum in greater than the square of the third side then the
angle they contain is acute. If the sum is less then, as he says later in the section, the
angle is obtuse. (In this latter case, however, one can be sure that the perpendicular
from the obtuse angle falls within the base and the method, explained below in
Problem 71, to find its area is still of use.)
In Problem 71 he finds the area of a scalene triangle with acute angles, using the
example of the triangle whose sides are 13 and 15, and whose base is 14. He divides
the procedure into two steps. He first calculates the length of the two segments
into which the altitude of the triangle divides the base. The method is based on
theorems 12 and 13 of Elements II, which relate the sides of a triangle to the
segments into which the altitude divides the base.22
Thus, he interprets Euclid’s geometrical language in terms of arithmetic and
calculates half the difference of the squares of the sides containing the
angle,15 13
2 2
2 ¼ 28, and divides this by the base, 14, to obtain 2. He adds this to half
the base, 7, to obtain 9 and subtracts it from 7 to obtain 5. These are the two
segments into which the perpendicular from the vertex opposite the base divides the
base.
He then calculates the length of the altitude using either of these two segments
and the side adjacent to it by the Pythagorean Theorem. Then, of course, one can
use the length of the altitude and the length of the base, 14, to obtain the area of the
triangle.

21
His discussion is strikingly similar to that of Heron, which he could have found in al-Nairizi’s
work.
22
Ibn ‘Abdūn, however, does not discuss a justification for the method.
10 Practical Geometry of Measurement 115

Fig. 29

To find the area of a circle, given its diameter (in Problem 114) he says to
subtract, from the square of the diameter, the seventh of the square and half of the
seventh. The reader may easily verify that this implies the value 3 1/7 for π. Indeed,
in Problem 115 he gives that number as the one to multiply by the diameter to
obtain the circumference.
We turn now to Ibn ‘Abdūn’s discussion of measuring quantities associated with
a segment of a circle.23 Two fundamental quantities associated with a segment are
the length of its base, i.e., the chord bounding the segment, and what he calls its
“arrow,24” that is to say the straight line cut off from the diameter perpendicular to
the base between the base of the segment and its circumference. In Fig. 30 DE is the
arrow of segment AEBD and ADB is its base.

Fig. 30

23
This does not appear to be connected with any material from Heron.
The Arabic word is “sahm.” We shall meet this word again in Chap. 5.
24
116 3 Geometry in the Islamic World

Fig. 31

To find the diameter of a circle given the base and arrow of a segment.
He begins, in Problem 123, with segments less than a semicircle (Fig. 31a) and,
for the sake of discussion, he assumes the arrow is two and the base is eight. He
begins by finding the diameter of the circle containing the segment by halving the
base of the segment, which gives 4. Then square 4 to get 16 and divide that by the
arrow, 2, to get 8. Add 8 and the length of the arrow (which, recall, is 2) to get 10.25
And that is the diameter of the circle.
This result is a special case of the beautiful theorem stated in Elements III, 35,
which says that if two chords of a circle intersect at a point inside the circle, then the
product of the segments of one chord is equal to the product of the segments of the
other chord.

To find the length of arc of a segment given its arrow and the diameter of
the circle:
Ibn ‘Abdūn then attacks the problem of finding the length of the arc of the
segment.26 His procedure is to take the sum of one-seventh of the arrow and the
arrow, then add the result to the diameter. From this one subtracts the positive
difference between the arrow and the radius. If we denote the length of the arc (in
radians) as q, the diameter as D, the base as c, and the arrow as a then Ibn ‘Abdūn’s
rules amount to calculating:
 2
D ¼ a þ 2c  a and q = [(a + a/7) + D] − (r − a), or, as we would simplify it,
2 a + a/7 + r.
The rule is exact for a semicircle, when the arrow is equal to the radius.

25
Al-Khwārizmī gave the same rule in the ninth century in the chapter on measurement in his
Algebra.
26
Since the problem assumes one knows the diameter and the arrow it would be possible to use the
versin function (which we discuss in Chap. 5) to calculate the chord length.
10 Practical Geometry of Measurement 117

If the segment is larger than a semicircle (Fig. 31b) he tells us to add the
difference between the arrow and the radius, to obtain (a + a/7) + 2 r + (a − r), i.e.,
2 a + a/7 + r. (Ibn ‘Abdūn’s avoidance of negative quantities forced him to give
two different rules for what is, in the end, the same formula.)
In Problem 125, for the segment larger than a semicircle Ibn ‘Abdūn takes the
example of a segment of a circle whose base is 8 and whose arrow is 8 (Fig. 31b).
For the diameter this gives him 8 + 16/8, i.e., 10. The length of the arc is 22 1/7 of
those units of which the radius is 5.
He does not carry out the example for the symmetric case of finding q for a
segment less than a semicircle, in which a = 2, but if one uses his rule for this case
one gets 9 2/7. Adding these two values one obtains 31 3/7, which is exactly equal
to 10(3 1/7). The quantity in parentheses is, as we mentioned above, Ibn ‘Abdūn’s
value for π and the value that Archimedes gave in his Measurement of the Circle as
an upper bound for that constant.
So Ibn ‘Abdūn’s values for the length of arcs less than a semicircle are consistent
with those for arcs greater than a semicircle and his rule for the circumference of the
circle. But, there is a serious problem. Ibn ‘Abdūn’s rule for arc length necessarily
gives a value larger than the radius, whatever the angle. It is only for angles over
80° that Ibn ‘Abdūn’s values are acceptable approximations. (For example, for 80°
Ibn ‘Abdūn’s value is 10.5 % too high, but for 90° it is only 5.7 % over the true
value and is exactly right for an angle near 106°.)

Areas of circular segments


Finally, to find the areas of segments less than a semicircle Ibn ‘Abdūn pre-
scribes the rule expressed in modern terms by the formula q2  D2  2c  ða  D2 Þ, and in
the case of segments greater than a semicircle one adds the two products rather than
subtracting them, i.e., q2  D2 þ 2c  ða  D2 Þ. In both cases, the first product represents
the area of the sector determined by the arc q. In the first case, however, one wants
to subtract from that area the area of the triangle bounded by the chord and the two
radii to its endpoints, and in the second case one wants to add that area to complete
the area of the sector to that of the segment. Thus, the approach is quite correct, but
the areas cannot be more accurate than the measurement of the arcs.27

Excercises

1. Use the symptoms of the parabola and hyperbola given earlier to show in the
“Third Reduction” in Sect. 4 that T lies on both the hyperbola and parabola.
2. In an Arabic manuscript found in Bankipore, the following construction of a
regular nonagon inscribed in a circle is given: Let a circle with center D be

27
Again, al-Khwārizmī gave the same rule in the ninth century in the chapter on measurement in
his Algebra. He does not, however, say how to measure the length of the arc.
118 3 Geometry in the Islamic World

Fig. 32

Fig. 33

quartered by two perpendicular diameters AE, ZH and let AB be a chord equal


to a radius (Fig. 32). Let BTG be drawn so that it cuts the diameter ZH at T and
the circle at G and so that TG = AB. Then TD is equal to the side of a regular
nonagon inscribed in the circle ABGH. (1) Show that the point T is obtained by
a verging construction. (2) Prove that TD is the side of a regular nonagon in the
circle. (Hint: If GL is perpendicular to DE and GM is perpendicular to DZ show
that GL = DM = TM.)
3. The following method for constructing a segment of a parabola is found in
North African Arabic manuscript on burning mirrors now housed in the British
Library. Let a line segment PR be given (Fig. 33), whose midpoint is Q, and at
Q erect a segment QD perpendicular to PR. Divide DQ and RQ into the same
number of equal segments. Suppose the points from D to Q are A, B, … and the
points from Q to R are A′, B′, At each of A′, B′, … erect perpendiculars to RQ.
Now let a straightedge passing through P and A intersects the perpendicular
Exercises 119

through A′ at S, let the straightedge through P and B intersects the perpen-


dicular through B′ at T, etc. Then the points S, T, … on the perpendiculars all
lie on a parabola with vertex D. Using R in place of P, A′′ in place of A′, etc.,
we obtain the other half of the parabola. Prove the validity of these statements.
4. In many textbooks on analytic geometry the student is asked to prove that if A
and B are two given points in the plane then the set of all points X so that
|XA| – |XB| = k, where k is a constant and |XA|, |XB| are the distances of X
from A and B, is a hyperbola. Use this fact to prove the validity of the fol-
lowing method, which Ibrāhīm gives for constructing a hyperbola. Let A be the
center of a given circle and B be any point outside this circle. Consider the
collection of all points X so that X is the center of a circle tangent to the given
circle and passing through B. This collection is one branch of a hyperbola.
5. Show on the basis of Euclid XIII,9 that if AD is a radius of the circle and if
IDIA = AD2 then IA is the side of a regular decagon in the circle.
6. In a letter to Abū al-Jūd, al-Bīrūnī asks for a proof that Heron’s construction
for the side of a regular heptagon is not exact. Show this, but also show
that Heron’s construction would be in error by less than 2 mm in a circle of
radius 1 m.
7. Show that the steps of Abū Sahl’s analysis and the construction of the regular
heptagon are equivalent, i.e., not only does the existence of Δ(ABG) imply the
division of a line segment such that (1) and (2) of the “Second Reduction” are
satisfied but the converse is true as well.
8. Show that in Fig. 9 the exterior angle of Δ(BAD) at B equals 3  ^ D.
9. In the rusty compass construction of a square in a given circle Abū a1-Wafā’
assumes that his compass opening is equal to the radius of the circle. Show
there is no loss of generality in this, i.e., if the construction can be done in this
special case, then it can be done in the general case.
10. Given a circle ABG find a rusty compass construction that locates its center.

Bibliography

Bellosta, H. 1991. “Ibrāhīm ibn Sinān: On Analysis and Synthesis.” Arabic sciences and
philosophy 1(2): 211–233.
Berggren, J.L. 1984. “An anonymous treatise on the regular nonagon.” Journal for History of
Arabic Science 5: 37–41.
Hogendijk, Jan P. 1984. “Greek and Arabic constructions of the regular heptagon.” Archive for
History of Exact Sciences 30:197–330.
Hogendijk, Jan P. 1996. “Al-Mu‘taman‘s Simplified Lemmas for solving ‘Alhazen’s Problem’” in
From Baghdad to Barcelona, (ed. J. Cassuleras and J. Samsó), vol. 1, pp. 59–102.
Norman, Jane and Stef Stahl. 1979. The Mathematics of Islamic Art: A package for teachers of
mathematics. The Library and Teacher Resource Center in The Ruth and Harold D. Uris Center
for Education, New York, 1979. Email inquiries to: education@metmuseum.org.
Rosenfeld, B.A. 1988. A history of non-euclidean geometry: Evolution of the concept of a
geometric space. Trans. A. Shenitzer. New York, Berlin, etc.: Springer-Verlag.
120 3 Geometry in the Islamic World

Sabra, A.I. 1982. Ibn al-Haytham’s Lemmas for solving ‘Alhazen‘s Problem’. Archive for History
of Exact Sciences 26: 299–324.
Toomer, G.J. 1976. Diocles on burning mirrors; The Arabic translation of the lost greek original.
Berlin, Heidelberg: Springer-Verlag.
Winter, H. J. J., and W. ′Arafat, “Ibn al-Haitham on the Paraboloidal Focussing Mirror” and “A
Discourse on the Concave Spherical Mirror of Ibn al-Haytham”. Journal of the Royal Asiatic
Society of Bengal, Science 15 (No. 1) (1949), 25–40; and 16 (No. 1) (1950), 1–16, respectively.
Woepcke, F. “Analyse et Extrait d’un Recueil de Constructions Géométriques par Aboûl Wafâ”,
Journal Asiatique (Ser 5), 5, Feb.–March (1855), 218–359.
This contains an exposition, evidently based on a student’s notes in Persian, of Abū a1-Wafā’’s
rusty-compass constructions.
Chapter 4
Algebra in the Islamic World

1 Problems About Unknown Quantities

Many ancient mathematical works contain problems requiring the discovery of an


unknown quantity. Sometimes this is a geometrical magnitude which is related by
the conditions of the problem to known magnitudes, one example being the problem
solved in Euclid’s Elements II,11 of dividing a given line segment AB into segments
AG and GB so that the rectangle whose sides are AB and GB is equal to the square
whose side is AG (Fig. 1). Here there is one known magnitude, AB, one unknown
segment, AG (since GB = AB – AG) and the one condition that AB  GB = AG2.
The reader will recall that this is the division of a line segment into “the section” we
spoke of in Chap. 3 in our discussion of Problem 5 of Abū al-Wafā”s treatise in
which a pentagon was inscribed in a circle.
Another example is found in Prop. 4 of Archimedes’ On the Sphere and
Cylinder, Book II, where Archimedes solves the problem of cutting a sphere by a
plane so that the volumes of the two segments are to one another in a given ratio. In
both of the above problems the unknown quantity is a geometrical magnitude, but
examples where the unknown quantity is a number are abundant. For example,
in a cuneiform text written in Mesopotamia, when the region was ruled by the
successors of Alexander the Great, there is a problem asking for the number that,
when added to its reciprocal, produces a given number. Long before this,
Babylonian scribes, as early as 1800 B.C., were able to solve problems that lead to
quadratic equations, and, although the Babylonians would say “I have added the
area and six times the side of my square and it makes 27” instead of “x2 + 6x = 27”
the procedure that they used for solving the problem is essentially the one that we
use today.
The numerical procedures of the Babylonians are found again in the Greek
writers Heron (fl. A.D. 60) and Diophantos (fl. A.D. 300), where sometimes more
than one unknown is required. This is the case in the following problem solved by
Diophantos in his Arithmetica: To find three (rational) numbers so that the product

© Springer Science+Business Media New York 2016 121


J.L. Berggren, Episodes in the Mathematics of Medieval Islam,
DOI 10.1007/978-1-4939-3780-6_4
122 4 Algebra in the Islamic World

Fig. 1

of any two added to the third gives a square. Other times it would be a single
unknown subject to a variety of conditions, as in this problem found in the Hindu
writer Bhaskara, around A.D. 1150: “Which number is it, which, being divided by
six, has five for a remainder, and divided by five has a remainder of four, and
divided by four has a remainder of three, and divided by three leaves two?.”
For many of the Greek writers, however, the treatment of problems of finding
unknowns was geometric. Thus, in his division of the line segment into “the section,”
Euclid is dealing with a geometric version of the problem of solving the quadratic
equation x2 + ax = a2; for, if we let AB = a and AG = x (Fig. 1), then the condition

AB  GB ¼ AG2 becomes aðaxÞ ¼ x2

and this implies a2 = ax + x2. This requires that, given a segment of length a, we
construct a segment of length x so that x satisfies the condition x2 + ax = a2.
Euclid, however, does not think of the numerical concept of “the length of a
segment,” but instead he looks on x2 + ax = x(x + a) as a rectangle having sides
x and x + a. He literally “completes the square” when he uses Prop. II,6 of his
Elements to the effect that if the square of side a/2 be added to this rectangle,
(x + a)x, then the result is the square of side x + a/2. Then x is equal to the side of
this square, (x + a/2), less the known a/2.
The case of “the section” is one example of how Greek geometers of the classic
period saw such quadratic relationships as those we write as x2 + ax = a2 as
statements about areas. To deal with these statements they had a body of theorems,
e.g., those in Book II of the Elements that allowed them to complete these areas to
yield known squares, whose sides then gave the desired line segments. Many
scholars believe that a historical reason for the Greek emphasis on geometry is that
pffiffiffi pffiffiffi
irrational numbers, such as 2 and 5, necessarily enter whenever one solves
quadratic equations. Since the Greeks lacked even a definition of such numbers they
could not treat them rigorously and so they used the geometrical magnitudes
themselves when they wanted rigorous arguments.

2 Sources of Islamic Algebra

All these parts of Greek mathematics were known to mathematicians of the Islamic
world. The translation of Euclid done by al-Hajjāj late in the eighth century was
followed by several improved translations, culminating in that of Thabit ibn Qurra
late in the ninth century. Thābit also revised an earlier, imperfect translation of
Archimedes’ Sphere and Cylinder, so this work too was available in a good Arabic
version from the tenth century onwards. Finally, the first seven books of
2 Sources of Islamic Algebra 123

Diophantos’ Arithmetica were translated into Arabic by Qusṭā ibn Lūqā of Baalbek,
in present-day Lebanon, probably in the mid-ninth century. Both Qusṭā and Abū
al-Wafā’ wrote commentaries on this work.
Another great source of the Islamic algebra was the Hindu civilization, where,
from the late fifth century A.D. onwards, we have abundant evidence of a highly
developed mathematics. There are many parallels between the Hindu mathemati-
cians and a Greek writer in the numerical tradition like Diophantos. Both used
abbreviations of words to stand for unknowns; but, whereas Diophantos used only
one abbreviation, the Hindu writers used many—the first syllables of the Sanskrit
words for the various colors. Hindu writers had a special sign to denote negative
numbers, and Diophantos, too, had a special sign for subtraction. Finally, both
Diophantos and the Hindus were interested in indeterminate equations, that is,
equations in several unknowns admitting a possibly infinite number of solutions.
One of the greatest early Hindu mathematicians was Brahmagupta, who lived in
the first half of the seventh century A.D. and whose astronomical work, the
Brahmasphuta-siddhanta, has two of its 24 devoted to mathematics. In this work,
Brahmagupta states clearly the rules for multiplying signed numbers and recognizes
that the solutions to some of his problems may be negative numbers. He follows the
earlier Āryabhata (fl. A.D. 500) in giving the general solution to what we would
write in the form ax + by = c as x = p + mb, y = q – ma, where x = p, y = q is a
particular solution and m is any integer. He also showed how to use what we call
“the Euclidean algorithm” to obtain particular solutions p and q. Thus
5x + 12y = 29 has the solution x = 1, y = 2, and the general solution is
x = 1 + 12 m, y = 2–5 m. He also discussed the difficult problem of solving
x2 = 1 + py2, known today as Pell’s equation.
The Muslims learned early in their history of the Hindu achievements in algebra, for
Brahmagupta’s astronomical work was one of those that Indian scholars brought to the
caliph al-Manṣūr around A.D. 770, and it was translated by al-Fazārī into Arabic.
Sanskrit astronomical works were written in the inverse form (perhaps to facilitate
memorization), and the translation must have been no easy task for al-Fazārī. For
example, al-Bīrūnī refers to the habit the early translators of Indian material had of
leaving certain words untranslated, simply spelling them out in Arabic.
The Muslim mathematicians were as ready as their Babylonian and Hindu
predecessors to appreciate the effectiveness of numerical procedures that these
nations possessed for solving quadratic or indeterminate equations. As we have
seen, they had inherited both the Babylonian sexagesimal system and the decimal
system of the Hindus, and these efficient systems provided a good foundation for
numerical mathematics.
On the other hand, the Greek geometrical approach had behind it the authority of
men whom most Muslim scientists admired immensely, and the geometrical rules
had been proved beyond any doubt to be true. Here it was not a case of checking a
numerical answer to see that it solved the problem; rather, there was a general
theory based on sound axioms that yielded proofs of the validity of the methods.
In the next section, we shall see how these two approaches, the numerical and
the geometric, were combined to create a new science.
124 4 Algebra in the Islamic World

3 Al-Khwārizmī’s Algebra

Out of this dual heritage of solutions to problems asking for the discovery of
numerical and geometrical unknowns Islamic civilization created and named a
science—algebra. The word itself comes from the Arabic word “al-jabr,” which
appears in the title of many Arabic works as part of the phrase “al-jabr wa
al-muqābala.” One meaning of “al-jabr” is “setting back in its place” or “restoring,”
and the ninth century algebraist al-Khwārīzmī, although he is not always consistent,
uses the term to denote the operation of restoring a quantity subtracted from one
side of the equation to the other side to make it positive. Thus replacing
5x + 1 = 2–3x by 8x + 1 = 2 would be an instance of “al-jabr.” The word “wa” just
means “and,” and it joins “al-jabr” with the word “al-muqābala,” which means in
this context replacing two terms of the same type, but on different sides of an
equation, by their difference on the side of the larger. Thus, replacing 8x + 1 = 2 by
8x = 1 would be an instance of “al-muqābala.”
Clearly, with the two operations any algebraic equation can be reduced to one in
which a sum of positive terms on one side is equal either to a sum of positive terms
involving different powers of x on the other, or to zero. In particular, any quadratic
equation with a positive root can be reduced to one of three standard forms:

px2 ¼ qx þ r; px2 þ r ¼ qx; or px2 þ qx ¼ r, with p, q, r all positive,

a condition that runs through the whole medieval period in Islamic mathematics.
We shall meet it again in the work of ‛Umar al-Khayyāmī, and it is the rule in
Western mathematics as well through the early sixteenth century. Thus the science
of al-jabr wa al-muqābala was, at its beginning, the science of transforming
equations involving one or more unknowns into one of the above standard forms
and then solving this form.

3.1 Basic Ideas in Al-Khwārizmī’s Algebra

One of the earliest writers on algebra was Muḥammad b. Mūsā al-Khwārizmī,


whose treatise on Hindu reckoning we referred to in Chap. 2. His work on algebra,
The Condensed Book on the Calculation of al-Jabr wa al-Muqābala, enjoyed wide
circulation not only in the Islamic world but in the Latin West as well.
According to al-Khwārizmī there are three kinds of quantities: simple numbers
like 2, 13, and 101, then root, which is the unknown, x, that is to be found in a
particular problem, and wealth, the square of the root, called in Arabic māl. (A
possible advantage of thinking of the square term as representing wealth is that
al-Khwārizmī can then interpret the number term as dirhams, a local unit of cur-
rency.) Another word used for “root” by many writers is “thing” (shay’). In these
terms al-Khwārizmī could list the six basic types of equations as
3 Al-Khwārizmī’s Algebra 125

Roots equal numbers (nx = m).


Māl equal roots (x2 = nx).
Māl equal numbers (x2 = m).
Numbers and māl equal roots (m + x2 = nx).
Numbers equal roots and māl (m = nx + x2).
Māl equals numbers and roots (x2 = m + nx).

All equations involving all the three kinds of quantities and having a positive
solution could be reduced to one of types (4)–(6), the only ones with which
al-Khwārizmī concerns himself.

3.2 Al-Khwārizmī’s Discussion of x2 + 21 = 10x

In following al-Khwārizmī’s discussion of type (4) above we shall use modern


notation to render his verbal account. He discusses this type in terms of the specific
example x2 + 21 = l0x, which he describes as “māl and 21 equals 10 roots,” as
follows (translation adapted from F. Rosen):
Halve the number of roots. It is 5. Multiply this by itself and the product is 25.
Subtract from this the 21 added to the square (term) and the remainder is 4. Extract
its square root, 2, and subtract this from half the number of roots, 5. There remains
3. This is the root you wanted, whose square is 9. Alternately, you may add the
square root to half the number of roots and the sum is 7. This is (then) the root you
wanted and the square is 49.
Notice that al-Khwārizmī’s first procedure is simply a verbal description of our
rule
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
s  ffi
10 10 2
 21;
2 2
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
and his second procedure describes the calculation of 5 þ 52  21, but since all
quantities are named in terms of their role in the problem whenever they appear
(For example, “5” is called “the number of roots”), his description of the solution is
quite as general, if not so compact, as our
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
r ffi
n n2
 m:
2 2

In fact, al-Khwārizmī’s generality is reflected in the remarks that continue those


quoted above.
“When you meet an instance which refers you to this case, try its solution by
addition, and if that does not work subtraction will. In this case, both addition and
126 4 Algebra in the Islamic World

x 21/x

Fig. 2

subtraction can be used, which will not serve in any other of the three cases where
the number of roots is to be halved.”
“Know also that when, in a problem leading to this case, you have multiplied
half the number of roots by itself, if the product is less than the number of dirhams
added to māl, then the case is impossible. On the other hand, if the product is equal
to the dirhams themselves, then the root is half the number of roots.”
In the first of the above paragraphs, al-Khwārizmī recognizes that the case we
are dealing with is the only one where there can be two positive roots. In the second
paragraph, he remarks that there is no solution when what we call the discriminant
is less than zero and he says that when (n/2)2 = m the only solution is n/2. Finally,
he remarks that in the case px2 + m = nx it is necessary to divide everything by p to
obtain x2 + (m/p) = (n/p)x, which can be solved by the previous method. This
shows, by the way, that his coefficients are not restricted to whole numbers.
What distinguishes al-Khwārizmī and his successors from earlier writers on
problems of the above sort is that, following the procedures for obtaining the
numerical solutions, he gives proofs of the validity of these same procedures, proofs
that interpret x2 + 21, for example, as a rectangle consisting of a square (x2) joined
to a rectangle of sides x and 21/x (Fig. 2).

4 Thābit’s Demonstration for Quadratic Equations

Preliminaries
Al-Khwārizmī presents his proofs in terms of particular equations, but Thābit ibn
Qurra in his work gives the demonstrations in general, and for that reason we shall
follow him rather than the earlier al-Khwārizmī.
The first two cases, x2 + px = q and x2 + q = px, sufficiently indicate Thābit’s
approach. In the proofs he uses two theorems from Euclid’s Elements, which we
now state and prove.
Book II, Prop. 5. If a line AE is divided at B and bisected at W then the
rectangle AB  BE plus the square on BW is equal to the square on AW (Fig. 3a, b).
Note that in this proposition B may be on either side of the midpoint W. The two
parts of Fig. 3 show these two cases and are drawn so that GAEM is a rectangle of
sides AE and AG(= AB). The rectangle AB  BE which the theorem speaks of is equal
to the shaded rectangle since AB = BD. The foregoing proposition deals with a line
segment bisected and divided internally. The next proposition deals with a line seg-
ment extended, which we could look on as being bisected and divided externally.
4 Thābit’s Demonstration for Quadratic Equations 127

Fig. 3

Fig. 4

Book II, Prop. 6. If a given line BH is bisected at W and extended in a straight


line BA then rectangle AH  AB plus the square on BW is equal to the square on
AW (Fig. 4).
Proof of II,5. In the case where B is between W and E the truth of the theorem is
most clearly seen by drawing the diagonal GB in the square ABDG and the line
WKL parallel to BD. Finally, draw the line FKH parallel to AB. ABDG is a square
since AG = AB, so WBHK is also a square. Now consider the L-shaped figure
WEMDHK. Such a figure was called a “gnomon” by the Greeks. In the case of the
gnomon WEMDHK, which is made up of the rectangle AB  BE plus the square on
BW, it is what remains when KHDL is taken away from the rectangle WEML. Also
KHDL = AWKF, by Euclid I,43, while WEML = WAGL, since AW = WE. Thus
the gnomon

WEMDK ¼ WEML - KHDL ¼ WAGL - AWKF ¼ FKLG,

and FKLG is equal to the square on AW since it is a square and its side FK = AW.
The proof for the other case of this theorem is similar and is left as an exercise.
Proof of II,6. Construct the square on AB, say ABDG, and the rectangle DBHL.
Then GAHL equals the rectangle AH  AB since AB = AG. The parallel to DB
128 4 Algebra in the Islamic World

through W divides DBHL into two congruent rectangles, since BW = WH.


If WHLK is placed above AB so that WK coincides with AB then we see that the
original rectangle with sides AH, AB is equal to the gnomon EGKWBF; but, if the
square BF2 is added to this figure, the result is the square on GK. The proof is now
complete since the square on GK is the square on AW, because AW = GK.

4.1 Thābit’s Demonstration

Thābit begins his discussion of the validity of the procedures for solving quadratic
equations by discussing the solution of what he calls the first basic form, “māl and
roots equal numbers.” In following, the discussion the reader should know that
Thābit’s term “roots” corresponds to the modern “px” and that he uses the words
“the number of roots” to refer to the coefficient “p.” In Fig. 5 let the square ABDG
represent māl (so its side AB is a root), and suppose the unit of linear measure is
chosen so that BH represents the number of roots. The area DEHB then represents
the term “roots,” and so GEHA represents māl and roots. According to Euclid II,6,
then, GEHA plus BW2 is equal to AW2. But māl and roots is known (for it is
“numbers”), and BW2 is also known since BW is half the given number of roots.
Hence AW2, and thus AW, is known. But x = AB = AW – BW, so x is known.
Finally, Thābit shows the correspondences between the geometric solution given
and the algebraic solution.

1
2 BH = BW $ half the number of roots,
Square on BW $ square of the above,
Rectangle on HA  AB $ numbers,
Square on AW $ sum of the two preceding,
AW $ square root of the sum,
AW – BW = AB $ the square root less half the number of roots.
He adds (p/2)2 to each side to obtain

p2 p2
x2 þ px þ ¼ qþ :
2 2

With his geometrical interpretation of all terms as areas he is able to apply


Euclid II,6 to the left-hand side to obtain
 p2 p2
xþ ¼ qþ
2 2
4 Thābit’s Demonstration for Quadratic Equations 129

Fig. 5

Fig. 6

and therefore x is determined by


rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
p2 p
x ¼ qþ 
2 2

Thābit then shows the validity of the solution for the second basic form, “māl plus
numbers equal to roots,” as follows (Fig. 6a, b). Again, the square ABDG represents
māl and on AB extended the point E is chosen so that, a unit of measure being given,
the length of the line AE is equal to the number of roots. (Thābit in fact shows that
such a point E must lie on the extension of AB.) Now, bisect AE at W and construct
the rectangle with sides GA and AE. Notice that GA is “root” and AE is the number
of roots, so the foregoing rectangle represents “roots,” which is “māl plus numbers,”
and the square (BG) is māl. (We use two opposite corners to denote a quadrilateral.)
Thus when we subtract māl the remaining rectangle is “numbers.” According to
Euclid II,5, “numbers” plus the square on BW is equal to the square on AW. But this
square is known, since AW is half the number of roots, and “numbers” is also known,
so the remaining term, the square on BW, is known, and hence BW itself is known.
Since AW and WB are both known, AB (which is “root”) is also known—in the first
diagram as AB = AW + WB, and in the second as AW – WB.
Thābit does not go into details on the algebra, since he evidently feels the
correspondences between the algebra and the geometry are sufficiently clear from
the previous case.
130 4 Algebra in the Islamic World

5 Abū Kāmil on Algebra

5.1 Similarities with al-Khwārizmī

A writer who was active about the time of Thābit’s death in 901 is Abū Kāmil,
whose epithet “the Egyptian calculator” tells us practically all we know about him.
His work Algebra became quite popular and both the Muslim writer al-Karajī in the
late tenth century and the Italian Leonardo of Pisa, known as Fibonacci, in the late
twelfth century made considerable use of Abū Kāmil’s examples.
There are many similarities between Abū Kāmil’s Algebra and that of
al-Khwārizmī. For example, he follows al-Khwārizmī in naming the basic quanti-
ties numbers, roots and mal. Like al-Khwārizmī’s work, the Algebra is entirely
verbal, and even the numbers are written out. In addition there is the same clas-
sification of equations, with the six types appearing in the same order and, for the
equations involving all three terms, there are the same examples. Finally, like
al-Khwārizmī, Abū Kāmil discusses the geometrical proofs of the procedures in
terms of specific examples rather than in general as Thābit does.

5.2 Advances Beyond al-Khwārizmī

Despite this Abū Kāmil’s work goes beyond that of al-Khwārizmī in giving general
statements of rules which al-Khwārizmī states by means of examples. In addition he
provides proofs of such rules for manipulating algebraic quantities as the following:

ð1Þ ða  pxÞ ðb  qxÞ ¼ ab  bpx  aqx þ pqx2 ;


ða  pxÞ ðb  qxÞ ¼ ab þ bpx  aqx  pqx2 :
pffiffiffiffiffiffiffiffiffi pffiffiffi pffiffiffi pffiffiffiffiffiffiffiffi pffiffiffi pffiffiffi
ð2Þ a  b ¼ a  b and a=b ¼ a= b:
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffi pffiffiffi pffiffiffiffiffi
ð3Þ a  b ¼ a þ b  2 ab:

The identities in (1) are obviously fundamental, while (2) and (3) taken together
allow one to write arithmetic combinations of square roots as square roots. In
addition, (2) is of some practical utility in that it allows us to calculate, say,
pffiffiffiffiffi pffiffiffi
13  5 by multiplying only whole numbers and taking only one square root,
pffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffi
namely 13  5 ¼ 65, which is now easy to see as a bit over 8, something not
pffiffiffiffiffi pffiffiffi
clear from 13  5.
In the case of (1), Abū Kāmil expresses the rule as, “When roots are added to
numbers or subtracted from them, on whichever side they are arranged (i.e., a + px
or px + a), then the fourth part is added”, where “the fourth part” is the product
of the roots, one by the other. Here he speaks only of “the fourth part,” i.e.,
(+px)  (+ qx) or (–px)  (– qx) since he assumes his readers know the signs of the
5 Abū Kāmil on Algebra 131

G B A
x 10–x

E Z

D H

Fig. 7

first three terms of these products. A little later he summarizes the rules of signs for
multiplication, namely (– a)(– b) = + (ab), (– a)(+ b) = –(ab), and (+ a)
(+ b) = + (ab), as, “The product, when the two terms are subtracted, is added; the
subtracted times the added is subtracted; the added times the added is added.”
Abū Kāmil carefully demonstrates rules like ax bx = ab  x2 and a  (bx) = (ab) 
x (where a, b are always specific numbers in the demonstrations), and then he shows
the truth of particular cases of (1), e.g.,

ð10  xÞ  ð10  xÞ ¼ 100 þ x2  20x:

Although Abū Kāmil gives an algebraic proof, based on the distributive law and
the rule of signs, he also gives the following geometric proof.
Proof (Fig. 7). Let the line GA represent the number 10 and GB the “thing,” x, and
complete the square (AD) as in the diagram. Then AB = ED = (10-x)2, so the square
(ZH) = (10 – x)2. Also, (GZ) = (GH) = 10x, so (EH) = (GH) – (EB) = 10x – x2.
Hence, the gnomon (EH) + (GZ) = 20x – x2. Since the large square is 100 it follows
that

ð10  xÞ2 ¼ ðZH) ¼ 100  ð20x  x2 Þ ¼ 100 þ x2  20x:

Abū Kāmil gives no explanation for the last equality, so he presumably feels it is
clear either from the rule of signs or from the figure.
The samples from Thābit and Abū Kāmil suffice to illustrate the demonstrations
algebraists of the Islamic world gave for algebraic rules by geometric theorems, and
we omit the demonstrations of (2) and (3).

5.3 A Problem from Abū Kāmil

Abū Kāmil’s work includes a great variety of problems, 69 in all, amounting to


almost 30 more than the 40 problems al-Khwārizmī explained. One of the most
interesting is Problem 61, which Abū Kāmil states as follows:
One says that 10 is divided into three parts and if the small one multiplied by
itself is added to the middle one multiplied by itself the result is the large one
132 4 Algebra in the Islamic World

multiplied by itself, and when the small is multiplied by the large it equals the
middle multiplied by itself.
Abū Kāmil thus speaks of three unknown quantities x, y, z (assumed to be
positive) satisfying the three conditions

10 ¼ x þ y þ z; z2 ¼ x2 þ y2 and xz ¼ y2 :

He first sets x = 1 and the conditions become

10 ¼ 1 þ y þ z; z2 ¼ 1 þ y2 and z ¼ y2 :

The last two yield (y2)2 = 1 + y2, and this quadratic in y2 he solves to obtain
rffiffiffiffiffiffi sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
rffiffiffiffiffiffi
1 1 1 1
z¼y ¼ þ 1 ;
2
so y¼ þ 1 :
2 4 2 4

Then

rffiffiffiffiffiffi sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
rffiffiffiffiffiffi
3 1 1 1
1þyþz ¼ þ 1 þ þ 1 :
2 4 2 4

If we call this quantity a then we should have a = 10, but it is obviously not
equal to 10; however, a  10/a = 10, so if we set 10/a = b then a  b = 10. This may
be written

ð1 þ y þ zÞ  b ¼ 10; i:e: b þ ðybÞ þ ðzbÞ ¼ 10:

Hence b, yb and zb solve the problem, and what Abū Kāmil has used here is
something known as “the rule of false position.” This algebraic device is, if not as
old as the hills, at least as old as the pyramids, since it is found in ancient Egyptian
texts. In its simplest form one would solve 5x = 24 by “If x = 1, 5  x = 5. What
must I multiply 5 by to make 24? The answer is 24 5 ¼ 4 5. Thus the true value of x is
4
 4
1 4 5 .” Of course in Abū Kāmil’s problem there are more variables and more
conditions, but in each condition (equation) every term is of the same degree (one
or two) and therefore the rule of false position holds valid.
We shall not follow Abū Kāmil’s calculation of the value of b, and shall only
remark that it is a virtuoso performance with the rules of algebra, including his use,
without comment, of the identity
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
11  x2  x2  1  x2  x2 ¼ 5  x2 x2 :
4 4
5 Abū Kāmil on Algebra 133

In the end he obtains b as a root of the equation


pffiffiffiffiffiffiffiffiffiffi
10x ¼ x2 þ 75  3125:

This equation is one of the six standard forms, and he solves it to obtain
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffiffiffiffiffi
b¼5 3125  50:

A similar procedure leads him to an expression for z, the largest root, as

rffiffiffiffiffiffiffiffi sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
rffiffiffiffiffiffiffiffi
1 1 1 1
2 þ 31  78  12
2 4 4 2

and from these two, knowing that x + y + z = 10, he is able to write down an
expression for y.
The solution to this problem, whose key steps are the use of the rule of false
pffiffiffi pffiffiffi
position and the identity for a  b shows that authors in the tenth century were
capable of rationalizing denominators in expressions of the form
pffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffi
a=ðc þ d þ e þ f Þ, which involved dealing with powers of the unknown as
high as the eighth and solving quadratic equations with irrational numbers as
coefficients.

6 Al-Karajī’s Arithmetization of Algebra

6.1 Introduction

Abū Kāmil’s work shows the development of an arithmetic of expressions of the


pffiffiffi
form a þ b q, where a, b, and q are rational numbers with q nonnegative. Although
Thābit and Abū Kāmil apply geometry to algebra, there is in Abū Kāmil’s treatment
pffiffiffi
of the numbers a þ b q a tendency to apply arithmetic to a subject that, long before
him, Euclid had treated geometrically in Book X of his Elements. We have also
seen increasingly complicated examples forcing algebraists to consider more
complicated expressions than those involving only “things” or “squares,” and both
of these tendencies are carried further in the algebra named The Marvelous, com-
posed by Abū Bakr al-Karajī.
Al-Karajī is one of the many remarkable Muslim scientists about whose life we
should like to know more than we do, but we do know that he worked in Baghdad
around the year A.D. 1000 and that in the first decade of that century he dedicated a
book on algebra to a vizier Fakhr al-Mulk. Sometime later he left for “the mountain
countries.” He must have been a man of wide interests, for among his writings
134 4 Algebra in the Islamic World

appear not only treatises on arithmetic, algebra, indeterminate analysis and


astronomy, but on surveying and on finding underground waters as well.
Although writers like Diophantos and Abū a1-Wafā’ indicated the possibility of
using arbitrarily high powers of the unknown, al-Karajī appears to have been the
first person to develop the algebra of expressions containing these powers. His point
of view is that unknown quantities, whether absolute numbers or geometrical
magnitudes, can be “roots,” “sides” or “things” (both of the latter corresponding to
our “x”), or they can be māl (x2), cube (x3), māl māl (x4), māl cube (x5), etc., where
each member is the product of “thing” by the previous member. These different
species of quantity al-Karajī calls “orders” (a terminology also used for the places
of the different powers of 10 in decimal arithmetic), and he observes that the
number common to all orders is 1 (since it is equal to all its powers). In addition,
corresponding to each order (xn) is the corresponding part (l/xn), with the property
that any order times its part equals 1. On this basis al-Karajī developed his program
of treating expressions like “māl māl and four cubes less six units” (x4 + 4x3–6)
and “five cube cubes less two squares and three units” (5x6 – [2x2 + 3]) by rules
modeled on the ordinary rules of arithmetic for adding, subtracting, multiplying,
dividing, and extracting square roots.
R. Rashed has referred to this modeling of the algebra of polynomials on
positional arithmetic as the “arithmetization of algebra”. Al-Karajī was one of the
pioneers in this process, and if his success in arithmetizing algebra was only partial
it is due less to lack of ingenuity than to the lack of a way of incorporating negative
numbers into the theory. Thus, although al-Karajī knew rules like a – (– b) = a + b,
for a and b positive, he evidently had not discovered the rule – a – (– b) = –(a – b).
This prevented him from extending his method for dividing two polynomials to
cover all cases, for his procedure, which we shall explain later, would, in general,
require the subtraction of one negative quantity from another. It also prevented his
discovery of a way to extract square roots of polynomials, for the same reason.
Students who have struggled with the law of signs may find comfort in learning that
at one time the discovery of these rules taxed the ingenuity of the best mathe-
maticians, and that the discovery of much of our elementary (pre-calculus) math-
ematics was a matter of considerable labor and many false starts.
We do, however, find the laws for dealing with signed magnitudes in the
writings of a physician named al-Samaw’al ben Yaḥyā ben Yahūdā al-Maghribī,
who was born in Baghdad, perhaps 70 years after al-Karajī died, and whose work is
a commentary on that of al-Karajī. In his work Al-Bāhir fi’l-Ḥisāb (The Shining
Book on Calculation), which he wrote when he was nineteen, al-Samaw’al states
the more troublesome parts of the rule of signs as follows:
… if we subtract a deficient (negative) number from a deficient number larger than it (i.e.
representing a greater deficiency) there remains the difference, deficient (e.g. –5 – (–2) = –
(5 – 2)), but in the other case there remains their difference, excess (– 2 – (– 5) = + (5 – 2)).
If we subtract an excess number from an empty order (one where there is a zero), the same
number, deficient, remains; but, if we subtract a deficient number from an empty order there
remains in it that number, excess.
6 Al-Karajī’s Arithmetization of Algebra 135

Thus, al-Samaw’al conceived of numbers as expressing either an excess (our


“positive”) or a deficiency (“negative”), so that, for example, a number expressing a
deficiency of 5 would be larger than one expressing a deficiency of 2.
Al-Samaw’al’s rules for subtracting powers, when expressed symbolically, say that
(– axn) – (– bxn) is equal to

ðaxn  bxn Þ; if a\b

and to

þ ðbxn  axn Þ; if a \ b.

With these rules al-Samaw’al could, and did, use the procedures we know today
for adding and subtracting combinations of powers by adding and subtracting like
terms.
Al-Samaw’al was born into a Jewish family, and it is a commentary on the
conditions in Baghdad at that time that he could find no one competent to teach him
mathematics beyond the first few books of Euclid’s Elements. So he completed his
studies of that work by himself and then studied the works of Abū Kāmil and
al-Karajī. In his autobiography he tells of a dream he had in 1163 that made him
convert to Islam. His life was spent as a traveling physician, who counted princes
among his patients, and he died in Maragha—in northern Iran—around 1180.
Of Al-Samaw’al’s 85 recorded works, which range over mathematics and
astronomy, medicine and theology, only a few have survived, and, of the surviving
mathematical works, we shall discuss only The Shining, which we have mentioned
earlier. Our account will deal with two sections from the first part on algebra,
namely those dealing with the law of exponents and dividing expressions composed
of different orders.

6.2 Al-Samaw’al on the Law of Exponents

The basis for al-Samaw’al’s rules for multiplication is the following chart. In
presenting this and other pieces from al-Samaw’al’s work we use, as a compromise
between the modern and medieval notation, abbreviations, such as “mcc” for “māl
cube cube” (x8) and “pcc” for “part of cube cube” (1/x6).
In this way the reader may gain a better feeling for mathematics in the
twelfth-century Muslim world.
To begin, al-Samaw’al sets down a chart aimed at teaching the reader how to
multiply or divide simple expressions, such as pmc and mmc. The columns of this
chart are headed by the usual Arabic alphabetic numerals, so A stands for 1, B for 2,
etc. (Chart 1). The fourth column to the left of the one headed “0,” namely the
1 one

headed “D,” is read: “When ‘thing’ is 2 (resp. 3) then ‘part of māl māl’ is 4 14
11
(resp. 9 9 )).”
136 4 Algebra in the Islamic World

Chart 1

The importance of the chart is that al-Samaw’al uses it to obtain an operational


equivalent of the law of exponents: For any two integers m and n, xmxn = xm+n. He
states the rule as follows, where “distance” means the “number of cells on the chart”
The distance of the order of the product of the two factors from the order of one
of the two factors is equal to the distance of the order of the other factor from the
unit. If the factors are in different directions then we count (the distance) from the
order of the first factor towards the unit; but, if they are in the same direction, we
count away from the unit.
Al-Samaw’al discusses several examples of this rule in which he makes good
use of the numerals above the orders. Thus in the example of pc times pmm he says
that since mm is fifth, starting from the unit, we count, starting at pc and moving
away from the unit, five orders and end up at pmmc; but, then he says
Opposite (above) the order of part of cube is 3 and opposite part of māl māl is 4. We add
them to obtain 7 and opposite (below) it is the order of part of māl māl cube.

Thus it was that a twelfth-century algebraist expressed what we write as


x–3x–4 = x–7. Again, al-Samaw’al writes (where we substitute numerals for his
number-words)
To find 3 parts of māl multiplied by 7 cubes we multiply 3 by 7 and 21 results. We find the
order of “cube” to be fourth from the unit, so we count four orders from the order of “part of
māl” towards the unit, so that the result … is 21 things.… If we want, we take the difference
of the numbers opposite the orders of the factors, namely 2 and 3, and we find 1, and
opposite 1, we find, in the direction of the factor that has the larger number opposite it, the
order of things.”

This latter rule we may write as xnx–m = x(n–m) for n > m.


Al-Samaw’al’s justification for his general rule is that if c = a  b then c:a = b :
1, so that the product, c, is related to a as b is related to 1. In particular, if b is
n orders to the right or left of the order of 1 then c must be n orders to the right or
left of the order of a. This is a nice example of giving a paraphrase of a rigorously
defined mathematical relation and using the paraphrase, interpreted the way one
feels is right, to deduce consequences of the relation. In the hands of one with a sure
feeling for the subject such heuristic reasoning is a fruitful method of discovery, and
so it was with al-Samaw’al.
With the above rules for multiplying the individual orders, al-Samaw’al has no
difficulty explaining how to multiply two expressions, each composed of various
6 Al-Karajī’s Arithmetization of Algebra 137

orders, simply by multiplying each term of one by all terms of the other and adding
up the results.
As for division, he notes that, “The division of a composite expression by a
single order is easy for him who knows the division of the single (by the single),
and the division of the single is easy for him who knows the multiplication of the
single.” (The first, of course, is because (a + b)/c = a/c + b/c, and the second
because a/c = d means a = d  c.) He says, however, that other cases are more
difficult, and that no one up to his time has solved them, but he has found a way and
gives the following example to clarify it.

6.3 Al-Samaw’al on the Division of Polynomials

The First Example


He sets the problem of dividing

20 cc þ 2 mc þ 58 mm þ 75 c þ 125 m þ 96 t þ 94 units
þ 140 pt þ 50 pm þ 90 pc þ 20 pmm

by 2c + 5t + 5 units + 10 pt. In modern symbols this asks for the computation of

20x6 þ 2x5 þ 58x4 þ 75x3 þ 125x2 þ 96x þ 94 þ 140x1 þ 50x2 þ 90x3 þ 20x4
2x3 þ 5x þ 5 þ 10x1

He says that to start, “We put the two terms in natural arrangement, and in each
empty order we put a zero.” (Al-Samaw’al’s procedure is obviously intended to be
used on the dust board, where erasure is easy, but space is at a premium, and it
proceeds by a series of charts. It adapts easily, however, to paper, where erasure is
not easy but space is ample, and nothing is lost if we combine his charts into one
(Chart 2).
The top row lists the names of the orders in their natural sequence from left to
right, and the row below it is the row of the answer, which is initially empty and is
filled in as the process proceeds. The rest of the chart is divided into horizontal
bands, each containing two rows, and each band, together with the two top rows,
constitutes one of al-Samaw’al’s charts. Thus we may speak of the first, second,
etc., charts, and the reader will understand what we mean.
In the first chart are the coefficients of the dividend, each under the name of its
order, and underneath, starting in the left-most column, are the coefficients of the
divisor. Notice that, whereas the names above the terms of the dividend correspond
to the names of the corresponding orders, this is not the case with the divisor, so
you must remember that the divisor begins, in this case, with cubes. Now, says
al-Samaw’al,
138 4 Algebra in the Islamic World

Chart 2

We divide the greatest order of the dividend (20 cc) by the greatest order of the divisor (2c)
and the result, 10c, we place in the order facing 75 (the order of cubes). We next multiply it
(10c) by the divisor, and we subtract the result of its multiplication by each order (of the
dividend) from what is above it.

That is, al-Samaw’al divides 20 cc by 2c to obtain 10c and then subtracts from
the dividend the product of 10c by the divisor. The old dividend is now replaced by
the remainder after this subtraction and al-Samaw’al says, “We copy the divisor one
order to the right, as we do in the Indian style of arithmetic, and obtain the second
chart.”
The reader no doubt recognizes the procedure from the division of two whole
numbers as Kūshyār ibn Labbān explains it, with its key sequence: divide leading
term by leading term, multiply, subtract, and shift to the right, but this time the
method is applied to divide algebraic expressions.
Now the sequence is repeated. In the Chart 2 we divide the leading 2 of the new
dividend by the 2 of the divisor, and the quotient, 1, we place in the column to the
right of the 10, in the row of the answer. Now we can forget which order the 2 of
the divisor represents, for the chart keeps track of it for us. Then “1 times the
divisor” is subtracted from the dividend, and the divisor is again shifted one place to
the right to obtain the third chart.
Three more iterations of the procedure result in the last chart, where clearly the
last row divides the row above it exactly, with a quotient of 2, and thus al-Samaw’al
obtains the quotient of
6 Al-Karajī’s Arithmetization of Algebra 139

20x6 þ 2x5 þ 58x4 þ 75x3 þ 125x2 þ 96x þ 94 þ 140x1 þ 50x2 þ 90x3 þ 20x4

by

2x3 þ 5x þ 5 þ 10x1

as

10x3 þ x2 þ 4x þ 10 þ 8x2 þ 2x3 :

Our present method of division attaches labels (x, x2, etc.) to the coefficients each
time they are written down in order to keep track of what they represent; but,
al-Samaw’al’s method labels the columns of an array once and then arranges the
coefficients in the labeled columns.

The Second Example


He follows the previous example of division by another example, in which he
divides expressions involving negative coefficients and makes good use of his rules
for signed quantities, including a subtraction – 20 – (–40) = 20 of the sort that gave
al-Karajī troubles.
The following, final, example illustrates al-Samaw’al’s insight. The problem is
that of dividing 20 m + 30t by 6 m + 12, and although we shall again give all his
charts combined into one, we will not repeat the explanation of the procedure
(Chart 3).
Al-Samaw’al calls the result so far, namely

1 1 2 1 1 1 1 1 2 1 1
3 þ 5   6  2  10  3 þ 13  4 þ 20  5  26  6  40  7 ;
3 x 3 x x 3 x x 3 x x

“the answer approximately,” and he then checks his work by writing out the
product of this expression by the divisor and subtracting the product from the
dividend to verify that the difference is the remainder 320x–6 + 480x–7.
By this point al-Samaw’al has seen the rule that governs the formation of the
coefficients of the quotient, for without further calculations he now writes down the
coefficients of all succeeding powers of x as far as that of x–28, which he correctly
gives as 54,613 13 (where the fractional parts of the numbers are written out in
words). The rule is, namely, that if a–n is the nth coefficient then an ¼  12 an2 :
However, apart from the insight such calculations show, the discovery of this
procedure of long division, which is in all its computations precisely our
present-day one, is a fine contribution to mathematics, and it seems to be a joint
accomplishment of al-Karajī and al-Samaw’al.
140 4 Algebra in the Islamic World

Chart 3

7 Al-Samaw’al on the Table of Binomial Coefficients

In Chap. 2 we referred to al-Karajī’s discovery of Pascal’s triangle. Since


al-Karajī’s treatise containing his discovery has not been found we shall take our
account of his discovery from al-Samaw’al’s account1 of the generation of the table
of binomial coefficients, where he names al-Karajī as his source. He begins with a
demonstration of the theorem that for any positive integer n; ða  bÞn ¼ an  bn , and
the expansion of ða þ bÞn for n = 3, 4 and 5.2 (In a previous chapter, al-Samaw’al
showed that for any two numbers a and b, a2b2 = (ab)2, so he has already estab-
lished the identity for n = 2.)

1
This is found in Chapter 4 of Section II of his al-Bāhir.
2
Our account relies on that given in Bajri et al.
7 Al-Samaw’al on the Table of Binomial Coefficients 141

Now, his first lemma is as follows:


(For) any four numbers, the product of the surface formed by the first and the second3 by
the surface formed by the third and the fourth is equal to the product of the surface of the
first and the third by the surface of the second and the fourth.

Symbolically one would write: If a, b, g, and d are any four numbers then
(ab)(gd) = (ag)(bd). (The reader may have noted that if we set a = b and g = d the
above result about squares of a product of two terms is an immediate consequence
of this first lemma.)
Al-Samaw’al’s proof uses a number of propositions from Euclid’s Elements, and
it runs as follows,: Set ab = e, ag = z, bd = h and gd = t. I claim that et = zh. By
definition of e and z and Elements VII, 17, e/z = b/g. Similarly h/t = b/g.4 Then, by
Elements VII, 17, e/z = h/t. Hence, by Elements VII, 19 (or VI, 16), et = zh. Thus,
by definition, (ab)(gd) = (ag)(bd), which was to be proved.
His second lemma states that the surface of two sides each cubed is equal [to the
cube] of their surface.5 Again, in modern terminology, (ab)3 = a3b3.
This follows easily from the first lemma and the identity (ab)2 = a2b2 which, as
we said, he had proved earlier in the book.
His third lemma states that if the number ab = ag + gb6 then the cube of ab is
equal to the sum of the cubes of each of the summands plus three times the products
of the square of each summand by the other.
On the basis of what he has proved so far, ab3 ¼ ab  ab2 ¼ ab  ðag2 þ gb2 þ
2ag  gbÞ and, by Elements II,1 this is equal to ag2 ðag þ gbÞ þ gb2 ðgb þ agÞ þ
2ag  gbðag þ gbÞ. Again simplifying each product by use of Elements II, 1 and
using the rule that for any numbers p, q, and r the product (pq)r = ( pr)q the desired
identity follows.7
He now proves lemma 4, which is the rule for expanding (p + q)4, the proof
proceeding in the same way from Lemma 3 that the proof of that lemma proceeded
from the result for the square.
Complementing this result, he shows that for any two numbers p and q,
(pq)4 = p4q4 and he states that, in a similar way the reader can show the corre-
sponding result for the fifth power.

3
He thinks of the product of two numbers, a and b, as the area of a rectangular surface with sides
a and b.
4
Since Euclid has two different theories of proportion in his Elements, one for magnitudes in
general in Book VI and another for numbers in particular it may be that al-Samaw’al had Elements
VI, 1 in mind here.
5
The terminology ‘surface’ for ‘product’ is found in Euclid’s Elements, even though ‘the surface of
a cube and another cube’ or ‘the surface of a surface’ makes no geometric sense.
6
He expresses this as saying that the number ab is divided at the point g. The practice of
representing numbers as line segments is very much in the tradition of Euclid’s Elements. VII – IX.
So a number written as a sum may be represented geometrically as a line segment, ab, divided at a
point, g.
7
Although al-Samaw’al does not prove (pq)r = (pr)q it follows from Lemma 1 if we recall that
r = r1.
142 4 Algebra in the Islamic World

He also states that a


person who understands what we have done. .. can demonstrate that for any number divided
into two parts the fifth power is equal to the fifth power8 of each of the two parts and the
product of each one by the fourth power of the other one taken 5 times and (the product of)
the square of each of them by the cube of the othertaken 10 times. And so on for the higher
powers.

Al-Samaw’al now turns to the task of representing the coefficients in the


expansion of (a + b)n as entries in a table.

Al-Samaw’al describes a triangular grid array in which the cells in the top row
are labeled, from right to left, by the names of the various powers: x, x2,. .., x12. In
each of the two cells directly below the right-most entry in the top row place a 1.
Then, copy the top 1 into the cell directly to the left of it. Then add the first 1 to the
1 below it to obtain 2, and write the 2 below the 1 in the second column. Then
copy the second 1 in the first column into the cell below the 2 in the second column.

8
He refers to the fifth power with the algebraic terminology māl cube and the fourth as mâl mâl,
although in the case of a simple square he uses the geometrical terminology for a square,
murabba’.
7 Al-Samaw’al on the Table of Binomial Coefficients 143

We thus have, reading down the second column from the right, 1 – 2 – 1, and this
completes the second column.
Al-Samaw’al then explains that the entries in the second column tell one that to
square a number expressed at the sum of two numbers, a + b, we multiply each of
the summands by itself once—which is what the top and bottom 1’s mean. And the
2 in the middle means that we multiply each term by the other (i.e., ab and ba)
and, so, (a + b)2 = a2 + b2 + 2ab.
To fill in the third column al-Samaw’al says to copy the 1 at the top of the
second column into the top cell of the third column. Then add the 1 at the top of the
second column to the 2 below it and put the sum (3) in the second cell of the third
column. Then add the 2 in the second column to the 1 below it and place the sum,
again 3, below the 3 in the third column. Finally, copy the final 1 of the second
column into the cell below the second 3 in the third column. One may now interpret
the entries in the third column, as al-Samaw’al explains, as saying that
(a + b)3 = a3 + 3a2b + 3ab2 + b3.
Rather than express the algorithm in general terms, al-Samaw’al proceeds with
the cases of exponents 4 and 5 to conclude, for the fifth column (which, in his
terminology, is the case of māl cube) that (a + b)5 = a5 + 5a4b + 10a3b2 + 10
a2b3 + 5ab4 + b5. He fills in the rest of the table up to the 12th power (i.e., “cube
cube cube cube”).
Clearly al-Karajī (and al-Samaw’al after him) understood the way to generate the
table of binomial coefficients for exponent n recursively from those for exponent
n-1. We shall meet this table again, in a different context, in Chap. 7.

8 Algebra in the Maghrib

In terms of our discussion about Arabic names in Chap. 1, al-Samaw’al had the
nisba “al-Maghribī,” i.e., “the man from the Maghrib.” In his case the nisba most
likely referred to the origins of his family, but certainly, by the time al-Samaw’al
was active, in the mid-twelfth century, the Maghrib had produced a number of
well-known writers on algebra. Among them were Aṣbagh b. al-Samḥ (979–1035),
from Cordova, and Abū Bakr al-Ḥaṣṣār, who was active both in Seville and the
Maghrib.
And, in his Book on Fundamentals and Preliminaries for Algebra9 written early
in his career, the Moroccan mathematician Aḥmad Ibn al-Bannā’ gave, at least for
positive exponents, very much the same rule for multiplying powers of an unknown
that one finds in al-Samaw’al. However, in his Summary, which we discussed in
Chap. 2, he gave a different procedure for dealing with exponents in the context of
multiplying or dividing monomials. In the section of this book that is devoted to

Kitāb al-uṣūl wa’al-muqaddimāt fi’l-jabr wa’al-muqābbala. See Djebbar, A. 1990.


9
144 4 Algebra in the Islamic World

finding unknowns, he explains a method for multiplying algebraic terms that does
not depend on al-Samaw’al’s use of an auxiliary chart.10
He begins by informing his readers that the exponent (‘uss) of the ‘thing’ is one
and continues that each māl contributes two to the exponent, and each cube three.11
To deal with powers beyond these three Ibn al-Bannā’ states that when one mul-
tiplies different terms involving these powers one adds the exponents of the mul-
tiplier and the multiplicand and the sum will be the exponent of the product. So, for
example, a term involving māl māl multiplied by another involving cube cube
would produce a term of exponent 2 + 2 + 3 + 3, i.e., 10. And when one multiplies
a known number by a given species the exponent of the product will be that of the
given species.
He immediately applies this to simplifying equations in which no term is just a
number, e.g., 3x2 þ 2x ¼ x3 þ 5x4 . He says that one should subtract the smallest
exponent (in this case 1) from the exponent of each term and, with what remains,
form an equation corresponding to the original equation.
He concludes with the law of signs: The product of two positives or two neg-
atives is a positive, and the product of a positive by a negative is a negative.

8.1 Ibn al-Bannā’ on Quadratic Equations

As did most writers on algebra, Ibn al-Bannā’ in his Algebra12 discusses the
solution of the canonical six equations of degree at most two. But his approach is to
reduce all the quadratics to equations of the form (x + a)2 = c. In this way he is able
to give a purely algebraic proof of the validity of the procedure for solving
quadratic equations.
Thus for an equation of the type x2 + 10x = 39 he says, “The procedure for this
type comes down to a method that changes its type by reducing it to the third of the
[simple] types we mentioned. For, if one adds to x2 + 10x the square of half the
number of roots, i.e., 25, the root of the result will be equal to the root of the square
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffi
plus half the number of roots i.e. x2 þ 10x þ 25 ¼ x2 þ 5.”13
This is, he says, “clear from what we said in the chapter on multiplication in
Part I of this book, that the square of a number [made up of a sum of two numbers]
is equal to the squares of its two parts plus two times the product of the one by the

10
Chapter 4 of Sect. 2 of Part II.
11
According to M. Souissi, it was Ibn al-Bannā’ who introduced the word ‘uss, whose
non-technical meaning is ‘foundation/basis,’ for ‘exponent.’’
12
For our information on this work we have relied on the edition of the Arabic text and French
translation by Djebbar, 1990.
13
This same equation and the algebraic method of solving it is found in al-Karajī’s book, Al-
Fakhrī, in his discussion of solving the three canonical equations of the second degree. (He refers
to his algebraic method as being that of Diophantos.)
8 Algebra in the Maghrib 145

other.” In other words, when we add (10/2)2 to the left-hand side of the equation
x2 + 10x = 39 we obtain a perfect square, (x + 5)2. And he expresses this by the
verbal equivalent of ((√x)2 + 5)2.
He then points out that the square of half the number of roots will be common to
both sides of the equation, so the equation becomes x2 þ 10x þ 25 ¼ 64, and since
the roots of both sides must be equal (x + 5) = 8. Thus, x = 3.

8.2 Algebraic Notation in the Maghrib

After the work of Ibn al-Bannā’ one finds the widespread use of a shorthand
algebraic notation in the Maghrib, which replaced the purely rhetorical algebra that
was so common in Eastern Islam. One of the earliest appearances was in a work by
Aḥmad ibn al-Qunfūdh, from the North African town of Constantine (in modern
Algeria). This work (Removing the Veil from the Methods of Arithmetic Operations)
is found in a unique fifteenth century manuscript and makes extensive use of
algebraic notation. Since he does not comment on his use of the notation it seems
probable that the use of such notation predates him.
The principle of the notation was to use truncated forms of Arabic words that we
introduced in the previous sections (often truncated to the initial letter) to represent
the unknown and its powers.14 Thus, the Arabic letter shīn (‫)ﺵ‬, the first letter of
shay’ (thing), represents the unknown, mīm (‫)ﻡ‬, the first letter of māl, represents the
square of the unknown, and kāf (‫)ﻙ‬, the first letter of ka‘b (cube) represents the third
power. The square root was represented by the Arabic letter jīm (‫)ﺝ‬, the first letter
of the Arabic word for root (jidhr).
The first five powers of the unknown, then, are

They (as well as the symbol for square root) were written above their coeffi-
cients, as below

14
This method of abbreviating words used frequently in mathematics has been called syncopation,
and can also be found in Byzantine manuscripts of Diophantos’s Arithmetica.
146 4 Algebra in the Islamic World

The symbol for equality, the letter lām (‫)ﻝ‬, appears in both equations above. In
the second equation the second symbol from the right is the Arabic word lā, which,
in an algebraic context, functions as a minus.
The image below shows two equations, of degrees six and nine respectively,
from the work of the fifteenth-century writer Aḥmad al-Qatrawānī, who was born in
Upper Egypt but taught in Tunis. It is the first example we know of an Arabic writer
extracting roots of polynomials using algebraic symbolism rather than the tables we
saw in al-Samaw’al’s work.

The reader may use the information we have given about algebraic symbolism to
read the top polynomial (left to right) as
81x6 þ 72x5 þ 106x4 þ 184x3 þ 89x2 þ 80x þ 64.
In the case of this polynomial, the text shows how to extract its square root.

9 ‛Umar al-Khayyāmī and the Cubic Equation

9.1 The Background to ‛Umar’s Work

An older contemporary of al-Samaw’al was ‛Umar al-Khayyāmī, whose treatment


of cubic equations is found in his book Algebra. He completed this work in
9 ‛Umar al-Khayyāmī and the Cubic Equation 147

Samarqand and dedicated it to the chief judge of that city, Abū Ṭāhir, and in the
preface to this work he refers to his harried existence up to then
I have always desired to investigate all types of theorems …, giving proofs for my dis-
tinctions, because I know how urgently this is needed in the solution of difficult problems.
However, I have not been able to find time to complete this work, or to concentrate my
thoughts on it, hindered as I have been by troublesome obstacles.

When ‛Umar did have the security to concentrate on a problem his powers of
intellect were remarkable. One of his biographers, al-Bayhaqī, tells how ‛Umar read
a book seven times in Isfahan and memorized it. When he returned he wrote it out
from memory, and subsequent comparison with the original revealed very few
discrepancies. However, ‛Umar’s feats of intellect were by no means confined to a
remarkable memory, as we shall see in the sections from his great work, Algebra.
In his introduction to this work ‛Umar mentions that no algebraic treatment of
the problems he is going to discuss has come down from the ancients, but that
among the modern writers Abū ‘Abdallāh al-Māhānī wrote an algebraic analysis of
a lemma Archimedes used in the problem from his work Sphere and Cylinder II,4
which we mentioned earlier, the problem of cutting a sphere by a plane so that the
volumes of the two segments of the sphere are to one another in a given ratio.
Archimedes showed that this problem can be solved if a line segment a can be
divided into two parts b and c so that c is to a given length as a given area is to b2. If
we let b = x, so c = a – x, the proportion may be written as x3 + m = nx2, where
m is the product of the given length and area. Khayyam tells us that neither
al-Māhānī, who lived from 825–888 (and would therefore have been contemporary
with al-Khwārizmī), nor Thābit could solve this equation, but a mathematician of
the next generation, Abū Ja‛far al-Khāzin, did solve it by means of intersecting
conic sections. Then, following Abū Ja‛far, various mathematicians tried to solve
special kinds of these equations involving cubes, but no one had tried to enumerate
all possible equations of this type and solve them all. This ‛Umar says, he will do in
this treatise.

9.2 ‛Umar’s Classification of Cubic Equations

We now give an account of some parts of his treatise Algebra, and we emphasize
that although we shall speak of “equations” and “coefficients” ‛Umar did not write
these symbolically, for he used only words, even for the numbers.
In the first part of his treatise ‛Umar lists all types of equations in which no term of
degree higher than three occurs. In ‛Umar’s equations all terms appear with positive
coefficients so that, whereas we would see x3–3x + 8 = 0 and x3 + 3x – 8 = 0
as being of the same type, ‛Umar viewed them as being different. He would have
expressed the first as “Cube and numbers equal sides” (x3 + 8 = 3x) and would have
seen it as distinct from “Cube and sides equal numbers” (x3 + 3x = 8). Thus, he
arrives at 25 species of equations, and, in the remainder of the treatise, he shows how
these may be solved—11 by Euclidean methods and 14 by conic sections. For each
148 4 Algebra in the Islamic World

of these 14 species ‛Umar gives a short section showing how the conics can be used
to produce a line segment from which solids that satisfy the required relation can be
constructed. The student who reads the English translation by D. S. Kasir will find
‛Umar’s arguments quite clear and will enjoy, as well, the many interesting asides on
the history of various types of equations.

9.3 ‛Umar’s Treatment of x3 + mx = n

Preliminaries
The equation we are going to discuss is “cube and sides equal a number,” that is,
“x3 + mx = n” and, to understand the section that we have chosen to present from
this work, the reader must recall that if ABC is a parabola with vertex B and
parameter p and if x is any abscissa and y the corresponding ordinate then y2 = p  x.
‛Umar begins with a lemma about solid figures called parallelopipeds, solids
with three pairs of parallel faces (Fig. 8a). When all the faces are rectangles
(Fig. 8b), as in the case of a brick, the solid is called a rectangular parallelopiped.
One face of a parallelopiped is arbitrarily designated as its base, and ‛Umar’s
lemma concerns the case in which the base is a square.
Lemma. Given a rectangular parallelopiped ABGDE (Fig. 9), whose base is the
square ABGD = a2 and whose height is c, and given another square MH = b2,
construct on MH a rectangular parallelopiped equal to the given solid.
Solution. Use Euclidean geometry to construct a line segment k so that a:b = b:
k; and then construct h so that a:k = h:c. Then the solid whose base is b2 and whose
height is h is equal to the given solid.
Proof. a:b = b:k implies a2:b2 = (a:b)  (a:b) = (a:b)  (b:k) = a:k, but a:k = h:
c, and so a2: b2 = h:c, and this implies a2c = b2 h. Thus the solid whose base is b2
and whose height is h is equal to the given solid a2  c, and that is what we wanted
to show.
Algebraically this lemma asks for the root of a2  c = b2  x, given a, c and b, and
Khayyām’s constructions obtain this solution (a2  c/b2) by first obtaining k = b2/a
and then h = (ac)/k = a2  c/b2. The principal fact, taken as known by ‛Umar, is that
given any three straight-line segments a, b, c it is possible to find a fourth segment

Fig. 8
9 ‛Umar al-Khayyāmī and the Cubic Equation 149

Fig. 9

Fig. 10

d so that a:b = c.d. The segment d is called “the fourth proportional”, and Fig. 10
shows how d can be constructed.

9.4 The Main Discussion

‛Umar now comes to his first nontrivial equation, which he describes as “Cube and
sides equal a number,” i.e., the case we would write as x3 + mx = n, where m and
n are positive. For this he gives the following procedure: Let b be the side of a
square that is equal to the number of roots, i.e., b2 = m, and let h be the height of the
rectangular parallelopiped whose base is b2 and whose volume is n. (The con-
struction of h follows immediately from the previous lemma.) Now take a parabola
(Fig. 11) whose vertex is B, axis BZ and parameter b, and place h perpendicular to
BZ at B. On h as diameter describe a semicircle and let it cut the parabola at D.
From D drop DE perpendicular to h and the ordinate DZ perpendicular to BZ. Then
DZ = EB and with y = BE it follows that y3 + my = n.
Proof. Let BZ = x. By the properties of the parabola y2 = bx and, by the
properties of the circle, x2 = y(h – y). But the first equality may be written as
x:y = y:b and the second as x:y = (h – y):x. Thus (h – y):x = x:y = y:b, or, inverting,
b:y = y:x = x:(h – y). Hence b2:y2 = (b:y)(b:y) = (y :x)(x:(h – y)) = y :(h – y), and
thus b2  (h – y) = y  y2, i.e. b2  h – b2 y = y3. Therefore, if we add b2 y to both
sides it follows that b2h = y3 + b2  y. If we then substitute for b2  h its equal, n,
150 4 Algebra in the Islamic World

B x Z
b
y

E D
h

Fig. 11

and for b2 its equal, m, we may conclude that y3 + my = n, which was the equation
we wanted to solve.
We can write the argument a bit more briefly as y2 = bx implies y4 = b2x2, and
x = y(h – y) implies b2x2 = b2y(h – y). Thus, y4 = b2x2 = b2y(h – y) and so,
2

because y 6¼ 0,
y3 ¼ b2 ðh  yÞ; i:e: y3 þ my ¼ y3 þ b2  y ¼ b2  h ¼ n

and the equation is solved.

9.5 ‛Umar’s Discussion of the Number of Roots

Throughout the discussion ‛Umar is careful to warn the reader that a particular case
may have more than one solution (or, as we should say, more than one positive real
root) or that it may have no solutions. What happens in any given case depends on
whether the conic sections he is using intersect in none, in one or in two points. For
example, he obtains the solution to x3 + n = mx by intersecting a parabola and
hyperbola and notices that the two curves may not intersect, in which case there
would be no solution, but if they do then they either intersect tangentially or at two
points. In our modern terminology, we would express this by saying the equation
x3 + n = mx either has no positive real solution or two of them. In the latter case,
the two could be a single root, a repeated root, corresponding to a factor (x – a)2, or
pffiffiffi
two different roots. Again, in the case of x3 + n = mx2 he notes that if 3 n  m then
p ffiffi

there is no solution. For, if 3 n  m then
pffiffiffi3 pffiffiffipffiffiffi2 pffiffiffi2
n ¼ 3 n ¼ 3 n 3 n m 3 n : ð1Þ
pffiffiffi
This implies that if x is any solution then x [ 3 n for

x3 þ n ¼ mx2 implies mx2 [ n: ð2Þ


9 ‛Umar al-Khayyāmī and the Cubic Equation 151

Thus, combining (1) and (2) we obtain


pffiffiffi2 p ffiffiffi2 pffiffiffi
mx2 [ n  m 3 n ; so mx2 [ m 3
n ; i:e x [ 3 n:
pffiffiffi pffiffiffi
Thus 3 n  m implies for any solution x, x > 3 n.
pffiffiffi
On the other hand, for any solution, x3 < mx2, so that x \ m \ 3 n, and this
p ffiffi
ffi p ffiffi

contradicts x [ 3 n. Hence ‛Umar has shown if 3 n  m then x3 + n = mx has no
(positive real) solution.
The reader should be aware that we have presented ‛Umar’s argument by means
of our modern symbolic algebra, a product of the European Renaissance. ‛Umar’s
argument uses either geometrical magnitudes or numbers interpreted geometrically,
and the only mathematical symbols he uses are letters to denote points on geo-
metrical diagrams. Thus, for example, in the case of the preceding argument, ‛Umar
pffiffiffi
expresses the condition 3 n  m by saying, “Let AC represent the number (m) of
squares and describe a cube equal to the given number (n), the side of which is h….
If h is equal to AC the problem will be impossible because…”(‛Umar’s pleasure at
having discovered that certain cases that earlier workers had thought impossible are
in fact quite possible is evident at several times in the treatise.)
However, having stressed the geometrical form of ‛Umar’s argument, we must
emphasize also that ‛Umar looked on his work as being a contribution to algebra.
After the customary invocation of Allah and prayers for His blessings on His
prophet Muḥammad he says, “One of the branches of knowledge needed in that
division of philosophy known as mathematics is the science of algebra, which aims
at the determination of numerical and geometrical unknowns….” Further he begins
his first chapter by saying,
Algebra. By the help of God and with His precious assistance I say that algebra is a
scientific art. The objects with which it deals are absolute numbers and (geometrical)
magnitudes which, though themselves unknown, are related to things which are known,
whereby the determination of the unknown quantities is possible… What one searches for
in the algebraic art are the relations that lead from the known to the unknown, to discover
which is the object of algebra as stated above.

If these relations that lead us from known to unknown happen to stem from the
properties of geometric figures the problem is no less algebraic in ‛Umar’s view. It
is the use of the given relations to search for the unknown that is the hallmark of
algebra—nothing else.
In fact, even though ‛Umar’s treatise expresses the solutions as line segments
and not as numbers depending on the coefficients of the equation, we know ‛Umar
wanted to find such numbers for he writes
As for a demonstration of these types, if the object of the problem is an absolute number,
neither we nor any of the algebraists have succeeded, except in the case of the first three
degrees, namely number, thing and square, but maybe those after us will.

‛Umar’s hope was fulfilled, for in the early part of the sixteenth century in Italy a
group of algebraists put together different pieces of the puzzle, and in 1545 the
152 4 Algebra in the Islamic World

physician and astrologer Gerolamo Cardano published in his Ars Magna (The Great
Art), just as ‛Umar had, a case-by-case analysis of the cubic equation, but this time
the roots were expressed not as line segments but as numbers depending on the
coefficients of the equation. Instead of conic sections, Cardano employed identities
like (a – b)3 + 3ab(a – b) = a3 – b3 to show how the solution of the cubic could be
obtained from a solution of an associated quadratic followed by the extraction of
cube roots—both of which could have been done numerically long before ‛Umar’s
time. The formula Cardano published for a root of the equation x3 + px = q is
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
3 q
q2 p3ffi 3 q rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi q2 p3
x¼ þ þ   þ þ
2 2 2 2 2 3

and in this way ‛Umar’s hopes were fulfilled, over four centuries later.

10 The Islamic Dimension: The Algebra of Legacies

To illustrate further mathematical methods in the service of Islam we turn, as in the


chapter on arithmetic, to the latter half of al-Khwārizmī’s Algebra and a problem
from the science of legacies (‛ilm al-waṣāyā). This science, the reader may recall,
requires the application of religious law and algebra to calculating shares of an
estate when a legacy to a stranger is involved.
Al-Khwārizmī gives the following example of the use of algebra in the case that
there is a legacy
A man dies, leaving two sons and bequeathing one-third of his estate to a
stranger. His estate consists of ten dirhams of ready cash and ten dirhams as a claim
against one of the sons, to whom he has loaned the money.
The relevant parts of Islamic inheritance law are:
The natural heirs can only refuse to pay that part of a legacy by which it exceeds one-third
of the estate, so in this case the legacy must be paid;
The amount by which the outstanding loan to the son exceeds what would be his legal share
is treated as a gift to the son; and
The gift precedes the legacy and the legacy precedes the natural shares.

To solve this problem, al-Khwārizmī lets x be the legal share of each son. Since the
gift must be paid first, only the legal share is taken out of the 10 dirham loan and added
to the 10 dirhams of ready cash to form the estate 10 + x. The stranger than gets
one-third of this, and each son gets x, so (10 + x)/3 + 2x = 10 + x. Thus x = (10 + x)/3,
so that 23 x = 3 13. Then, when each side is increased by its half, 23 x becomes x and
3 13 becomes 5. (Recall that many arithmetic texts had separate sections on halving
a number.) Thus the legal share is 5, the stranger gets (10 + 5)/3 = 5, and the gift is 5.
Exercises 153

Fig. 12

Exercises

1. Show that the fourth type of quadratic equation, x2 + n = mx, is the only one of
al-Khwārizmī’s six types that can have two positive solutions.
2. Find a proof similar to the first case in the case of the second figure of Euclid
II,5 (Fig. 3b).
3. With reference to Thābit’s discussion of the second basic form of the quadratic
equation, show that if a2 + q = pa, then p > a.
pffiffiffi pffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffi
4. Make use of Fig. 12 to show that a þ b ¼ a þ b þ 2 ab and draw your
pffiffiffi p ffiffi
ffi p ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
p ffiffiffiffiffi
own diagram to prove that a þ b ¼ a þ b þ 2 ab 
5. Verify that the three quantities in Problem 61 of Abū Kāmil’s book, b, yb and
zb, not only satisfy the condition that their sum is ten, but they also satisfy the
other two conditions of the problem.
6. Suppose f(x,y,…, w), g(x, y,…, w),…, k(x, y,…,w) are homogeneous polyno-
mials and f is of degree 1. If c is a nonzero constant show that the conditions
f(1, Y,…, W) = c, g(1, Y,…, W) =  = k(1, Y,…, W) = 0 are satisfied if and
only if, with x = (d/c)X, y = (d/c) Y, etc., f(x,y,…,w) = d and g(x, y,…, w) =
 = k(x, y,…, w) = 0. Also identify all these functions and constants in Abū
Kāmil’s problem.
7. Show that the root Abū Kāmil found for the equation arising from Problem 61
is positive, but less than 5, while the other root of the equation exceeds 5.
Deduce that the root Abū Kāmil has located is the correct value of b = 10/a.
8. Use al-Samaw’al’s procedure for division to find the quotient of 2x3 – 11x2 –
13x – 5 by 2x – 5.
9. Use al-Samaw’al’s algorithm to compute the entries in the 13th column of the
table of binomial coefficients, i.e., the coefficients in the expansion of (x + y)13.
10. Solve the following problem from al-Khwārizmī. A man has two sons and
leaves 10 dirhams in ready cash and 10 dirhams as a claim against one of the
sons, to whom he has loaned the money. To a stranger he leaves one-fifth of his
estate plus one dirham. Compute the amount each party will receive.
154 4 Algebra in the Islamic World

Bibliography

Abū Kāmil, Shujā‛ b. Aslam. 1966. Algebra (transl. and comm. by M. Levey). Madison, WI:
University of Wisconsin Press.
Bajri, S.J., and Hannah C. Montelle. 2015. “Revisiting Al-Samaw’al’s table of binomial
coeffcients: Greek inspiration, diagrammatic reasoning and mathematical induction”. Archive
for History of Exact Sciences 69(6): 537–576.
Djebbar, A. 1990. Le livre d’algèbre d’Ibn al-Bannā’ (1256–1321), édition critique, traduction
française et analyse. In Mathématiques et mathématiciens du Maghreb médiéval (IXe-XVIe
siècles) : Contribution à l’étude des activités scientifiques de l’Occident musulman. Doctoral
thesis, Université de Nantes.
Djebbar, A. 2005. L’algèbre arabe, genese d’un art. Paris: Vuibert/Adapt.
Gandz, S. 1938. “The algebra of inheritance”. Osiris 5: 319–391.
al-Khayyāmī, ‛Umar. 1931. The Algebra, Transl. and comm. D. S. Kasir as The Algebra of Omar
Khayyam. New York.
al-Khwārizmī, Muḥammad b. Mūsā, 1831. The Algebra of Muhammed ben Musa, Trans. F. Rosen.
London. (Reprinted by Georg Olms Verlag, 1986.).
Rashed, R. 1975. “Récommencements de 1’Algèbre au XIe et XIIe Siècles”. In The cultural
context of medieval learning, ed. J.E. Murdoch, and E.D. Sylla, 33–60. Dordrecht: Reidel.
Sesiano, J. 1977. “Les Méthodes d’analyse indéterminée chez Abū Kāmil”. Centaurus 21(2):
89–105.
Winter, H.J.J., and ‘Arafat, W. 1950. “The Algebra of ‛Umar Khayyam”, Journal of the Royal
Asiatic Society of Bengal. Science, 16:27–70.
Chapter 5
Trigonometry in the Islamic World

1 Ancient Background: The Table of Chords and the Sine

The branch of elementary mathematics whose origins most clearly lie in astronomy
is trigonometry, for there is no trace of this subject until Hellenistic astronomers
devised geometrical models for the motion of the sun, moon, and five known
planets that required calculating the values of certain sides and angles of a triangle
from other, given, ones. Astronomers of ancient India also used the Greek models
and therefore faced the same mathematical problems, and it is the astronomical
handbooks, or commentaries on them, by Greek and Indian authors, that furnish
most of our record of the early history of trigonometry.
Of course, it is a poor problem that inspires only one solution, and the problems
of astronomy, being very good problems indeed, drew forth a wonderful variety of
solutions which ranged from the construction of numerical sequences to the
methods of descriptive geometry. Among these methods were those that we rec-
ognize as being trigonometric, and in order to provide a historical context for the
Islamic contributions we turn to a brief survey of the developments in Greece and
India.
As early as about 150 B.C. Hipparchos of Rhodes had composed an early
version of a trigonometric table for use in his astronomical researches. The most
thorough Greek treatment of trigonometry, however, is contained in The Almagest,
which is, as we have mentioned earlier, an astronomical work written in the early
part of the second century A.D. by the astronomer Ptolemy of Alexandria. The
word “Almagest,” is an Arabic rendering of a Greek word megistē (with the Arabic
definite article al—put in front) meaning “the greatest,” for Greek writers called
Ptolemy’s Mathematical Arrangement by the name The Great Arrangement, and so
important did his work become that later astronomers, perhaps punning on the dual
sense of a Greek word that could mean either “big” or “great,” called it simply “The
Greatest.” Islamic writers transliterated the Greek name into Arabic and in the
twelfth-century European authors Latinized the Arabic to give us “Almagest.”

© Springer Science+Business Media New York 2016 155


J.L. Berggren, Episodes in the Mathematics of Medieval Islam,
DOI 10.1007/978-1-4939-3780-6_5
156 5 Trigonometry in the Islamic World

In The Almagest Ptolemy supplies tables and rules to allow the user to answer,
for each of the sun, moon or five planets then known, the question, “Where will it
be at a given time?” So that the user may understand how these tables and rules
were composed, Ptolemy constructs the underlying geometrical models, whose
parameters he derived from observations by ingenious mathematical methods.
These models were so successful that they formed the basis for scientific astronomy
to the time of Copernicus.
Many of Ptolemy’s mathematical methods depend on a table that he places in
Book I of The Almagest and which he calls “A TABLE OF CHORDS IN A
CIRCLE”. In Fig. 1 we have reproduced the three columns of part of the table. The

left-most column simply lists the arcs of a circle, beginning with an arc of 12 and

proceeding by steps of 12 up to 180°. The next column lists, for each arc h, the
length of the chord that subtends the arc in a circle of radius 60. The length is given,
using sexagesimal fractions, in terms of the 60 units in the radius. The third column
tells the average increase in the chord length per minute of arc, from one entry to
the next, and is used for purposes of linear interpolation. Thus, for example, row 3

reads “Crd 1 12 = 1;34,15 and for each of the next 30′ of arc add 0;1,2,50 to the
chord length.”
A table such as this, together with knowledge of how to use it, was all the
trigonometry that the ancient Hellenistic astronomer had available to solve plane
trigonometric problems. However, this one table was powerful enough so that even
as late as the thirteenth century the Muslim astronomer Naṣīr al-Dīn al-Ṭūsī dis-
cusses its use in his great work called The Transversal Figure.
In Chap. 2 of Part III Naṣīr al-Dīn begins by remarking that since any triangle
ABG may be inscribed in a circle its sides may be regarded as the chords sub-
tending its angles, or, more properly, subtending the arcs opposite these angles
(Fig. 2). He then says that both in astronomical operations and in geometry itself it
is necessary to find some angles and sides of triangles from other given ones, and
this may be done either by arcs and chords or arcs and sines. Our immediate
concern is with his use of arcs and chords to solve right triangles. (We have solved a
triangle when we have obtained the values of all its sides and angles.)
Naṣīr al-Dīn remarks, first of all, that if only one acute angle of a right triangle
ABG with the right angle at B (Fig. 3) is known then all the angles are known, for
the two acute angles must together total 90°; however, only the ratios of the sides
will be known. Here he is doubtless referrring to Prop. VI,4 of Euclid’s Elements,
which says that if two triangles have equal angles then the sides about equal angles
are proportional.
Thus one must know at least one side to solve the triangle, and Naṣīr al-Dīn first
assumes he knows two (say a, g or a, b in Fig. 3), so he can calculate the third from
the Pythagorean theorem. To calculate the magnitude of an acute angle in this case
imagine a new unit, u, so that b = 120u, and recalculate the sides to this scale. (This
change of scale to a diameter of 120 units runs all through ancient trigonometry,
because the tables of chords and sines were constructed on the assumption of this
diameter.) If a = n  u then one may look down the column headed “Chords” in a
1 Ancient Background: The Table of Chords and the Sine 157

Fig. 1 TABLE OF CHORDS. Taken from Toomer: Ptolemy’s Almagest, copyright by


Springer-Verlag 1984, reproduced by permission of Springer-Verlag
158 5 Trigonometry in the Islamic World

Fig. 2

Fig. 3

table of chords in a circle to find the entry n and the entry next to it will be the
_
magnitude of BG.
It was a well-known theorem from Euclid’s Elements (Book III, Prop.20) that an
angle inscribed in a circle is half the central angle subtending the same arc. From
_
this, ^A ¼ 12 BG, and as before, ^G can be calculated since ^A þ ^G ¼ 90 .
1 Ancient Background: The Table of Chords and the Sine 159

In the third case, when one is given only one side one must also know an acute
angle A. Then ^B may be calculated as 90° – ^A, so that all angles are known.
Again, it is only changing scale to assume the hypotenuse has 120 units and set up
the .proportion involving the known side and either unknown side:
_ _
GB:GA = Crd(GBÞ:Crd(GAÞ:

All arcs are known since, according to the theorem of Euclid quoted in the
second case, above, each arc is twice the opposite angle, which is known. And one
of the two sides is known. Thus we may calculate the other side from the above
proportion, and this concludes Naṣīr al-Dīn’s presentation of how to solve any right
triangle by means of a table of chords.
Thus a table of chords in a circle may be used in place of a sine table to solve
triangles. Indeed, since in Fig. 4,
_ _
2 Crd(2ABÞ 2 Crd(AEÞ
1 1
AG
sin h = = = ;
AO 60 60
_
a table of chords for AE from 0° to 180° is equivalent to a table of sines for h from 0°
to 90°. In addition, relations like cos(h) = sin(90° – h) and tan h = sin h/cos h show
how all trigonometric functions can be expressed in terms of the sine function, so a
single table of chords can be made to serve for all the trigonometric functions. Thus,
with such a table we can determine three unknown quantities in a triangle from three
known quantities, whenever this is possible with trigonometric functions.

Fig. 4
160 5 Trigonometry in the Islamic World

Of course, what is possible in theory may be awkward in practice, for one


function is usually no substitute for six, each one of which is tailor-made for a
certain kind of application. In addition, as we saw in the example, there is often the
inconvenience of having to double an angle in order to calculate the size of an arc
from that of an inscribed angle. However, so far as we know, these inconveniences
never motivated the Hellenistic astronomers to abandon a tool that was used as
early as the second century B.C. by Hipparchos.
It is, rather, the astronomers of India who introduced the function that is basic to
modern trigonometry, namely the sine, and in an Indian astronomical handbook
dating from the fourth or fifth centuries of our era, and known as the Surya

Siddhanta, the values of this function are given in Sanskrit verse for every 3 34 of

arc, from 3 34 to 90°. The radius of the reference circle was taken to be 3438′, where
1
the minute is a unit of length equal to 60 of the length of 1° of arc on the circle. With
this Hindu achievement we have come near to the beginnings of the Islamic
reception of Indian astronomy, for it was in the eighth century, at about the time
Islam had spread from Spain to China, that Indian astronomical texts were trans-
lated into Arabic at the court of the caliph al-Manṣūr in Baghdad.
Soon Muslim astronomers were producing their own astronomical handbooks,
based on Indian and Greek models but incorporating elements of particular concern
to Islamic civilization, e.g., first visibility of the crescent of the new moon and the
direction of Mecca. Such handbooks, written by almost all important Islamic
astronomers, were called zījes, a Persian word taken over into Arabic. The word
originally signified a “thread” or “chord,” and then a set of these, as in the warp of a
fabric. By analogy, astronomical tables, presenting the appearance of a whole set of
parallel lines separating the columns, came to be known by the same word.

2 The Introduction of the Six Trigonometric Functions

Among the ways in which authors of the Islamic world extended the ancient
methods in trigonometry was to define and use all six trigonometric functions, as
follows:
_
(1) The Sine (Fig. 4). This was defined for an arc AB of a circle with center O and
radius R as the length of the perpendicular AG from A onto OB. Clearly, this
is also the length of the perpendicular BD from B onto OA, and its length
_
depends on R. In fact, if SinR AB denotes the sine of an arc AB in a circle of
radius R, then this medieval sine function is related to the modern function by
_
the rule SinR AB = R  sin AB, and to Ptolemy’s chord function as SinR
_ _
AB = 12 CrdR 2 AB. (We shall use capital letters to denote the ancient and
medieval functions and small letters for the modern trigonometric functions.)
(2) The Cosine. Again, Muslim authors described this as a length and not a ratio.
2 The Introduction of the Six Trigonometric Functions 161

_ _
If CosR AB denotes this function for arcs AB < 90° in a circle of radius R then
_  _ 
CosR AB ¼ SinR 90 AB ;

and this is the length OG in Fig. 4. The function was always called “the Sine
of the complement of the arc” and was not tabulated separately.
(3) and (4) The Tangent and Cotangent. Both were originally conceived of as
lengths of certain shadows, the tangent being the shadow of a horizontal rod
mounted on a wall (for a given altitude of the sun) and the cotangent the
shadow of a vertical rod (gnōmōn in Greek and miqyās in Arabic) of standard
length. Thus, in Figs. 5 and 6, the shadow lengths are, respectively, R  tan h
and R  cot h, where R is the length of the rod and h the angle of elevation of
the sun over the horizon.
However, by the tenth century both functions were defined, as we find them in
Naṣīr al-Dīn’s work, according to Fig. 7. Here BD and AG are perpendicular to OB
and EK is perpendicular to EO. Then

Fig. 5

Fig. 6
162 5 Trigonometry in the Islamic World

Fig. 7

_ _
TanR AB ¼ DB and CotR AB ¼ EK:

Muslim authors tabulated the Tangent function, but since they recognized that
the Cotangent is just the Tangent of the complement they did not tabulate it sep-
arately. Beyond the preceding relation, Naṣīr al-Dīn states the following pair of
relations:
_ _ _
1. Tan AB=R ¼ Sin AB=Cos AB. (Notice that when R = 1 all the functions
become their modern counterparts and the relation is the familiar
_ _ _
tan AB ¼ Sin AB=Cos AB.)
_ _ _ _
2. Tan AB=R ¼ R=Cot AB. (When R = 1 this becomes tan AB ¼ 1=Cot AB, a
relation familiar to us.)
(5) and (6) The Secant and Cosecant. These functions were seldom tabulated,
_
but Naṣīral-Dīn defines them, with reference to Fig. 7, as Sec AB = OD and
_
Csc AB = KO. In the terminology of the Arabic writers these were called “the
hypotenuse of the shadow” and “the hypotenuse of the reversed shadow”
respectively, and these names may be explained by reference to Figs. 5 and 6
and the fact that “hypotenuse” refers to the line joining the gnomon’s tip to the
end of the shadow. The Secant is the hypotenuse in Fig. 5 and the Cosecant is
the hypotenuse in Fig. 6. Naṣīr al-Dīn observes that, since the triangles DBO
and AGO are similar, DB/DO = GA/AO, so
_ _
Tan AB Sin AB
_ ¼ :
Sec AB R
3 The Seventh Trigonometric Function 163

Fig. 8

3 The Seventh Trigonometric Function

In addition to these six standard trigonometric functions there was a seventh


function, known as the versed sine (now abbreviated to ‘vers’), which could be
calculated for a given arc by subtracting the cosine of the arc from the radius of the
circle containing the arc (Fig. 8). As were the other six functions, it was a length,
not a ratio, and its value depended, therefore, on the radius of the circle used to
calculate the trigonometric functions. (We will follow our practice with the sine
function and write ‘Vers’ when we are referring to the medieval function.)
The Vers was tabulated as early as the fifth century C.E. in India, where it seems
to have arisen and where one of its names was ‘reversed Sine.’ Another name was
“arrow,” which, one understands by imagining (in the figure above) double the
arcBG as a bow and double the Sine of arc BG as its bowstring. One easily sees the
Versine as the arrow of that bow.1 The function, together with that name, went over
into Arabic both as the Arabic equivalent of “arrow,” sahm, or, also, “reversed
Sine,” jaib mankūs. And both versions of the name went over into Latin, as sagitta
and sinus versus respectively.

The Arabic word qaus can mean either ‘arc’ or ‘bow’, depending on the context.
1
164 5 Trigonometry in the Islamic World

In Arabic the Vers appears in the Astronomical Tables of al-Khwārizmī,2 whose


contributions to arithmetic and algebra we have also discussed in Chapters 2 and 4.
Al-Khwārizmī specifies, for example, that for an arc of less than 90° “it [the arc] is
to be subtracted from ninety and the Sine of the remainder is to be taken, and then
subtracted from 60.” (And it says something of the persistence of tradition that
al-Khwārizmī, who introduced the decimal place-value system into Islamic science,
wrote the entries in his Astronomical Tables using the sexagesimal place-value
system, since it was the only system at that time that could handle fractions in the
same way it handled whole numbers.)
We now jump forward from al-Khwārizmī 400 years and follow the treatment of
the Vers and one of its applications in the work of the Moroccan astronomer, Abū
al-Ḥasan al-Marrakūshī (d. 1262),3 titled Collection of Beginnings and Results in
the Science of Timekeeping.
As the title suggests, al-Marrakūshī begins at the beginning, and in Chapter 10 of
Book I, he begins his treatment of trigonometry with definitions of the chord, Sine,
Cosine and Versed sine of an arc, and how one may find the arc of a circle
corresponding to any of those four quantities. Of the Versed sine he says, “The
arrow is a perpendicular drawn from the end of the arc onto the Sine of the arc,” and
he points out that the Sine, arrow and chord of an arc form a right triangle.
Moreover, he remarks (because of the similarity of triangles) that the product of the
Vers of an arc by the diameter of the circle is equal to the product of the chord of
the arc by itself, and that the product of the Vers of an arc by the other part of the
diameter containing it is equal to the square of the Sine of the arc. Both of these
rules he describes as ‘very useful.’
After this introduction and an exposition of Ptolemy’s rules4 for computing the
chords of 72°, 60°, 12°, 6°, 3°, 1 ½° and ¾° he then presents a table of Sin and Vers
of arcs of a circle progressing by steps of ¼° from ¼° to 90° calculated to 3
sexagesimal places. We show here the last few values of the table, and the reader
will note that Sin90° = Vers90° = 60, since the two are(perpendicular) radii of a
circle of radius 60.
Following this, al-Marrakūshī gives a table of the arcSin, and just before the end
of the chapter says:
We have also prepared, for the artisan, a small table, which gives the approximate value of the
Sine of an arc or of the arc corresponding to a value of the Sine. Its simple and quick use may
substitute for the use of tables, as we ourselves have proved in many graphical operations,5
where the difference between the true value and the approximate value is not noticeable.

2
Adelard of Bath, in the early twelfth century. translated into Latin the reworking of these tables
and rules for their use by the Cordovan astronomer Maslama of Madrid, who died in 1070 A.D.
We have used the English translation of that text in Neugebauer (1962).
3
He was, therefore, roughly a contemporary of another great mathematician and astronomer, Naṣīr
al-Dīn al Ṭūsī, whose work we discuss elsewhere in this chapter and who worked at the eastern end
of the Islamic world.
4
For an account of Ptolemy’s procedure see Aaboe, A. 1997.
5
This probably refers to geometrical constructions rather than by calculations using tables.
3 The Seventh Trigonometric Function 165

The diagram below shows the small table with its nine columns, the first column
being for the first 30 degrees of arc.

Parts of the Quadrant Max. 30 20 10 5 5 5 5 5 5 90


Corresponding parts of the total sine, 30 16 6 2 1/2 2 1 1/2 1 2/3 1/3 60

The entry in the left-most column of the second row says that value of the Sine
of an arc of at most 30 degrees (when the radius has 60 parts) is the same as the
number of degrees. The next column to the right says that after 30° the next 20
degrees correspond to an additional 16 parts. And the following column says that
after 50° the next 20 degrees correspond to an additional 6 parts. Thus, the
right-most column says that the 5° from 85 to 90 contribute an additional one-third
part. For finer divisions the user—who would have been well trained in mental
arithmetic—would, of course, rely on linear interpolation, which we have discussed
in Chap. 2.6
Also in Book I, al-Marrakūshī applies the Vers to solve a problem connected
with calculations needed to construct a planishperic astrolabe. We shall explain this
connection in Chap. 6, where we discuss the astrolabe, but here we—as did
al-Marrakūshī in his Chap 867—shall treat it purely as a mathematical problem,
which is as follows:
Let ABGD be a circle with perpendicular diameters AB and DG passing through
_
the center, E. (Fig. 9) Let AW be a known arc on the quadrant AG, and draw lines
AW and BW. Let AW extended meet DG extended at Z and let WB meet DG at H.
Assuming that the radius of the circle is 60 find the lengths EH and EZ.

Fig. 9

6
The difference between the value of a Sine shown this short table and the value in a modern sine
table (normalized to a unit radius) is never more than .002.
7
See Sédillot and Sédillot, pp. 347–348.
166 5 Trigonometry in the Islamic World

(Although al-Marrakūshī does not mention the line, we have drawn the line WI,
_ _
perpendicular to AB. By definition WI = Sin(AW) and AI = Vers(AW).)
_ _
To find EG, multiply Sin(AW) by 60 and divide the product by Vers(AW).
_ _
To find EH, multiply Vers(AW) by 60 and divide the product by Sin(AW).
The reader is invited to finish the proof.
Al-Marrakūshī gives no proof of these rules but they follow from the facts that
AEZ and AWB are similar right triangles, and both are similar to triangle AIW. He
_
does work an example, however, for the case when AW = 30°, and, in fact,
_
includes a table of the values for AW ranging, by steps of 1°, from 1° to 180°.
The completion of the ancient systems of trigonometry to one based on the six
functions we now use, as well as the Vers, made trigonometry much simpler, and
therefore more useful, than it had been before Islamic times.

4 Abū a1-Wafā”s Proof of the Addition Theorem for Sines

We have seen from the above discussion that Islamic astronomers were aware of
seven trigonometric functions, which are, in every case, constant multiples of the
modern ones. Also, from the late tenth century onward, following the work of Abū
al-Wafā’, they were aware of the possibility of taking R = 1 in defining the Sine,
Cosine and Tangent. Thus Abū al-Wafā’ may be regarded as the first to have
calculated the modern trigonometric functions, and the simplifications such a step
yields may be seen by comparing two statements, each equivalent to the familiar
addition theorem for sines:

sinða  bÞ ¼ sin a  cos b  cos a  sin b:

The first is an ancient form of this law from Ptolemy’s Almagest, where Ptolemy
_ _ _
gives rules (see Fig. 10) for finding (1) Crd(AB – AC), from Crd(AB) and Crd
_ _ _ _ _
(AC), and (2) Crd(AC + CB) from Crd(AC) and Crd(CB). These may be repre-
sented as
_ _  _  _ 
ð1Þ Crd AB  AC  Crd(180 Þ ¼ Crd(ABÞ  Crd 180  AC
_  _ 
 Crd AC  Crd 180  AB ;and
 _   _   _ 
ð2Þ Crd 180  AB  Crd(180 Þ ¼ Crd 180  AC  Crd 180  CB
_ _
 Crd AC  Crd CB .
4 Abū a1-Wafā”s Proof of the Addition Theorem for Sines 167

Fig. 10

_ _
In case (1), Crd(180°) = 120, Crd(AB), and Crd(AC) are all known and (see
_ _
Fig. 10) Crd(180° – AC) and Crd(180° – AB) may both be calculated from these
and the Pythagorean theorem since:
 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
_  _
Crd 180  AC ¼ ðCrdð180 ÞÞ2 ðCrdðACÞÞ2 ;

and
 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
_  _
Crd 180 AB ¼ ðCrd(180 ÞÞ2 ðCrd(ABÞÞ2 :

_ _
Hence five of the six terms in (1) are known so we may solve for Crd(AB – AC).
_ _
In Case (2) one solves for Crd(180° – AB) and from it calculates Crd(AB) by the
Pythagorean theorem. Thus in both cases the desired quantities may be calculated,
but with some effort.
Compare statements (1) and (2), which are not easy to remember, with the
following elegant statement and unified proofs found in Abū al-Wafā”s Zīj
Almajisṭī.
Calculation of the sine of the sum of two arcs and the sine of their difference when each of
them is known. Multiply the sine of each of them by the cosine of the other, expressed in
sixtieths, and we add the two products if we want the sine of the sum of the two arcs, but
take the difference if we want the sine of their difference.
168 5 Trigonometry in the Islamic World

For this statement, which may be expressed by the modern formula given earlier,
Abū al-Wafā’ gives the following proof, which refers to Figs. 11 and 12. (Since he
takes the radius of the circle to be unity we may write his trigonometric functions
without capital letters. The reference to the trigonometric functions “expressed in
sixtieths” does not mean the radius is 60 but only that his tables use sexagesimal
fractions.)
Let there be two arcs AB, BC of a circle ABCD, and suppose we know the sine of
each arc. I say that the sine of their sum, as well as that of their difference, is known.
Join the three points A, B, C to the center O and from B drop BT, BH, perpendicular
to the radii OA, OC, respectively, and then draw HT. In addition, prolong BH, BT to
meet the circle at D, Z respectively. Since radii that are perpendicular to chords bisect
these chords, BH = HD and BT = TZ. Thus triangles BHT, BDZ are similar and so
_ _ _ _ _ _
DZ = 2TH. In Fig. 11, ZBD = 2 AC because ZB = 2 AB and BD = 2 BC and in
_ _ _ _ _
Fig. 12, DZ = 2 AC, for the same reasons and the fact that AB – BC = AC. Thus
_ _
TH = 12 DZ = 12 Crd(2 AC), so TH = Sin AC. To complete the preliminaries draw
BN perpendicular to TH.
The key to Abū al-Wafā”s proof is his observation that since the angles BTO,
BHO are right the four points B, T, H, and O lie on the circumference of a circle of
diameter BO, by Euclid’s Elements III,31. (We have drawn this circle with dashed
lines in both Figs. 11 and 12.) In both cases (Figs. 11 and 12) the angles BHT, BOT
subtend the same chord. In the first case (Fig. 11) they are on the same side of the
chord BT so they are equal. In the second case (Fig. 12) they are on opposite sides

C D

B
H

A O
T

Fig. 11
4 Abū a1-Wafā”s Proof of the Addition Theorem for Sines 169

N
C
H

A
T O

Fig. 12

of the chord BT, so they are supplementary. In this case the supplement of ^ BHT,
namely ^ BHN, equals ^ BOT. In both cases, then, the triangles BHN, BOT are
right triangles with equal angles at H, O so the triangles are similar and thus
_ _
BH/HN = BO/OT. Since BH = sin(BC), OT = cos AB and BO = 1 it follows that
_ _
HN = sin BC  cos AB. In addition, the triangles BNT, BHO are similar because the
angles N and H are right while the angles T and O are equal, being constructed on
the chord BH in the circle through B, H, T and O. Thus

BT BO _ _
¼ where BT ¼ sin(ABÞ; BO ¼ 1 and OH ¼ cosðBCÞ:
TN OH0
_ _
Thus TN = sin(AB)  cos(BC). Finally, in the case of Fig. 11,
_ _  _
sin AB þ BC ¼ sinðACÞ ¼ TH ¼ TN + NH
_ _ _ _
¼ sinðABÞ  cosðACÞ + sinðBCÞ  cosðABÞ:

In the case of Fig. 12,


_ _  _
sin AB  BC ¼ sinðACÞ ¼ TH ¼ TN  NH
_ _ _ _
¼ sinðABÞ  cosðBCÞ  sinðBCÞ  cosðABÞ:
170 5 Trigonometry in the Islamic World

Fig. 13

5 Naṣīr al-Dīn’s Proof of the Sine Law

Naṣīr al-Dīn introduces the Sine Law for plane triangles to provide a basic tool for
solving them, and in this section we shall see how he proves the law and how he
applies it to find unknown parts of triangles from known ones.
The Sine Law. If ABC is any triangle then c/b = Sin C/Sin B.
Figure 13 illustrates the case when one of the angles B or C is obtuse, and
Fig. 14 the case when neither B nor C is obtuse, so that one of them is acute. In
either case prolong CA to D and BA to T so each is 60 units long and, with centers
B, C, draw the circular arcs TH and DE. If we now drop perpendiculars TK and DF
to the base BC, extended if necessary, then TK = Sin B and DF = Sin C. (In the
case of Fig. 13 both of these statements are obvious, but in the case of Fig. 14 the
reader must remember that Sin(^ B) = Sin(180° – ^ B).) Now draw AL per-
pendicular to BC. Since triangles ABL, TBK are similar AB/AL = TB/TK, and
since triangles ACL and DCF are similar, AL/AC = DF/DC but DC = 60 = TB, so,
if we multiply the left and right sides, respectively, of these two proportions, we

Fig. 14
5 Naṣīr al-Dīn’s Proof of the Sine Law 171

obtain the proportion AB/AC = DF/TK. Therefore c/b = Sin C/Sin B, and this
proves the Sine Theorem.
Since Naṣīr al-Dīn’s sine function is simply 60 times the modern one, the above
theorem holds for the modern function as well. We may rewrite the theorem as
c/sin C = b/sin B = a/sin A, a form it is often given in today, and it may be most
easily remembered as the statement that in a given triangle the ratio of any side to
the sine of the opposite angle is constant.
Naṣīr al-Dīn uses this theorem to solve all possible triangles systematically, as
follows.
Case 1. Two angles and one side known:
If two angles (A and B) are known then the other angle C = 180° – (A + B) is
also known (Fig. 15), but at least one side must be given as well since, as we said
earlier, a triangle cannot be determined from its angles alone. Since all angles are
known, however, we may suppose, without loss of generality, that the known side
is c. Then, by the Sine Theorem,

c Sin C c Sin C
¼ and ¼ :
b Sin B a Sin A

In each proportion three out of the four terms are known and so the remaining
terms, a and b, may be found.
Case 2. Angle and two sides known:
If only one angle is known then two sides must be known. If one of these is
opposite the known angle then without loss of generality, the known are c, C and a.

Fig. 15
172 5 Trigonometry in the Islamic World

Fig. 16

Then A is determined by c/a = Sin C/Sin A, and, since two angles are now known,
we are back to the previous case, which we have already shown how to solve.
If, on the other hand, neither side is opposite the known angle then, without loss
of generality, the known are B, a, c (Fig. 16). In this case consider the perpen-
dicular AE from A onto a. Then in the right-angled triangle BEA the side c and the
angle B are known and the side AE may be found as described in the section on
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
solving right triangles. Then EB = BA2  AE2 and CE = a – EB, so in the right
triangle AEC two sides AE and EC are known and the remaining side AC (= b) and
the angle C may be calculated. Then A = 180° – (B + C) and all parts of triangle
ABC are determined.
Case 3. Three sides known:
If no angles are known then the three sides a, b and c are given. In this case
(Fig. 16 also) Naṣīr al-Dīn says to calculate the perpendicular, h, from A onto
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
a according to BE = (c2 + a2 + b2)/2a, and then h = c2  BE2 . (In modern terms
the first equality follows from the Law of Cosines, but in the ancient and medieval
world it would have been viewed as a consequence of Elements II, 12 and 13.)
Naṣīr al-Dīn calls this “the usual rule” for calculating the perpendicular.
Since EC = BC – BE, in the two right triangles BEA, CEA the three sides of each
are known. His solutions of right triangles show how to find the angles of
5 Naṣīr al-Dīn’s Proof of the Sine Law 173

such triangles, so B and C may be calculated as in that section and, from them,
A = 180° – (B + C).

6 Al-Bīrūnī’s Measurement of the Earth

An elegant application of elementary trigonometry was done by al-Bīrūnī when he


was traveling with King Maḥmūd of Ghazna in what is now northwest India, but
was in al-Bīrūnī’s time known as al-Hind. In this section we shall follow his
account of the method as he gives it in his great work on mathematical geography,
On the Determination of the Coordinates of Cities.
He begins with an account of a method Ptolemy gives in his Geography, which
requires a geodetic survey to determine the distance along a great circle connecting
two places of known latitude and longitude. Ptolemy’s method is, in fact, a gen-
eralization of that of Eratosthenes, who measured along a meridian. Al-Bīrūnī
introduces his method with his characteristic dry humor, “Here is another method
for the determination of the circumference of the earth. It does not require walking
in deserts.”
Al-Bīrūnī states that the astronomer Sanad ibn ‘Alī was with the caliph
al-Ma’mūn on one of his campaigns against the Byzantines and used this method
when they came to a high mountain near the sea. Since the method assumes one
knows how to determine the height of the mountain, al-Bīrūnī first explains how to
do that. The problem is nontrivial since a mountain is not a pole and therefore we
cannot easily measure the distance from us to the point within the mountain where
the perpendicular from its summit hits ground level.
To measure the height of a mountain al-Bīrūnī first requires that we prepare a
square board ABGD whose side AB is ruled into equal divisions and which has pegs at
the corners B, G. Then, at D, we must set a ruler, ruled with the same divisions as the
edge AB and free to rotate around D. It should be as long as the diagonal of the square.
Set the apparatus as in Fig. 17 so that the board is perpendicular to the ground and the
line of sight from G to B just touches the summit of the mountain. Fix the board there

Fig. 17
174 5 Trigonometry in the Islamic World

and let H be the foot of the perpendicular from D. Also, rotate the ruler around D until,
looking along it, we sight the mountain peak along the ruler’s edge DT. Now since AD
is parallel to EG, ^ ADT = ^ DEG, and therefore the right triangles ADT and GED
are similar. Thus TA: AD = DG: GE, and since of the four quantities in this pro-
portion only GE is unknown we may solve for GE = AD (DG/TA). However, since
both ^ EGZ + ^ DGH and ^ EGZ + ^ GEZ are equal to right angles it follows that
^ DGH = ^ GEZ, and thus the two right triangles DGH and GEZ are similar, so
that GE:EZ = DG:GH. This means that we may solve for the single unknown
EZ = GE (GH/DG), which is the desired height.
Adjusting the ruler through such a small angle as ADT is going to present
problems, but al-Bīrūnī tells us he used the method to obtain reasonable results, so
we shall take his word for it that the method is not totally impractical.
In fact, he says that “When I happened to be living in the fort of Nandana in the
land of India, I observed from an adjacent high mountain standing west of the fort a
large plain lying south of the mountain. It occurred to me I should examine this
method there.” He is referring to the method of finding the circumference of the
earth, and it is the following, illustrated in Fig. 18.

Fig. 18
6 Al-Bīrūnī’s Measurement of the Earth 175

Let KL be the radius of the earth and EL the height of the mountain. Let ABGD
be a large ring whose edge is graduated in degrees and minutes, and let ZEH be a
rotatable ruler, along which one can sight, which runs through E, the center of the
ring. An astrolabe, which we describe in the next chapter, would be perfectly
suitable for this, using the ruler and scale of degrees on the back of this instrument.
Now move the ruler from a horizontal position BED until you can see the
horizon, at T, along it. The angle BEZ is called the dip angle, d. From L on the
earth, imagine LO drawn so LO is tangent to the earth at L. By the law of sines
applied to D(ELO)

EL : LO = Sin(O):Sin(E) = Sin d:Sin(90  dÞ:

Since the two angles, as well as EL, the height of the mountain, are known, we
may determine LO; but, TO = LO, since both are tangents to a circle from a point O
outside it. Also, since EL and LO are known, it follows from the Pythagorean
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
theorem that EO = EL2  LO2 is known, and hence ET = EO + OT is known.
Again, by the Law of Sines, ET:KT = Sin(d): Sin(90° – d). Since KT, the radius of
the earth, is the only unknown quantity in this proportion, we may solve for KT and
so find the radius of the earth.
As al-Bīrūnī said, he tried the method on a mountain near Nandana in India
where the height, EL, was 652;3,18 cubits and the dip angle was 34′. (Note the very
small angles and the rather optimistically accurate height.) These give for the radius
of the earth 12,803,337;2,9 cubits. Al-Bīrūnī takes the value 3 17 for p and arrives at
the value 80,478,118;30,39 cubits for the circumference of the earth. He than
divides this by the 360° of the meridian circle, and then divides the quotient by the
4,000 cubits in a mile to obtain the value of 55;53,15 miles/degree on a meridian of
the earth. This, he notes, is “very close” to the value of 56 miles/degree determined
by a geodetic survey undertaken at the time of al-Ma’mūn. It doubtless gladdened
al-Bīrūnī’s heart to show that a simple mathematical argument combined with a
measurement could do as well as two teams of surveyors tramping about in the
desert.

7 Trigonometric Tables: Calculation and Interpolation

The successful application of the rules that Naṣīr al-Dīn gives for finding the
unknown parts of triangles from known parts depends not only on knowing the
relevant theorems but on having good trigonometric tables and knowing how to use
them. Accurate tables were necessary, not only to advance the study of such sci-
ences as astronomy and geography but also to investigate such questions as the ratio
of the circumference of a circle to it s diameter. In the calculation of such tables,
Islamic scientists went far beyond their ancient predecessors, and the chart below
shows the increasing precision in the construction of trigonometric tables, as found
176 5 Trigonometry in the Islamic World

in the zījes of some of the major scientists. (The fourth line of the following table,
for example, should be read as saying that around the year 1030 in his zīj al-Bīrūnī
tabulated both the sine and tangent functions, the former in steps of 15′ and the
latter in steps of 1°. In both cases the results were accurate to four sexagesimal
places.)

Year Individual Function(s) Intervals Places


850 Ḥabash al-Ḥāsib Sin, Tan 1° 3
900 Abū Abdullah al-Battanī Sin 1 3
2 °
1000 Kūshyār ibn Labbān Sin, Tan l′, l° 3
1030 al-Bīrūnī Sin, Tan 15′, 1° 4
1440 Ulūgh Beg Sin, Tan 1′ 5

Consider what the calculation of a set of sine tables such as those of Ulūgh Beg
involved. First, 60 entries must be done for each of 90 degrees, so a total of 5400
entries must be calculated. Moreover, the table will be only as accurate as the basic
entries from which the others are calculated, for, from the value of Sin(l°) and from
those of a few other basic Sines, the trigonometric formulae already discussed will
yield a table of Sines for all integer values of n°. The half-angle formulae may then
 
be used to fill in the intervals of 12 or 14 and for divisions finer than this some sort of
interpolation would have been used, which, of course, still demands computation.
Bear in mind, also, that the tables often came provided with auxiliary columns,
recording the increment or decrement from one line to the next to aid the user in
performing interpolations, and you will understand the amount of labor required, in
a pre-calculator era, to produce such a set of tables.

8 Auxiliary Functions

The application of trigonometry to construct astronomical tables often requires the


repeated calculation of the same combination of trigonometric functions, but for
various values of the arguments. For example, in calculating tables of solutions to
problems in spherical astronomy expressions that frequently occur are
 
Tan h  Sin e Tan e  Sin h
and arc Tan ;
R R

where e = 23 12 (approximately), R is the radius of the circle used to define the
trigonometric functions, and h runs over a certain number of degrees. Thus, in
computing tables for use in astronomy, timekeeping and regulating the times of
prayer Muslim astronomers noticed that, in the course of the computations, they
8 Auxiliary Functions 177

were computing the same things over and over again. Only the value of the
argument was varying.
It soon occurred to some of the astronomers to lessen the labor of computing
tables of functions by computing a set of tables of those auxiliary functions that
appeared often. As early as the mid-ninth century Ḥabash al-Ḥāsib tabulated,
among other functions, the two given above, for R = 60. Later in the same century
al-Faḍl al-Nayrīzī again tabulated the two functions above for R = 150, a parameter
common in Indian trigonometry. Then, late in the tenth century, al-Bīrūnī’s teacher,
Prince Abū Naṣr, tabulated four such auxiliary functions with R = 1, perhaps as
much to show the utility of this value of the radius as for any other reason. At about
the same time, in Ibn Yūnus’ works, we find auxiliary functions which are functions
of two arguments, but in which one of them (the latitude) assumes only the value of
the latitude of Cairo or, sometimes, of Baghdad. The next major step in the cal-
culation of auxiliary tables was taken by the fourteenth century astronomer
Muhammad al-Khalīlī of Damascus, He generalized Ibn Yūnus’ tables and calcu-
lated the following two-argument functions:
1. f(u, h) = R  Sin h/Cos u, h = 1°, …, 90°; u = 1°, …, 55° as well as 21;30° (the
latitude of Mecca) and 33;31° (the latitude of Damascus).
2. g(u, h) = Sin h  Tan u/R, for h, u as above.
3. G(x, y) = arc Cos(xR/Cos y), x = 1,…, 59, y = 0°, 1°,…, n(x)°, where n(x) is the
largest integer such that x does not exceed Cos(n(x)). (Recall that Cos h is a
decreasing function from 0° to 90°.)
Al Khalīlī’s tables, which contain over 13,000 entries, allow one to solve any of
the fundamental problems of spherical astronomy for any latitude and so represent
general solutions to these problems. In the following chapter we shall see that
al-Khalīlī also provided such a solution in the case of the direction of Muslim
prayer.

9 Interpolation Procedures

Out of the mass of labor in computing the tables described in the previous sections
two parts require particular mathematical insight: (1) Calculating Sin 1°; and
(2) Framing rules to allow the user to interpolate from the tables. Zījes contain a
variety of ingenious methods of interpolation, and we shall discuss an example of
second-order interpolation. Following that we shall study an interative method of
al-Kāshī that allows the rapid calculation of Sin(l°), accurate to as many places as
one wishes.
178 5 Trigonometry in the Islamic World

Fig. 19 a, b Transcription of part of al-Bīrūnī’s Sine Tables

9.1 Linear Interpolation

Figure 19 shows part of the table of sines from al-Bīrūnī’s Mas’ūdic Canon
together with a transcription of the table from the alphabetic notation into our
numerals. Al-Bīrūnī’s sine function is very close to the modern one, since for each
h, 0 < Sin h < 1. One difference is a conceptual one in that we view the sine as a
function of a central angle, while al-Bīrūnī saw it as a quantity that depends on the
arc. A further difference, not apparent from the table, is that al-Bīrūnī’s function
assumes only positive values for, in the Canon, he tells the reader that:
9 Interpolation Procedures 179

(1) When 90° < h < 180° then sin h = sin(180° – 0), a fact we have already seen
Naṣīr al-Dīn use in his proof of the Sine Theorem.
(2) When 180° < h < 270° then sin h = sin(h – 180°).
(3) When 270° < h < 360° then sin h = sin(360° – h).
Thus the values of al-Bīrūnī’s sine function are the absolute values of our
modern function, and since neither al-Bīrūnī nor any other Muslim astronomer had
the convenience of negative numbers, it was often necessary to break an argument
up into several cases in order to specify in each case the direction in which a given
line segment or arc should be taken.
In Fig. 19 there are four columns, headed “Degrees/Minutes,” “The Sines,” “The
Corrections,” “The Differences,” explained as follows. The first entry in Col. 1 is
15′ and each succeeding entry is 15′ more than the previous entry, so Col. 1 lists
arcs h, by increments of 15′ from 15′ to 90°. Column 2 then gives, for a given arc h,
the value of sin h, so that, for example, the third row of this column may be read as
saying sin 45′ = ;0,47,7,21. (The reader may want to think for a minute to discover
what the next two columns contain.) Column 3, headed “Corrections” simply gives
for each value h the value 4D(h), where D(h) is recorded in the fourth column as sin
(h +15′) – sin(h).
Exercise 5 makes it clear that the third column in al-Bīrūnī’s tables facilitates
linear interpolation, and such interpolation works very well for the sine function. (In
fact, if the reader converts the sexagesimal value for sin(l°22′) obtained from
Exercise 5 to decimal values the result is 0.02385051, rounded to eight places,
whereas the correct value, rounded to eight places, is the only slightly higher value
.02385057.)
In general, linear interpolation works well over small intervals for functions
whose rate of growth does not change too much over the interval. However, for a
function such as the tangent function, which has a vertical asymptote at 90°, linear
interpolation does not produce satisfactory results and more refined methods are
necessary. Such methods were developed early in the history of trigonometry, in
fact in India before the Islamic era.
Muslim scientists also needed interpolation for other purposes. For example,
they needed ephemerides, sets of tables giving the positions of the sun, moon and
planets at equally spaced intervals (e.g. of one day or five days) throughout the year,
but to compute such a set of tables was a great deal of work, often involving the
production of various auxiliary tables and further computations based on these. In
order to lighten this labor, Muslim astronomers would use interpolation procedures
in compiling the tables, just as we know they used them in applying the completed
ephemerides to find the position of a planet at a time not entered in the tables. In the
next chapter, when we survey some of the tables used in spherical astronomy,
which could have anything from 30,000 to 250,000 entries, the reader will see other
ways in which interpolation methods were used.
180 5 Trigonometry in the Islamic World

9.2 Second-Order Interpolation

For the present, however, we are going to study a method of interpolation that
appears in al-zīj al-Ḥākimī composed by the Egyptian astronomer Abu 1-Ḥasan ibn
Yūnus, the son of an eminent Egyptian historian who became one of the great
astronomers of the Islamic world. We do not know precisely when he was born, but
since his father died in 958 Ibn Yūnus witnessed not only the conquest of Egypt by
the Fatimid kings in 969 but also the foundation of the city of Cairo by the same
dynasty, a dynasty which claimed descent from Fāṭima, daughter of the Prophet
Muḥammad.
Ibn Yūnus counted as his patrons at least two of the Fatimid kings, al-‘Azīz, who
reigned for roughly 20 years until 996, and al-Ḥākim, a firm believer in astrology
who considered himself to be God.
However, Ibn Yūnus yielded to no one in eccentricity. Al-Ḥākim himself told
the following story of Ibn Yūnus’ indifference to custom. It seems that Ibn Yūnus
came into his presence one day carrying a pair of heavy shoes. He sat near al-Ḥākim
for awhile, while the ruler eyed the shoes—objects which court etiquette required
be left outside the throne room. Finally, Ibn Yūnus kissed the ground, put the shoes
on, and left. Another time Ibn Yūnus and a fellow astronomer went up into the
Muqaṭṭam hills outside of Cairo. They observed Venus for some time, and then,
according to the biographer Ibn Khallikān, “Ibn Yūnus took off his cloak and
turban. Then he put on a woman’s red cloak and a red veil, and took out a lute. This
he played on, with incense burning in front of him. It was a remarkable sight.”
(Transl. from King (1972).) In addition to his expertise on the lute Ibn Yūnus
achieved renown as a poet, and several of his works were contained in anthologies.
Ibn Yūnus named his great set of astronomical tables al-zīj al-Ḥākimī (The
Hakemite Zīj) after his patron al-Ḥākim, who probably appreciated Ibn Yūnus more
for the reputed accuracy of his astrological predictions. The historian Ibn Abī
Ḥajala tells the following story:
Another example of his correct (astrological) predictions took place when al-Ḥākim had
given him a house. He said, “Prince of the Believers, I desire that you give me another
house.” (Al-Ḥākim) asked why. He said, “Because water will destroy it (the one I have) and
everything in it.” Al-Ḥākim gave him another and he moved out (of the first) the next
morning. Three days later a mighty torrent came down on Cairo from the mountain and
threw down palaces and houses—a frightening event such as had never been experienced
before—and the above-mentioned house was among what was destroyed, as he had pre-
dicted (King 1972).

The same source says Ibn Yūnus predicted the day of his death and, after he had
locked himself in his house, he told his servant girl, “Iḥsān, I have locked what I
shall never open.” He then took some water and began washing the ink off his
manuscripts, and finally, continually reciting the Quranic verse “Say God is one,”
he died. This was in the year 1009.
The story of Ibn Yūnus washing the ink off his manuscripts is consistent with
another story that after his death his son sold all his books, by the pound, in the
9 Interpolation Procedures 181

Plate 1 Sine tables attributed to Ibn Yūnus. Extract for arguments 22° (on the right) and 23° (on
the left). The columns on the left of each page are for interpolation. The horizontal argument is the
number of degrees (22° and 23° in this case) and the vertical argument is minutes (from l′–30′ and
31′–60′). (Taken from MS Berlin Staatsbibliothek, Ahlwardt 5752 fols. 13v-14r (Lbg 1038).)

Cairo soap market. After the washing there may have been little left to do with them
but sell the remainder for scrap.
Ibn Yūnus intended that the interpolation procedure he described be used in
trigonometric tables calculated for intervals of 30′ (see Plate 1). (However, the
extract from his Sine tables (shown in Plate 1) shows that he tabulated the Sine for
intervals of 1°.) In the following account we shall consider an arc h + k′ where h is
an integer number of degrees and 0′ < k′ < 60′, and LSin will denote the value of
the Sine obtained by linear interpolation. The procedure, then, is the following:
(1) Use linear interpolation between successive degrees to find LSin(h + k′) and
LSin(h + 30′).
(2) From the table find Sin(h + 30′).
(3) Define the “base for the interpolation” to be 4(Sin(h + 30′) – LSin
(h + 30′)) = B. (Ibn Yūnus has already observed that Sines found by linear
interpolation are always less than the actual values.)
(4) Calculate B  k′  (60′ – k′) and note B  k′ – (60′ – k′) = B  k′  (60 – k:)/3600.
(5) For the value of Sin(h + k′) take LSin(h + k′) + B  k′  (60′ – k′).
It seems that the discoverer of this method had the idea of beginning with linear
interpolation over a given l°-interval and then correcting by an amount that would,
at the midpoint of the interval, bring the computed value to the true value. Perhaps
182 5 Trigonometry in the Islamic World

it was experience with tables and interpolation, rather than geometrical arguments,
that taught astronomers that not only is LSin h less than Sin h but that, over a given
interval, the difference between these two would be greatest (or nearly so) at the
midpoint. The problem, then, is to find a function, f, defined from 0′ to 60′ so that
f(0′) = f(60′) = 0 and f(30′) is the greatest value, for then

Sinð# þ 300 Þ  LSinð# þ 300 Þ


LSinð# þ k 0 Þ þ f ðk 0 Þ 
f ð300 Þ

will be a good rule for interpolation. It produces the value Sin(h + 30′) when
k′ = 30′, and elsewhere a value that lies between the value obtained by linear
interpolation and the true value. In fact, one may take f(x) = x(60 – x), and any
competent mathematician from the time of Euclid onward would have realized that
the value of the product x  (a – x) is greatest when the two factors x and a – x are
equal. When a = 60, this implies that x = 30, which yields the maximum value of
f as 900″.
Ibn Yūnus makes no attempt to show that this rule produces better approxi-
mations than linear interpolation, and it is typical of mathematics in the ancient and
medieval worlds that no attempt was made to axiomatize these numerical methods
or to give proofs of their validity. For the scientists of Ibn Yūnus’ time these rules
were only procedures, doubtless arrived at by plausible reasoning, but not proved,
which the practitioner would find worked. See Plate 1 for an extract from the Sine
tables composed by Ibn Yūnus.

10 Al-Kāshī’s Approximation to Sin(1°)

Interpolation procedures such as the one discussed in the previous section provide
but one example of mathematical practice leading theory. Another example is
featured in this section, an iterative solution to a third degree equation.
When Ptolemy approximated Crd(l°) in his Almagest he made use of the
inequality:
   
2 3  4 3
Crd \Crd(1 Þ\ Crd ;
3 2 3 4

which gave him an approximation to Crd(l°) correct to two sexagesimal places,


since the two extreme terms of the inequality both begin 1;2,50. However, the
method has an inherent limitation in that, whatever bounds one uses on either side,
they will agree only to a certain number of places (since they are not equal), and
there is no possibility of getting greater accuracy without finding new bounds.
10 Al-Kāshī’s Approximation to Sin(1°) 183

Refinements of Ptolemy’s method were used by mathematicians of the Islamic


period to find an approximation of Sin(l°), but it was Jamshīd al-Kāshī who, in
Samarqand early in the fifteenth century, discovered a method which will provide
arbitrarily close approximations to Sin(l°), a method based on two relationships:

Sin(3hÞ ¼ 3Sin(hÞ  ;0,4(Sin h)3 ; ð1Þ

which, when h = 1°, becomes

Sin(3 Þ = 3 Sin(1 Þ  ;0,4(Sin 1 Þ3 ;

and on

Sin(3 Þ = 3;8,24,33,59,34,28,15, ð2Þ

which is exact as far as it goes. This value for Sin(3°) can be determined as exactly
as necessary because Euclidean procedures allow one to find both Sin(72°) and Sin
(60°) from the constructions of the sides of a regular pentagon and an equilateral
triangle in a given circle. These Euclidean constructions, when translated into
algebraic equations, demand nothing more than the solution of first or second
degree equations. The roots of these may be expressed in terms of, at worst, square
roots, which may be approximated to any desired accuracy. Then the formula for
the sine of the difference of two arcs, which we have seen was known to Abū
al-Wafā’ in the tenth century, would yield Sin(12°) = Sin(72° – 60°) to any desired
accuracy, and this, by repeated use of the half-angle formula, would yield in turn
Sin(6°) and Sin(3°). When we substitute this value of Sin(3°) into the equation and
write x for Sin(l°) we obtain, after a bit of arithmetic, the fundamental relation:

x3 + 47,6;8,29,53,37,3,45
x=
45; 0

a cubic equation one of whose roots is Sin(l°).


Al-Kāshī knows this equation has a root near 1, so we may write the root as 1;a,
b,c,… where a,b,c,… are the successive “sexagesimal places” of the root. Then
substituting this value for x we obtain

ð1; a; b; c; . . .Þ3 + 47,6;8,29,53,37,3,45


1; a; b; c; . . . =
45; 0

So, if we subtract 1 from each side,

ð1; a; b; c; . . .Þ3 + 47,6;8,29,. . .


; a; b; c; . . . = 1
45; 0

which simplifies to
184 5 Trigonometry in the Islamic World

ð1; a; b; c; . . .Þ3 + 2,6;8,29,. . .


; a; b; c; . . . = :
45; 0

Place by place the two sides are equal, so in particular the first sexagesimal place
of the right-hand side must be a. However, because 45,0 is so big and the root (and
hence its cube) is near 1 the first place of the right-hand side does not depend on the
value of a, To convince ourselves of this we can evaluate the right-hand side at 1;59
(or even 2) instead of at 1, and when we do we find that

23 + 2,6;8,29,. . .
¼; 2ð58 or 59Þ; . . .
45; 0

Thus to find out what a is we need only evaluate

ð1Þ3 + 2,6;8,29,. . .
¼; 2ð49 or 50Þ; . . .
45; 0

so that a = 2.
Then again,

ð1; 2; b; c; . . .Þ3 + 47,6;8,29,53; 37; 3; 45


1; 2; b; c; . . . = ;
45; 0

and al-Kāshī now uses the fact that the second digit on the right will not depend on
b, because of the size of the divisor 45,0. (One can check this by evaluating the
right-hand side with x = 1;2 and x = 1;3 to get 1;2,49,39,… and 1;2,49,43,…
respectively.) Clearly, then, we may set x = 1;2 and obtain b = 49.
If, now, we write f(x) = (x3 + 47,6;8,29,…)/45,0 then we may state al-Kāshī’s
idea as follows: Since f(x) increases so slowly near 1 the value of the nth digit of f(x)
does not depend on the value of the nth digit of x but only on the value of the first
n – 1 digits of x. We have seen that, at least for the first two digits, the idea works,
but will it always work?
Al-Kāshī does not address this question but continues computing to the ninth
sexagesimal place (60–9) and obtains the result Sin(l°) = 1;2,49,43,11,… 17, and
one may check that for this value of x, f(x) = x, very nearly. Thus, al-Kāshī dis-
covered a method for approximating Sin(l°) which will produce a value as near the
true value as one wishes.
Methods such as the one al-Kāshī used are called iterative methods, which
means that one begins with certain data, in al-Kāshī’s case Sin(3°) and an
approximation (generally one that is fairly crude but at least near) to the correct
answer. One then uses the data and the initial approximation in a given procedure to
arrive at a number. This number is then taken as the new approximation, and it,
together with the data, is then put into the same procedure for a second round of
computation. This computation produces another approximation which, together
10 Al-Kāshī’s Approximation to Sin(1°) 185

with the data, is again put into the procedure, etc. If the procedure is effective, the
successive results will approach nearer and nearer to one value, which will be the
value that solves the problem. In this case we say that the procedure converges and
the algorithm, or procedure, is an effective one.
In fact, in al-Kāshī’s algorithm the results of the successive approximations do
approach the value of Sin(1°). The method is taught today in courses on numerical
analysis under the name of “Fixed-point iteration,” where it is proved that the
procedure al-Kāshī used to find a solution A to the equation x = f(x) will converge
provided the curve y = f(x) is a smooth one and the initial approximation is chosen
in a neighborhood of A where the tangent line to the curve has a slope of absolute
value less than 1. The proof of the convergence of the algorithm does not go beyond
the mean value theorem of the differential calculus.
Al-Kāshī did not, of course, know this theorem, nor is there evidence that he
concerned himself with the proof of convergence. He was working on the interface
of the exact sciences and mathematics, the region that has been responsible his-
torically for so much of the growth in mathematics, and his concern was to find
methods which would provide solutions to problems important in astronomy. This
he did, and, as we have seen, he did it well.
Hogendijk and Rosenfeld (2002/3) contains a commentary on al-Kāshī’s
method, one composed by a colleague of his in Samarqand, and, for the present, this
is as close as we can come to reading al-Kāshī’s own account.
We have seen in the preceding sections that Muslim mathematicians organized
trigonometry into a systematic discipline, whose theory rested on a full complement
of six functions and a variety of powerful results, such as the Law of Sines and
certain key trigonometric identities. In addition, the application of this theory was
made possible by extensive, highly accurate trigonometric tables, including tables
of auxiliary functions, a variety of methods for interpolation—a technique that has
been nicely characterized as “reading between the lines,” and iteration. Thus
trigonometry takes its place, alongside algebra and arithmetic, as part of our her-
itage from Islamic mathematics.

Exercises

1. Discover the value of p implicit in the ancient Indian value of 3438′ for the
radius of a circle, where, as we said in the text, the minute is taken to be the
same length as a minute of arc on the circle.
2. In Nasīr al-Dīn’s discussion of solving a right triangle when two sides are
known, let the sides be a, b, and g as in Fig. 3, and set b = 120u. Show that to
express a and b in terms of the new unit, u,
186 5 Trigonometry in the Islamic World

a must be taken to be ða=bÞ120u

and

g must be taken to be ðg=bÞ120u:

3. Calculate the values u, a and g (according to Problem 2) for a 3-4-5 right


triangle, and use Ptolemy’s table of chords (Fig. 1) to calculate the smallest
angle of such a triangle.
4. Use Naṣīr al-Dīn’s explanation of how to solve right triangles, together with
Ptolemy’s table of chords, to solve the 3–4–5 right triangle.
 
5. Use the value for Crd 12 given in Ptolemy’s table of chords to estimate the
value of p. How accurate is this estimate?
6. Use al-Bīrūnī’s sine table to calculate sin(l°22′) from the formula

sinðl 220 Þ ¼ sinðl 150 Þ þ ð; 7ÞCðl 150 Þ;

where C(h) is the value recorded in the third column of Fig. 19b.
In general show that for any h, 0 < h < 90°, if h′ is the largest multiple of 15′
that is less than h then, then

sinðhÞ ¼ sinðh0 Þ þ ðhh0 ÞC(hÞ:

(Here, h – h′ must be read not as “minutes” but as sixtieths of the parts that make
up the radius.)
7. Use Ibn Yūnus’s procedure to calculate a value for Sin 1°22′, given the tabular
values Sin(1°) = 1; 1,49,45, Sin(1°30′) = 1;34,14,13 and Sin (2°) = 2;5,37,17.
Show that your answer lies between the value produced by linear interpolation
and the true value. Show also that for any value of the argument h the value of
Sin h produced by Ibn Yūnus’ procedure lies between that produced by linear
interpolation and the true value. (The values are from al-Bīrūnī’s Mas‘ūdic
Canon, discussed in Sect. 9 above.)
8. Prove each of the following statements from al-Marrākushī’s exposition of the
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Vers function. (1) Sin2h + Vers2h = Crd2h; (2) Crd (180°- h) = d2  Crd2 h,
where d is the diameter of the reference circle; (3) dVersh = Crd2h.
9. Derive the final form of al-Kāshī’s cubic equation as we have given it in the text
from the form:

Sinð3 Þ ¼ 3 Sinðl Þ - ;0; 4ðSin 1 Þ3 ;

(Hint: It helps to use the fact that (;0,4)/3 = ;0,1,20.)


Exercises 187

Fig. 20

10. If x2 = 2 show that x = (x + 2/x)/2. Use al-Kāshī’s method and the initial
pffiffiffi
approximation x = 1 to calculate three more approximations to 2. Heron of
Alexandria, who wrote in Greek ca. A.D. 60, recommends this method for
approximating square roots in his Metrica. (Dr. C. Anagnostakis pointed out to
me that this procedure is an example of al-Kāshī’s algorithm.)
11. Let g(x) = x3 + 4x2 – 10, and show that the equation g(x) = 0 has a root in the
interval (1,2). Now use algebra to show that this root is also a root of the
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
equation x = f(x), where f(x) = 10=ðx þ 4Þ: Finally, starting with an initial
approximation x = 1 to the root of this equation, use al-Kāshī’s method to find
the successive decimal places of the root. Conclude that, to three decimal
places, the value of the root is 1.365.
12. The tenth century writer Abū Ṣaqr al-Qabīṣī, in the context of arguing that the
height of any known mountain is negligible compared to the radius of the earth,
gives the following method of finding the height of a mountain. Justify it.
Let BGD be the surface of the earth and AB the height of the mountain
(Fig. 20). From the two points G,D, whose distance apart is assumed known,
one measures the angles AGB and ADB. Then

GD  Sin(D)
AB = :
Sin(90 D)Sin(90 G)Sin(D)=Sin(G)

Since all quantities on the right are known we can calculate AB.

Bibliography

Aaboe, A. 1954. “Al-Kāshī’s Iteration Method for Sin(l)”. Scripta Mathematica 20: 24–29.
Aaboe, A. Episodes from the Early History of Mathematics. The Mathematical Association of
America,1997. Hamadanizadeh, J. “A Survey of Medieval Islamic Interpolation Schemes”. In:
From Deferent to Equant: A Volume of Studies of the History of Science in the Ancient and
Medieval Near East, Dedicated to E. S. Kennedy. (King, D. A. and G. A. Saliba, eds.). New
York: New York Academy of Sciences. 1987, pp. 143 – 152.
188 5 Trigonometry in the Islamic World

King, D. A. The Astronomical Works of Ibn Yūnus. (Unpublished Yale Ph.D. thesis.) New Haven:
Yale University, 1972.
Neugebauer, O. 1962. The Astronomical Tables of al-Khwarizmi. Copenhagen: Dansk
Videnskabernes Selskab 4: 2.
Rosenfeld, B. and J. P. Hogendijk. “A Mathematical Treatise Written in the Samarqand
Observatory of Ulugh Beg.” Zeitschrift für Geschichte der Arabisch-Islamischen
Wissenschaften. 13 (2002/2003), 25 – 65.
Sédillot, J.-J. and L.-A, Sēdillot. Traité des instruments astronomiques des Arabes. Reprinted by
Institut für die Geschichte der arabisch-islamischen Wissenschaften. Frankfurt a.M., 1984.
al-Ṭūsī, Naṣīr al-Dīn (transl. by C. Caratheodory). Traité du Quadrilatère. Constantinople: 1891.
Van Brummelen, Glen. The Mathematics of the Heavens and the Earth: The Early History of
Trigonometry. Princeton, New Jersey: Princeton University Press, 2009.
Chapter 6
Spherical Trigonometry in the Islamic
World

1 The Ancient Background

The problems of spherical trigonometry concern the sizes of circular arcs or angles
on the surface of a sphere, and their relationships to each other. In applications, the
sphere was either the celestial sphere or the earth, the former being a sphere that
was thought to contain the fixed stars and to have such a large radius that, in
relation to it, the earth was no more than a dot. However, the radius was finite and,
in principle, its magnitude could be calculated.
In the theoretical treatment of the geometry of the surface of sphere, the ana-
logues of the straight lines on a plane are the great circles, which are the inter-
sections of the surface of the sphere with any plane through its center. Also
important are parallel circles, which are formed by the intersection of the surface of
the sphere with planes that do not pass through its center.
In ancient Greek mathematics and astronomy a circle was a two-dimensional
plane surface and its bounding curve was always called its circumference, not
simply a circle. Similarly, a sphere was a geometrical solid and its surface was
always so-called, never simply a sphere. Hence, any two great circles on a sphere
would intersect in a straight line through the center of the sphere, not just in two
diametrically opposite points.
In the case of the terrestrial sphere, the important great circles are the equator and
meridians, and the parallels of latitude are parallel circles. In the case of the celestial
sphere some important great circles are the celestial equator, the ecliptic and the
horizon, which we shall define and discuss in the next section. It is these inter-
pretations of great circles and parallel circles that give the subject of spherical
trigonometry its utility both in astronomy and in mathematical geography.
It was the Greeks who first investigated the geometry of the surface of the
sphere, and they called the subject “spherics” (ta sphaerika), a term we shall also
use in this chapter. Surviving treatises of Autolykos, who was probably a

© Springer Science+Business Media New York 2016 189


J.L. Berggren, Episodes in the Mathematics of Medieval Islam,
DOI 10.1007/978-1-4939-3780-6_6
190 6 Spherical Trigonometry in the Islamic World

contemporary of Euclid in the fourth century B.C., show that the following basic
facts were known long ago (Fig. 1a–c).
1. Any two great circles of a sphere bisect each other.
2. Given any two diametrically opposite points on the sphere, consider all great
circles joining these two points. Then there is a unique great circle lying on the
plane perpendicular to all those great circles. Conversely, given any great circle
there are two diametrically opposite points called its poles such that any circle
through these poles is perpendicular to the given circle.
3. Since any two given great circles of a sphere bisect each other it follows that
they intersect at diametrically opposite points A, B. Then (2) guarantees that
there is a unique great circle with A, B as poles. The smaller arc that is cut off
from this circle by the two given circles measures the smaller angle a between
these circles.
Menelaos, who conducted astronomical observations in Rome and lived a few
decades before Ptolemy, is the first writer we know of to base his study of the
geometry of the sphere on spherical triangles. In his work Spherica, a spherical

Fig. 1
1 The Ancient Background 191

triangle is defined as “the area enclosed by the arcs of (three) great circles on a
sphere, each arc being less than a semicircle,” and in Book III of the Spherica we
find a theorem that is not only the first theorem of spherical trigonometry but, for
the Greek writers, the only theorem of that science.1 It has become known as
Menelaos’s Theorem, and may be stated in a slightly modernized form as follows
(Fig. 2):
_ _
Let AB and AG be two arcs of great circles on the sphere, and let two other arcs
_ _
GD and BE meet within the angle the first two arcs enclose, say at Z. In addition,
let all four arcs be less than semicircles. Then,
 _  _ h  _   _ i h  _   _ i
Crd 2GA : Crd 2EA ¼ Crd 2GD : Crd 2ZD  Crd 2ZB : Crd 2BE :

As the reader may suspect, this is but one of the many cases of the theorem.
Ptolemy states one more, and medieval Islamic writers obtained 72 cases in all of
what they called the “figure of the complete quadrilateral”. For our purposes it suffices
to note that it often requires ingenuity in completing a few given arcs to obtain a
configuration to which one could apply the theorem, while a single triangle would be
much easier to find. (For example, compare Exercises 5, and 6 of this chapter.)
In addition to Menelaos’s Theorem, other methods for finding arcs or angles on
the sphere were also of some advantage, either in getting reasonable approximations
easily or in demonstrating basic facts to the beginning student. One such method is
simply to construct a good model of a sphere and then engrave on it the important
great or parallel circles as well as the positions of the important stars (in the case of

Fig. 2

1
Ironically, it is not about spherical triangles!
192 6 Spherical Trigonometry in the Islamic World

the celestial sphere) or the major geographical features (in the case of the earth).
Then to solve problems of spherics one could simply use a substance such as
colored wax or chalk to mark on any additional arcs or angles needed and then
measure the ones desired.
Increasingly, sophisticated variations on the above practical approach were
described by a series of writers. For example, in the ninth century, Qusṭā ibn Lūqā
wrote a treatise On the Sphere with a Frame, and in the twelfth century ‛Abd
al-Raḥmān al-Khāzinī described an automated device of this sort in a treatise titled
On the Sphere that Moves by Itself (This treatise is described in Lorch 1980). Such
devices were known not only to the Muslim authors but in the ancient world as well,
and al-Khāzinī’s treatise belongs to the tradition of powered models that can be
traced back to Archimedes’ moving model of the sun, moon, and planets rotating
around the earth. A simpler version of the same idea was the solid sphere surrounded
by a framework of graduated rings, each corresponding to an important circle. Such
an instrument is described in Ptolemy’s Almagest, where it is called “an astrolabe”,
but we know it as an armillary sphere (from the Latin armilla = bracelet), for so the
rings around the sphere would appear). Plate 1 shows an illustration of an armillary
sphere, with the earth at the center, from a manuscript in the Vatican.

2 Important Circles on the Celestial Sphere

Since in much of the following, we shall be concerned with the applications of


spherics to the celestial sphere we shall here introduce the reader to some of the
most important circles and angles on that sphere, and we shall begin with the one
that is most apparent to the reader. In our explanation we shall use the capitalized
word “Star” to denote a star, the sun, the moon, or any of the five naked-eye
planets2 when the discussion applies equally to all of these objects.
When one looks around on a wide plain one sees in all directions the line
bounding sky and earth. Anything in the heavens that is above this line is visible
and anything below it is invisible, for which reason the ancient Greeks called it
horizōn, that is the “bounding” or “defining” circle, and from this we take our
phrase “horizon circle.” The points on the celestial sphere which are directly over
our head and directly underneath our feet are the poles of this circle. The modern
names for these poles, zenith and nadir, respectively, come from the Arabic words
samt, “direction” (of the head), and nāzir “opposite” (the feet). The great circles
joining these two points are called altitude circles, and one of these that passes
through the north and south points of the horizon is called the meridian of our
locality. The reader who imagines a star and the altitude circle that passes through it
will easily see that the smaller arc of that circle between the star and the horizon,

2
These planets, the ones that can be seen without a telescope, are Mercury, Venus, Mars, Jupiter
and Saturn.
2 Important Circles on the Celestial Sphere 193

Plate 1 An illustration of an armillary sphere taken from MS Vatican Borg. ar. 817, fol. lr
(Reproduced courtesy of the Vatican Apostolic Library.)

measured in degrees, is a reasonable measure of the altitude of the star, i.e., its
degrees above the horizon, and hence the name altitude circle for that great circle
passing through the zenith and the star. Again, the lesser of the two angles between
the altitude circle and the local meridian measures how many degrees one must turn
194 6 Spherical Trigonometry in the Islamic World

from the N–S line to see the star, and this angle is called the azimuth of the star,
again from the Arabic word samt, “direction” (of the star).
One of the most striking daily phenomena in the heavens is the rising and setting
of the sun, moon and stars. The fixed point in the sky around which they appear to
rotate each day is called a (celestial) pole, north or south depending on whether one
is in the northern or southern hemisphere, for there are two poles, of which only one
is visible to a given observer. (Note that stars sufficiently near the visible pole never
set.) The circle a star makes as it rotates around the pole in the course of 24 h is
called its day-circle, and that part of its day-circle that is visible is called, for
obvious reasons, its arc of visibility.
The great circle perpendicular to all the great circles through the north and south
poles is called the (celestial) equator. If one imagines these poles and circles
somehow visible in the heavens then the equator would be seen just rotating into
itself, going around the earth every 24 h, but the great circles joining the poles and
perpendicular to the equator would be seen actually rotating around the heavens,
always passing through the fixed poles. Any Star lies on one of these circles and the
smaller arc of the circle contained between some star and the equator is called the
declination (symbolized by “d”) of the star, which may be thought of as measuring
the altitude of the star relative to the equator. Also, the angle between the circle
passing through the star and the meridian is called the hour-angle of the star, and is
useful in telling local time. For example, when the star is the sun the hour angle
measures the hours before or after noon, using the conversion 15° = 1 h.
Finally, if one watches the skies a bit before sunrise one will see some star rise
above the eastern horizon just before it becomes too light to see the stars any more.
In other words, the sun is near that star. After another week or so one will again see
a star rise just before sunrise, but it will not be the same star as before, which will
have risen some time earlier, so it appears that the sun has moved relative to the
stars, and it is now near another star. If one watches during the course of a year one
will see the sun go completely around the heavens, at the rate of about a degree
each day, and return to the same star. In fact, the sun appears to follow a great circle
around the heavens, and this circle is called the ecliptic (from the Greek ekleipein,
which means “to eclipse”). If one takes a narrow band, say 5° wide on either side of
the ecliptic, then not only the sun, but the moon and the five naked-eye planets all
appear to move relative to the stars within this band. The sun is always on the
ecliptic, so when the moon is on the ecliptic there is a possibility of an eclipse,
which is why the Greeks chose the name they did.
Since the ecliptic and equator are two great circles they intersect at diametrically
opposite points, where they form an angle of approximately 2312 , which is called
the obliquity of the ecliptic and is written with the Greek letter e. The points of
intersection are called the spring and fall equinoctial points (because when the sun
is at these points day and night have equal length). One of these points, for the
spring, is taken as the 0° point for longitudes on the ecliptic, which are measured in
the counterclockwise direction looking down on the ecliptic from the north.
2 Important Circles on the Celestial Sphere 195

The narrow band that surrounds the ecliptic on each side is called the zodiac and
is divided into twelve segments, each 30° long. The names of the signs, beginning
at the point of the spring equinox and running counterclockwise, are the following
(reading across the page):

Aries Taurus Gemini


Cancer Leo Virgo
Libra Scorpio Sagittarius
Capricorn Aquarius Pisces

According to this arrangement, the point which is 90° (counterclockwise, as one


looks down at the sphere from its north pole) from the beginning of Aries, and is
therefore the northernmost point on the ecliptic, is the beginning of the sign of
Cancer. Thus, the circle that point makes as it turns during the day is called the
Tropic (from the Greek tropos, “turning”) of Cancer. Likewise, the southernmost
point of the ecliptic (six signs away from the beginning of Cancer) is the beginning
of Capricorn, so its turning creates the Tropic of Capricorn.
It seems to be one of mankind’s enduring beliefs that the arrangement of the sun,
moon and various planets within the zodiac as well as the placement of the zodiac
relative to the horizon at the time of some event (the birth of an individual, the
founding of a city, the beginning of a military campaign) influence for good or ill
the outcome of the event. Thus, there arose a set of motives for study of the
movement of the heavenly bodies that included self-interest, in addition to such
motives of a practical or scientific nature as the construction of calendars or
reckoning time.

3 The Rising Times of the Zodiacal Signs

A typical problem of spherics, which makes use of both the equator and the ecliptic,
is finding the rising times of arcs on the ecliptic. Thus, Fig. 3 shows the celestial
sphere at a time when the point on the ecliptic of longitude k is rising above the
eastern horizon of a locality. Imagine the sun to be at this point, so it too is just
rising on some day of the year, and disregard the slow motion of the sun on the
ecliptic, which, as we said, amounts to a little less than l°/day. Then, at sunset of
that day, the sun will still have longitude k, but it will now be on the western
horizon. Since any two great circles bisect each other, half of the ecliptic will be
above the horizon at any time of the day, in particular at sunset, and this half of the
ecliptic will have risen during the day as the point where the sun is travels across
the sky. Hence, during each daylight period, 180° of the ecliptic rises over the
horizon. Thus, if we can tell, for any arc of the ecliptic, how long it takes for it to
rise over our horizon (called the rising time of that arc) we will be able to calculate
196 6 Spherical Trigonometry in the Islamic World

Fig. 3

how long daylight will last on any given day (provided we know where the sun is
on that day, of course). In ancient and medieval geography the length of the longest
day of the year at a given locality was one measure of the local latitude, and the
length of daylight was also important in telling time by the sun. Thus it is no
surprise that all Islamic zījes deal with this problem.
One way in which the zījes treat the problem is as follows: To calculate the rising
time of an arc that stretches from k to k′ it is sufficient to be able to calculate the
rising times of the arcs from 0 to k and from 0 to k′, for the rising time of the given
arc will be the difference of these two rising times. Thus, we can solve the problems
of rising times if we can calculate the rising time of any given initial arc from 0 to k.
This is not an easy problem because the heavens rotate around the poles of the
equator, not the poles of the ecliptic, and as a consequence equal arcs of the ecliptic
do not rise in equal times. However, for the same reason, equal arcs of the equator
do rise in equal times and therefore, since 360° rise in a day, 1° of the equator rises
every 4 min. Thus, if we want to find the rising time of an arc of the ecliptic from 0°
to k° we need only compute how many degrees of the equator have risen in the
same time and multiply that number by 4 to obtain the rising time in minutes.
For a locality on the earth’s equator, the celestial equator will pass through the
zenith, so it will be at right angles to the horizon. In this case, the arc of the equator
that has risen with the arc of the ecliptic from 0° to k° is called the right ascension
of that arc and is written a(k). When the locality is not on the equator then the
equator makes an acute angle with the horizon equal to 90°− /, where / is the local
latitude. In this case, the arc of the equator that has risen is written a/(k) and is
called, for obvious reasons, the oblique ascension of k.
Methods for computing a/(k) are found in astronomical treatises of the
Babylonians, Greeks, and the Islamic world and one of the fascinating aspects of
the history of various schemes for computing them is the survival of very old
methods for over 1000 years, long after more sophisticated approaches were
available. We shall say no more of some of the oldest methods, which employed
arithmetical sequences, but later in this chapter we shall see some of the results of
quite sophisticated calculations used by astronomers of the Islamic world. Of
course, we can get results that yield a good insight into the problem by means of the
spherical models we described earlier.
4 Stereographic Projection and the Astrolabe 197

4 Stereographic Projection and the Astrolabe

However, of such spherical models, al-Bīrūnī wrote that to be of any use they must
be of considerable size, but that this very feature of size also means that they are
rarely found and difficult to transport and manipulate. As al-Bīrūnī puts it, in his
laconic style, “Thus, the difficulty in it corresponds to the good in it.” And, whether
it was for reasons of convenience or not, it seems that the astronomer Hipparchos of
Rhodes, the same man who composed the first table of chords that we know of,
wrote a treatise on a method that allows one to represent the surface of a sphere on a
plane so that circles on the sphere are represented by circles on the plane. This
method is now called stereographic projection (stereo = solid, graphein = to
describe), and although Hipparchos’ treatise on the subject is lost, Ptolemy’s The
Planispherium, written almost 300 years later, has survived.
Stereographic projection, familiar to any student of complex variables, may be
described as follows (Fig. 4): A great circle on the surface of a sphere, such as the
equator, is chosen and the plane p containing it will be the one onto which we will
map the surface of the sphere. To effect this mapping, we pick one pole of the great
circle, say the south pole, and then define for any point X 6¼ P on the sphere the
image point X′ on the plane p as the point where the line PX cuts p. Since for any
point X 6¼ P the line PX cuts p in only one point, the image X′ is uniquely defined
for each X 6¼ P. The effect of this mapping on various points X, X1, … on the
sphere is illustrated in Fig. 4.
The utility of the projection lies in the fact that it maps circles on the sphere to
straight lines or circles on p and that it preserves angles. Figure 4 makes it plain that
meridians are mapped onto straight lines through the center of the sphere, and also
that the equator, which is on p, is mapped to itself, as well as the fact that points
south of the equator (such as X1) are mapped outside the equator, while points north
of the equator are mapped inside it.
We do not know whether Hipparchos thought of making any instrument based
on stereographic projection, nor does Ptolemy’s treatise describe any such

Fig. 4 Drawing reproduced courtesy of Paul Mac Allister and Associates


198 6 Spherical Trigonometry in the Islamic World

instrument. However, the first known treatise on such an instrument was written by
Theon of Alexandria late in the fourth century A.D., so at some time prior to this
someone had applied stereographic projection not only to the sphere with its stars,
but to the system of its important great circles. The result was two disks, one
representing the starry sphere and the other the various great circles, in particular
the equator and horizon. When the disk carrying the coordinate net of great circles
rotates on top of that on which the stars are engraved, one has an anaphoric clock.
This ancient instrument was designed to show the time of day or night in the
seasonal hours used in the ancient world (explained in Sect. 5 of this chapter), The
Roman architect, Vitruvius, describes it in IX,8 of his De architectura, Some
evidence suggests that such an instrument may have been housed in the Tower of
the Winds, which was built in the Athens marketplace in the mid-first century B.C.
and is still standing.
However, if a disk carrying pointers to mark the star positions and mounted on a
brass framework rotates over a solid disk carrying the coordinate systems then we
have the astrolabe. (See Plates 2 and 3.) This instrument is more fully described as
“the planispheric astrolabe” since it is a plane, rather than a 3-dimensional, image of
the celestial sphere. But the usual name is just “the astrolabe,” and that is what we
shall call it in what follows.
The first reference to an astrolabe that clearly refers to what we understand by
that word is in a letter of Synesios, Bishop of Ptolemais (probably on the coast of
modern Libya), to his teacher Hypatia, the first woman mathematician whose name

Plate 2 The rete (the top star map) and plate (the disk of celestial coordinates under the rete) of
this astrolabe were made by the astrolabist Abū Balr b. Yūsuf around the year 1200. The mater
(the circular rim around the rete and plate) was made later, probably in the seventeenth century
Maghrib (Northwest Africa). Reproduced courtesy of the Trustees of the Science Museum,
London
4 Stereographic Projection and the Astrolabe 199

Plate 3 A universal saphea on the back of the astrolabe illustrated in Plate 2. The grid is based on
a projection for the latitude of the equator and is hence independent of the terrestrial latitude. With
the usual alidade as shown, the grid is useless. As devised in eleventh century Muslim Spain the
grid must be used together with an alidade fitted with a movable perpendicular cursor. Then one
can convert from one orthogonal coordinate system to another. (Reproduced courtesy of the
Trustees of the Science Museum, London.)

is known to us. Her father, Theon of Alexandria, did an important edition of


Euclid’s Elements and also wrote on the planispheric astrolabe. In fact, he seems to
have equipped Synesios’s astrolabe with a sighting device that allowed the user to
take the altitude of a star or the sun over the horizon—the basic piece of data fed
into the instrument. It was Theon’s treatise that was rendered into Syriac by Bishop
Severus Sebokht, who made the first known reference to Hindu numerals outside of
India, and through the Syriac it became known to the Arabic authors.
Since the stereographic projection maps the sphere (less the south pole) onto the
whole plane it is necessary to limit the size of the image, and this is done by
mapping only the part of the sphere (and its great circles) above the Tropic of
Capricorn, which has a latitude of approximately −23½°. Thus, the whole ecliptic is
represented on the plane, along with the equator, the Tropic of Cancer, the horizon
(or that part of it above the Tropic of Capricorn) and the circles of equal altitude
above the horizon and parallel to it. (These latter are still known by their Arabic
name, almucantar.) Fig. 5 shows the resulting map of the coordinate circles in
section, where capital letters represent points on the sphere and the letters with
dashes are their images under stereographic projection. (“S” represents the south
pole.) Solid lines indicate the equator and circles parallel to it, so AB is the section
of the Tropic of Capricorn, while CD is the section of the Tropic of Cancer. XY is
the section of the local horizon and the line ZW, parallel to it, is the section of an
almucantar. Notice that since, in the case shown, the southern edge of the horizon
200 6 Spherical Trigonometry in the Islamic World

Fig. 5

is below the Tropic of Capricorn, it is not mapped onto the plane, so that, for
example, X has no image.
Figure 6 shows the images of this mapping for the latitude 30°. The top of the
figure represents the south. The outer circle represents the Tropic of Capricorn; the
next smaller circle, concentric with it, represents the equator, and the inner one of
the three concentric circles is the Tropic of Cancer. The center of these circles is the
projection of the north celestial pole and it is represented by a small circle.
The circle south of the pole represents the zenith for latitude 30°, and the circles
clustering around it are almucantars at intervals of 20°, extending down to the
horizon itself. (Anything outside the horizon circle is below it on the globe and thus
invisible). As we said, the southern extremity of the horizon is below the Tropic of
Capricorn, and therefore the horizon (as well as some of the lower circles parallel to
it) is not represented by a complete circle. The curves (actually circular arcs) joining
the zenith to the horizon are the images of the azimuth circles, marking out intervals
of 10° on the horizon, proceeding clockwise from the north point, N, which is 0°.
On top of the plate just described is a circular star map, of the same diameter as
the lower plate, which shows certain important stars above the Tropic of Capricorn
as well as the circle of the zodiac. Only pointers indicating the positions of the stars
and a supporting fretwork are on the star map, and the rest of the brass plate is cut
away to allow the user to see the circles on the lower plate. A post goes through the
center both of the star map and the lower plate, and rotating the star map around the
post imitates the rotation of the heavens around the North Pole. Figure 7 shows the
star map, called rete (=“net”) by the Latin astrolabists, a less colorful term than the
Greek and Arabic equivalents, which mean “spider.”
The exposition above assumes that the stereographic projection used to produce
the planispheric astrolabe is both circle-preserving and angle-preserving, that is to
4 Stereographic Projection and the Astrolabe 201

Fig. 6

say it maps all circles on the sphere onto circles or straight lines on the plane and it
maps angles between circles on the sphere onto equal angles on the plane.3
Although it is likely that the Greeks, who first constructed the astrolabe,
observed at least that angles made by the meridians at the pole are mapped onto
equal angles, they did not remark on it. (The English mathematician, Thomas
Harriot (1560–1621), proved this property).4
As for the first property of the projection, that it maps circles onto circles or
straight lines,5 the Greeks must have known this. However, Ptolemy makes no

3
Mappings that preserve angles are called “conformal.” An angle on the sphere is an angle formed
by the planes of two intersecting circles.
4
Pepper, pp. 366–67. Harriot’s proof may be found on pp. 411–13 of the same paper. An elegant
proof of the conformality of stereographic projection may be found in Hilbert/Cohn-Vossen,
pp. 248-49.
5
In case the circle contains the pole of the projection.
202 6 Spherical Trigonometry in the Islamic World

Fig. 7

reference to it in his treatise on the instrument, The Planispherium, nor did any
other Greek writer. The fact was, however, stated and proved by Abū al-‘Abbās
al-Farghānī, who wrote The Complete [Book] on the Construction of the Astrolabe
probably in the years 856-7.6 Here, again, we see an advance in sophistication, in
that sense that it was not sufficient to observe that things seemed to work. One
wanted to prove they worked!
Al-Farghānī was evidently a friend of the Banū Mūsā, who subcontracted to him
the task that the Caliph al-Mutawakkil had assigned to them of constructing a canal in
a new city he was having built. Although al-Farghānī had successfully built the
Nilometer in Egypt he was not successful with the canal, but, fortunately for him (and
the Banū Mūsā), al-Mutawakkil was murdered before his mistake came to light.
Al-Farghānī was clearly more successful with his book on the astrolabe, for
Ibrāhīm b. Sinān, whose geometrical expertise we have shown in Chap. 3, praised it
as one of the recent books that “comprehend everything necessary for this [the
construction of the astrolabe].”
After a brief introduction about what is new in his book, al-Farghānī begins the
mathematics

6
According to Lorch (2005, p. 5).
4 Stereographic Projection and the Astrolabe 203

Fig. 8

Theorem 1 Suppose ABGD is a circle with diameter AG (Fig. 8), line EZ is


tangent to the circle at A, and BH is any chord of the circle. Draw lines GB and GH
and extend them to meet the tangent EZ at K and T, respectively. Then D(GKT) is
similar to D(GHB), [i.e., ∠TKG = ∠GHB and ∠KTG = ∠HBG].
Proof Drop a perpendicular from B onto GA and let it meet GA at L. Because GAB
and GBL are right triangles with a common angle at G the remaining angles, GBL
and GAB are equal. And because BL and AE are both perpendicular to GA they are
parallel. Hence the two angles, GBL and GKA, are equal. Hence ∠GKA = ∠GAB.
And ∠GAB = ∠GHB because they are inscribed angles that subtend the same arc,
BG. Hence, ∠GKA = ∠GHB. Therefore, two angles of triangle GTK are equal to
two angles of triangle GHB, and so the remaining angles GBH and GTK are equal.
Thus, D(GKT) is similar to D(GHB), which was to be proved.
Now al-Farghānī states and proves the second theorem.
Theorem 2 Let there be a cone ABGD (Fig. 9), whose base is a circle BGD.7
Around the cone is circumscribed a sphere, so the vertex of the cone and the
circumference of circle BGD lie on the surface of the sphere. From the vertex, A,
the line AZ is drawn as diameter of the sphere. One then imagines a plane, TZY,
tangent to the sphere at Z and one extends the surface of the cone beyond the base
BGD to meet the tangent plane TZY in a figure KLTN. Then KLTN is the cir-
cumference of a circle.8
Proof Let B be the point on the given circle nearest A and let BD be the diameter
of that circle, so D is the point on the given circle furthest from A. The circle
ABZH9 bisects both the sphere and the cone and its diameter is the diameter of the
sphere, AZ. Triangle ABD is in the common section of the cone and circle ABZH,
and let it be produced to the plane TQY. So line TY is the common section of that

7
As the figure shows, and the proof requires, BD, is a diameter of circle BGD in the sphere.
8
Since this is the content of Apollonius’s Conics, I, 5 it seems that al-Farghānī did not know of that
work.
9
Recall that circles are not curves but areas.
204 6 Spherical Trigonometry in the Islamic World

Fig. 9

plane and the tangent plane TZY. And let TK be the common section of the figure
KTLN and plane TZY.
Now, let L be an arbitrary point on the curve KTLN and M the midpoint of TK.
We shall show that ML = MK and, since L was an arbitrary point and the length
KM is fixed, this will show that KLTN is a circle.
In outline the proof goes as follows: Through L draw a plane, SLQ, parallel to
the plane of the given circle BGD. Now al-Farghānī cites a theorem of his con-
temporary, Muḥammad b. Mūsā, that says the intersection, SLQ, of the parallel
plane with the cone is a circle.10 Let the extension of D(ABD) in the cone cut circle
SLQ in SO. Also by Theorem 10 [FIX, since this refers to nothing that I can see
above.], SO is the diameter of circle SLQ. Let SO intersect KT at F.
Al-Farghānī now argues that since planes SLQ and KLTN are perpendicular to
AZT it follows that FL, which is part of their common section, is perpendicular to
AZT. Moreover, ∠AZT is right, since AZ is perpendicular to the tangent plane, YZT.
He now applies his first theorem to deduce that ∠ZTA is equal to ∠ADB. And,
by parallels, he concludes that ∠ADB is equal to ∠ASO. Hence, ∠ASO = ∠ZTA.
Since the vertical angles OFT and KMS are also equal it follows that triangle OFT
is similar to triangle KFS. From this it follows that FTFK = FOFS, which is equal
to FL2. And since the expression FTFK does not depend on the position of L on the
curve KLTN it follows that neither does FL. Hence KLTN is a circle.
One uses an astrolabe as follows. It is suspended from a cord attached to a ring,
such as that shown on the right in Plate 2. The cord is held by the hand so the

10
The result is Theorem 10 in On the Measurement of Plane and Solid Figures by the Banū Mūsā.
(Muḥammad was the oldest of the three brothers.) This theorem is found, phrased differently, in
Apollonius’s Conics I, 4.
4 Stereographic Projection and the Astrolabe 205

astrolabe hangs vertically. The ruler on the back (shown in Plate 3) pivots around the
central post and one may sight along it to a particular star and read the altitude of the
star from the scale of degrees around the rim. Suppose one finds the star Spica is 16°
above the horizon in the southwest. Rotate the star map until the pointer for Spica is
in the southwest, 16° above the horizon. (It will then be on the eighth almucantar of
the coordinate plate.) Now the astrolabe shows all stars in their correct positions, and
gives the altitude and azimuth of any star on the plate. It shows, in particular, which
stars are just rising/setting (i.e., have their pointers on the eastern/western horizon)
and which are below the horizon and therefore invisible.
The earliest surviving Arabic treatise on the use of the astrolabe is one written by
‛Alī b. ‛Isā, a scientist who was active in Baghdad around 830 and participated (as
we mentioned Chap. 1) in al-Ma’mūn’s survey to determine the circumference of
the earth. In addition, he took part in astronomical observations both in Baghdad
and in Damascus, so he must have had the experience necessary to write a treatise
dealing with a great variety of uses of the astrolabe—among them the following:
1. Determination of the longitude of the sun for a given date in the ecliptic.
2. Azimuth and altitude of any star.
3. Determination of ascendant, descendant, houses and other astrological uses (for
casting horoscopes).
4. Length of daylight or night, length of the unequal hours.
5. Time of day in equal or unequal hours.
We shall concern ourselves with the last of these.

5 Telling Time by Sun and Stars

To understand ‘Alī’s directions for telling the time of the night or day one must know
the two systems for recording time in the ancient or medieval world. The first system,
the popular one, divided each day and night into twelve equal parts, each of which was
one seasonal (or “temporal”) hour. Clearly the hours of the day would be longer in
summer and shorter in winter. Only on the equator, where all days and nights are equal,
would the hours not vary with the seasons. Elsewhere, it is only on the dates of the
spring and fall equinoxes, when the sun is on the equator, that the daytime hours are
equal to the nighttime hours. Thus, in the second system, where every day–night has
24 h of equal length, the hours are referred to as equinoctial hours. We shall say no
more about these, although one can also determine them readily with the astrolabe.
As for the problem of determining the seasonal hours Fig. 10 shows the celestial
_
sphere with a horizon ABG. On a given day of the year, the sun will describe a
_
circle parallel to the equator ADG. Let the part of the sun’s daycircle below the
_ _ _
horizon be ZEH , so that this is the path of the sun during the night. If ZE ¼ 12
1
ZEH
_
then the arc the sun travels in 121 the night is ZE . The corresponding point E on the
other arcs, one for each night of the year, forms a continuous curve on the sphere,
206 6 Spherical Trigonometry in the Islamic World

Fig. 10

and some of the Muslim astronomers realized that the curve is not a circle. Hence,
its image will not be a circle on the astrolabe. However, if we take the night arcs
only when the sun is on the Tropic of Capricorn, the Tropic of Cancer or the
equator, then we get only three points for each hour. The images of these three
points on the astrolabe determine the arc of the unique circle joining them, and it is
this arc that is labeled 1, with the successive arcs labeled 2, …, 11, and the eastern
horizon labeled 12, for when the sun is there it is the end of the twelfth hour of the
night, i.e., it is dawn. Of course, these circular arcs are only approximations to the
real curves, but most applications of mathematics demand some approximation, and
the circular arc is a reasonable one. This is how the lines for the night hours are
drawn on the astrolabe.
Now, if we wish to tell what hour of the night it is, we can proceed as follows.
Either from tables or from a scale on the back of the astrolabe, we find out where
the sun is in the zodiac on the day in question and mark that point on the zodiac
circle appearing on the star map. This sets the sun in the right position relative to the
stars for that day. Now find in the sky any of the 30 or so stars of those on the star
map, and use the alidade and scale along the rim of the back to take its altitude.
Note, also, whether it is to the east or west of the local meridian. Then set the star
map of the astrolabe so that the pointer corresponding to the star observed is on its
circle of altitude and east or west of the meridian as the case may be. Now the star
pointers are in the correct position relative to the horizon line, and so the sun,
correctly placed among the stars, is in a place faithfully representing its position in
the sky (of course below the horizon). All we need do now is look to see what
hour-line the sun is on (or near), and the number on that line tells us how many
hours of the night have passed.
5 Telling Time by Sun and Stars 207

We may now quote ‛Alī’s method for determining the time of day
The nadir is the degree (of the ecliptic) that lies exactly opposite the sun’s
position, the one seven signs after the one in which the sun is located. (‛Alī begins
counting “1” in the sign of the sun.) And one proceeds continuously until he finally
attains the seventh sign from his beginning point, which is then the position of the
nadir. Put the point corresponding to the sun on the altitude which you have found,
then look at the nadir (of this point), which falls on (or near) one of the hour lines. It
should be reckoned from the beginning point of the counting, and the point to
which you attain, on which the nadir falls, is the amount of hours and fractions
thereof past.
In order to shed some light on ‛Alī’s procedure, we have drawn in Fig. 11 the
equator and horizon, with the sun at a given point P on the ecliptic. We consider a
straight line through the center of the sphere and P. Since P and the center are on the
plane of the ecliptic, the point P*, where the line joining P and the center cuts the
sphere, is on the ecliptic and is the nadir. Now imagine P rotating in the direction
shown, parallel to the equator. As it does, P* rotates an equal amount in the same
direction. Moreover, because a great circle contains the diameter through any of its
points, P will be on a given great circle exactly when P* is. Our conclusion is that P
will sink below the horizon precisely when P* rises above the horizon, and that
whatever fraction of its daily circle, parallel to the equator, the sun, P, traces out to
reach the horizon, its nadir, P*, traces out precisely the same fraction of its parallel
circle in reaching the horizon. Thus, if P* falls on hour-line n, indicating 12-n hours
remain of the night, then, for P, 12-n hours remain of the day.

Fig. 11
208 6 Spherical Trigonometry in the Islamic World

Since the astrolabe is an analogue computer, which faithfully models circular


arcs and angles in the heavens, it can be used to solve any problem of spherical
astronomy. However, its accuracy is necessarily limited by the skill with which it
was fabricated, and it does not provide the most elegant solutions for all problems.
In fact no single method does (which is part of the charm of spherical astron-
omy), but spherical trigonometry, with its powerful and (often) easily stated rules,
was one source of several lovely solutions, and it is to the development of this
subject in the Islamic world that we now turn.

6 Spherical Trigonometry in Islam

There are three astronomers whose careers span the period during which most of
spherical trigonometry was developed and who, themselves, were responsible for
the most important results. The first of these is Ḥabash al-Ḥāsib, who was a con-
temporary of the great Arab scientist al-Kindī, and who was one of the astronomers
active in Baghdad under the patronage of the caliph al-Ma’mūn. The second is the
astronomer Abū al-Wafā’ al-Būzjānī, one of the ornaments of the Būyid court in the
middle and late tenth century, whose life and contributions to geometry and
trigonometry we have already recounted. The third is the prince Abū Naṣr Manṣūr
ibn ‛Irāq who was both teacher and patron of al-Bīrūnī in the latter part of the tenth
century. In his Keys to the Science of Astronomy al-Bīrūnī gives a lively account of
the controversies, misunderstandings, and accusations that accompanied the dis-
putes over priority in the discovery of some important theorems in spherical
trigonometry—particularly as they concerned the latter two astronomers mentioned
above.
In the course of his account, al-Bīrūnī has some unkind things to say of Abū
a1-Wafā’, but, despite al-Bīrūnī’s low estimate of Abū a1-Wafā’’s character, we
shall follow the proofs of two major theorems as Abū al-Wafā’ presented them in
his astronomical handbook Zīj al-Majisṭī.
The first of these results is the following “Rule of Four Quantities” (as it came to
be known later in the Latin West)
If ABG and ADE are two spherical triangles with right angles at B, D, respectively, and a
_ _ _ _
common acute angle at A then SinðBGÞ : SinðGAÞ ¼ SinðDE Þ : SinðEAÞ (Fig. 12).
_ _
Abū a1-Wafā’’s proof is as follows: Since the arcs AB and AG are arcs of great
circles the planes containing these arcs both contain the center of the sphere and so
intersect in a diameter d of the sphere. From G and E drop perpendicular straight lines,
_
inside the sphere, GH and ET, onto the plane containing AB, and in the plane con-
_
taining AG draw perpendicular straight lines GY and EK onto the diameter d. Then, it
is an easy consequence of Euclid’s Elements, XI, 11 that YH and KT are also
perpendicular to d. Thus, the angles GYH and EKT are equal, and so D(GHY) is
6 Spherical Trigonometry in Islam 209

Fig. 12

_
similar to D(ETK). Hence, TE:EK = HG:GY. But, EK ¼ SinðEAÞ and GY ¼
_ _ _
SinðGAÞ, while TE = SinðEDÞ and HG = SinðBGÞ, and substituting these into the
previous proportion yields the conclusion of the theorem.
An important application of the Rule of Four Quantities is found in Abū
al-Wafā’’s derivation of the Law of Sines for spherical triangles. Its discovery
simplified many problems concerned with arcs on the sphere and marked the full
emergence of spherical trigonometry because it was the first theorem to use spherical
angles. Other theorems used spherical triangles, but they dealt with only the sides.
Given the importance of the theorem it is not surprising that several authors
claimed credit for its discovery. Simultaneous discovery is not uncommon in
mathematics, since most problems and current methods are known to all workers,
and the case of the spherical law of sines seems to be another instance of it.
However, it appears that Abū al-Wafā’ was the first to publish it, in his Zīj
al-Majisṭī, and to use it, so perhaps the greater part of the credit for this important
advance goes to him.
The Law of Sines for spherical triangles says: If ABG is a spherical triangle with
sides a, b, g opposite the angles A, B, G then

SinðaÞ SinðbÞ SinðgÞ


¼ ¼ :
SinðAÞ SinðBÞ SinðGÞ

Abū a1-Wafā”s proof is as follows: In Fig. 13 let the spherical triangle,


210 6 Spherical Trigonometry in the Islamic World

Fig. 13

_
D(ABG), be given, and let GD be an arc of a great circle perpendicular to AB .
_ _ _ _ _ _
Extend AB and AG to AE and AZ , both of them quadrants, and extend BA to BH and
_ _ _
BG to BT , both quadrants. Then, A is a pole for the great circle EZ and B a pole for
_
the great circle TH Thus, by the second part of Basic Fact 2 in Sect. 1 the angles E,
H are both right, and triangles ADG and AEZ are spherical right triangles with a
common angle at B. Thus, by the rule of four quantities
_ _ _ _
SinðDGÞ SinðZE Þ SinðDGÞ SinðTH Þ
¼ _ and ¼ _ :
SinðbÞ SinðZAÞ SinðaÞ SinðTBÞ
_ _
However, as remarked earlier, A and B are poles of ZE and TH , respectively, so,
_ _
by definition of spherical angles, ZE ¼ ^A and TH ¼ ^B and so we may rewrite
the above equalities as
_ _
SinðDGÞ SinðAÞ SinðDGÞ SinðBÞ
¼ and ¼ :
SinðbÞ R SinðaÞ R
_
Eliminating SinðDGÞ and R from these two equations, we obtain Sin(a)/Sin(A) =
Sin(b)/Sin(B).
The proof of the remaining equality is entirely similar and the theorem is proved.
7 Tables for Spherical Astronomy 211

7 Tables for Spherical Astronomy

In the chapter on trigonometry, we mentioned that the auxiliary tables of trigono-


metric functions were computed by a series of astronomers from the ninth to the
fourteenth centuries, and we pointed out that one of the principal uses of such tables
was to aid in the computation of tables of functions important for spherical
astronomy. Our aim in this section is to provide details on these latter tables, for
they are one of the crowning achievements of numerical methods in medieval
mathematics.
A good example is the history of tables of oblique ascensions, which were
defined earlier in this chapter, and we shall begin with the tables of Ibn Yūnus. In
his Ḥākimī zīj this Egyptian astronomer tabulated these ascensions for each degree
of the ecliptic, and for each degree of latitude from 1° to 48°, to minutes of arc. This
means the computation of nearly 18,000 entries. D. A. King, who has investigated
these tables, reports that for the latitude of Cairo (30°) only one of the first 90
entries was in error by as much as 1 min. However, for the table as a whole, King
found that about one-third of the entries were incorrect by 1 min, and for a few
special cases, such as latitude 40°, the proportion of entries in error by 1 min
increased to two-thirds. There are also isolated errors of 2 or 3 min.
Two facts emerge from the information above: Ibn Yūnus performed an
impressive piece of numerical mathematics and, as the fair number of small but
significant errors makes clear; he used some mathematical methods to interpolate
between accurately computed values. However, as King’s analysis of the errors
makes plain, the interpolation procedure was not linear, so the tables of Ibn Yūnus
are evidence that in the Islamic world during the late tenth century there were
large-scale computational endeavors, using nontrivial mathematical formulas and
methods. (Apropos of the question why Ibn Yūnus stopped at latitude 48° King
conjectures that perhaps his feelings were the same as those expressed by Abū Naṣr
expressed in his auxiliary tables, namely, “I have made it (the table) for latitudes 1°
to 45° since amongst the inhabitants of the places whose latitude is greater than this
there is scarcely anybody who studies this sort of thing or even thinks about it.”)
Evidently, as the centuries passed, people of the northern latitudes became more
interested in such matters since Naṣīr al-Dīn al-Ṭūsī, in his Ilkhanī zīj, calculates the
oblique ascensions for each degree of the ecliptic, but this time for all latitudes from
1° to 53°, and not to minutes but to seconds. Then, some century and a half later
al-Kāshī calculated the oblique ascensions, again to seconds, up to latitude 75°, and
his patron, Ulūgh Beg, calculated these ascensions to latitude 50°, but to thirds. The
work of al-Kāshī and Ulūgh Beg exemplifies the quality of the best Muslim
achievements in the production of accurate, extensive scientific tables of functions
arising in spherical astronomy.
Our second example of tables of functions from spherical astronomy is the tables
for timekeeping. Ibn Yūnus compiled the first extensive collection of such tables for
the latitude of Cairo, for the purpose of determining the time by means of the sun
212 6 Spherical Trigonometry in the Islamic World

and stars for civil or astronomical uses, as well as regulating the times of the five
daily Muslim prayers. These times were defined in terms of the position of the sun
relative to the horizon, hence the composition of such tables is an exercise in
spherics applied to astronomy, i.e., in spherical astronomy. This science of time-
keeping (‛ilm al-miqāt in Arabic) gave rise to a group of astronomers who were
associated with major mosques and whose duty it was to tell the mueẓẓin when to
call the faithful to prayer.
Collections of such tables were quite large, that of Cairo containing 200 pages of
180 entries each. The following is a survey of the contents of this corpus as it is
found in the work attributed to Ibn Yūnus with the descriptive title Very useful
tables for finding the time since sunrise, the hour-angle and the azimuth of the sun
from the altitude of the sun. These tables fall into these main classes:
1. Auxiliary Tables of Functions for Spherical Astronomy
Among the 13 tables in this group are tables giving, for each degree of longitude
k of the sun, the declination of the sun, d(k), the length of daylight for the day
when the sun has longitude k, and the height of the sun when it is due south, due
east or due west.
2. Tables of Time Since Sunrise and the Hour Angle
These times are tabulated as functions of the longitude (on the ecliptic) of the
sun on the day in question and the altitude of the sun at the instant it is observed.
(One way to determine the altitude of the sun is by the sighting device on the
back of an astrolabe.) (See Plates 4 and 5 for an example of such tables.) In his
study of these tables, D. A. King points out that the often-repeated assertion that
Ibn Yūnus was the first to propound the so-called prosthaphairesis formula

1
cosðhÞ  cosðdÞ ¼  ½cosðh þ dÞ þ cosðh  dÞ
2

to facilitate computing the hour-angle from the altitude of the sun is a


misunderstanding by the nineteenth century French historian J.-B. Delambre.
3. Tables of the Azimuth of the Sun
These tabulate the azimuth as a function of the same arguments as the previous
tables used namely solar longitude and altitude. King remarks that these values
seldom deviate from the true values by more than 1 in the second digit, and they
are calculated for each degree of solar altitude up to a maximum of 83°. (This is
approximately the maximum for Cairo.)
4. Tables of the Altitude of the Sun for Certain Azimuths
These tabulate the altitude of the sun as a function of its azimuth and longitude—
which shows a certain excessive zeal for computing, since it is hard to think of
any purpose for which they would be really convenient, let alone necessary. Other
tables, very much in the same spirit as the foregoing, are
7 Tables for Spherical Astronomy 213

Plate 4 Two tables from the corpus of tables for timekeeping that were used in Cairo during the
medieval period. These display the time before midday when the Sun is in the direction of Mecca,
and the duration of evening twilight. Values are given in equatorial degrees and minutes for each
degree of solar longitude. (Taken from MS Dublin Chester Beatty 3673, fols. 8v–7r. Reproduced
courtesy of the Trustees of Chester Beattv Library, Dublin.)

5. Tables for Orienting Ventilators


Ventilators were used on the top of houses to draw cooling winds down into the
buildings, and certainly any book that could tell the residents of a city as hot as
Cairo how to do this could lay justifiable claim to the description “Very Useful.”
Indeed, the Iraqi traveler ‛Abd al-Laṭīf al-Baghdādī, who visited Egypt around
the year 1200, wrote of the Egyptians that they make the openings of their
houses exposed to the agreeable winds from the north. One sees hardly any
houses without ventilators. These ventilators are tall and wide, and open to
every action of the wind; they are erected carefully and with much skill. One can
pay between one hundred and five hundred dinars for a single ventilator, but
small ones for ordinary houses cost no more than one dinar each.
However, Ibn Yūnus’ Very Useful Tables did not tell the Cairenes in which
direction they should orient their ventilators. Rather, the book gave the altitude
of the sun when it was in the azimuth of the direction everyone used for
214 6 Spherical Trigonometry in the Islamic World

Plate 5 Extract from some anonymous fourteenth century (?) tables for timekeeping computed
for the latitude of Tunis, referred to as “Tunis, the protected (by Allah)” in the heading of the table
on the left. The tables display the time until noon as a function of solar meridian altitude and
instantaneous altitude. (Taken from MS Berlin Staatsbibliothek, (Ahlwardt 5754 fols. 23v–24r (We
1138).) Courtesy of the Staatsbibliothek Berlin

orienting ventilators. This is strange enough, but even stranger is the direction
the Cairenes used (and Ibn Yūnus prescribed) for orienting ventilators, namely
the direction of the rising sun at the winter solstice, which Ibn Yūnus calculated
to be 27°30΄ south of east, whereas modern data suggest that the optimal
alignment for the winds in Cairo is about 70° south of east. However, studies by
D. A. King have shown that in the medieval Islamic world a tradition of folk
astronomy associated the direction of the rising sun at the winter solstice was
associated with certain winds. Thus, this part of Ibn Yūnus’ treatise shows a
blend of folk astronomy with sophisticated calculations that is one of the
pleasures of this field of historical investigation.
6. Tables of the Duration of Morning and Evening Twilight
Twilight was defined in terms of the depression of the sun below the horizon,
and its determination is important because it defines times appropriate for
morning and evening prayers. Also, the use of these tables, in connection with
tables of the length of daylight, allowed the computation of the length of real
darkness for a given day, and Ibn Yūnus tabulates this function.
7 Tables for Spherical Astronomy 215

7. Tables for Afternoon Prayer


Although conventions on the time of afternoon prayer in the Islamic world
varied, Ibn Yūnus uses the convention that afternoon prayer begins at the time
after noon when the shadow of an upright rod in the ground equals the length of
its noon shadow plus the length of the rod. Ibn Yūnus tabulates, for each degree
of longitude of the sun, the altitude of the sun at the beginning of afternoon
prayer. (The permitted time for this prayer ends just before sunset.)
8. Tables of Corrections for Horizontal Refraction
In his Optics Ptolemy considered in a qualitative way the effects of atmospheric
refraction, especially at the horizon, but a table found in one of the manuscripts
of Ibn Yūnus’ Very Useful Tables makes it clear that Muslim astronomers of the
medieval period tried to get quantitative estimates of the effect of refraction. The
table under discussion applies to sunrise and sunset, but it appears in only one of
the manuscripts and contains blunders inconceivable in one as expert in
spherical astronomy as Ibn Yūnus. Thus, King feels the table was not composed
by Ibn Yūnus but was added, by a much less competent writer, on the basis of
remarks by Ibn Yūnus which he only half-understood.
The above, then, are descriptions of some of the tables found in the Very Useful
Tables, a corpus of tables that served Egyptian astronomers and timekeepers
until the nineteenth century.
In later centuries even more ambitious tables were undertaken. For example, in
1250 the Egyptian astronomer Najm al-Dīn al-Miṣrī tabulated the time since the
rising of a Star as a function of three quantities: (1) the maximum altitude of the
Star; (2) the instantaneous altitude of the Star; and (3) half of the arc of visibility.
These tables are computed for all declinations of the Star and for all terrestrial
latitudes and contain over a quarter of a million entries.
In the next century, in Damascus, the time-keeper at the Umayyad mosque,
Muḥammad al-Khalīlī, tabulated practically all the functions Ibn Yūnus tabu-
lated, but for the latitude of Damascus and for a different value of the obliquity
of the ecliptic. Perhaps it was the labor of doing all over again what Ibn Yūnus
had done for Cairo that inspired al-Khalīlī to compose the auxiliary tables we
mentioned in the chapter on trigonometry, tables which one could use to solve
the standard problems of spherical astronomy and which, therefore, would allow
the user to draw up a similar set of timekeeping tables for his latitude. In the
following section we shall see more of al-Khalīlī’s universal solutions to
mathematical problems arising within the context of Islam.11

11
Much more about the history of computed astronomical tables in medieval Islamic literature may
be found in King 2004–5.
216 6 Spherical Trigonometry in the Islamic World

8 The Islamic Dimension: The Direction of Prayer

The problem of finding the direction of Mecca relative to a given locality was a
product of the religion of Islam, for Mecca is the site of the Ka‛ba, the most sacred
spot in the Islamic world, and it is the direction to which Muslims must turn to say
their five daily prayers. This direction is called, in Arabic, al-qibla, and the problem
of its determination is an important one for Muslims. Accordingly, many of Islam’s
greatest scientists devoted some attention to its solution.
One of the greatest of these, al-Bīrūnī, wrote near the end of his definitive work
on mathematical geography, The Determination of the Coordinates of Cities, as
follows:
Though the determination of position is an end in itself, which satisfies an investigator, it is
our duty to find an application for such a determination that is beneficial to the populace of
the whole region whose longitude and latitude we have surveyed or to a particular section
of it exclusively. Let the universal benefit be the determination of the azimuth of the qibla.

It is typical of al-Bīrūnī’s tolerance for religions other than his own that he also
mentions the duty of the Jews to face Jerusalem and of the Christians to face east,
and he says his techniques will also be useful to them, and “I have no doubt that is
useful also to people of all faiths.”
In any case, al-Bīrūnī presents four methods for solving this problem. Although
a detailed exposition of one of these is beyond our purpose here, we shall describe
the problem and illustrate how it may be solved by spherical trigonometry.
Consider the problem on the earth’s surface. Figure 14 illustrates the situation
for a locality northwest of Mecca, where P is the North Pole, Z the locality in
question, M the position of Mecca, and WKFN the equator. (Since we shall con-
_
sider only arcs in what follows, and never straight lines, we will write XY for XY
without any ambiguity.) Thus PZ and PM are the local meridian and that of Mecca,
respectively, and ^PZM is the local azimuth of Mecca, which is the qibla.

Fig. 14
8 The Islamic Dimension: The Direction of Prayer 217

Moreover, ZM is the great circle distance (in degrees) from the locality to Mecca.
Since KZP is the local meridian, we will be facing north if we stand at Z and orient
ourselves so we face along ZP. Then, if we turn towards our right through angle
^PZM we will be facing Mecca, since ZM is the shortest route to that city. Hence,
to calculate the qibla of Mecca we must calculate ^PZM.
Clearly, if we are to find this angle we must know where we are and where
Mecca is, that is, we must know both our latitude, /, and that of Mecca, /M, as well
as the two longitudes, or at least their difference Dk (and whether Mecca is east or
west of the local meridian). From the latitudes we can determine their complements,
the arcs PZ and PM. Thus, if we are to find the qibla we must know those two arcs
as well as ^ZPM, that is, two sides and the included angle in D(ZPM). However, a
spherical triangle is determined by its two sides and the included angle, so these
data are sufficient to solve the problem.
It is apparent, however, that we cannot solve the problem by a single application
of the Sine Theorem to the spherical triangle PZM, for we do not know both an
angle and the side opposite to it. However, there is an approach that applies the Sine
Theorem to a series of spherical triangles. This was given by Ibn Yūnus without any
justification, but al-Bīrūnī both stated and justified it in his Mas‛udic Canon. It is
this account, as given in King (1986), which we shall follow.
First, however, the reader should recall that for any locality the altitude of the
visible pole (P) above the horizon is equal to /, the local latitude, and the distance,
measured in degrees, along the meridian between the zenith (Z) and the visible pole
is equal to /. These and other relations are illustrated in the circle of latitudes
shown in Fig. 15.
If we call the great circle whose pole is a point X “the horizon of X” then an easy
consequence of the definition of the poles of a great circle is that X is a point on the
horizon of Y if and only if Y is a point on the horizon of X. This principle has two
immediate consequences, which we leave as exercises: (P1) The horizons of two
non-antipodal points intersect in the poles of the great circle containing them; and
(P2) if Y lies on the horizon of X then X, its horizon and its antipode divide the
horizon of Y into four quadrants.
We now introduce al-Bīrūnī’s diagram for finding the qibla, as it appears in the
Mas‛udic Canon, although we have put the locality to the northwest rather than to
the northeast of Mecca. In Fig. 16, the circle KSN is the horizon of some locality
viewed from above, Z represents the local zenith, S the south and N the north, so
NZS is the local meridian. The point M is the zenith of Mecca, so that NK or
(equivalently) KS is the arc we need to find in order to know the qibla. Now, let
GFL be the horizon of Mecca, where F is an intersection of the horizon with the
local meridian, and let MHJ be the horizon of F. Finally, draw the great circle MPL,
where P is the north celestial pole.
Since M is a pole of GLJ, all three of ^MLG, MJ and ML are 90°, and since F is
a pole of MHJ both ^FHM and FH are 90°. Also PN = / ^MPH = D(k) and
218 6 Spherical Trigonometry in the Islamic World

Fig. 15

Fig. 16
8 The Islamic Dimension: The Direction of Prayer 219

PL = /M the latitude of Mecca, so MP = 90 – /M. By the Sine Theorem applied to


D(MPH)
sinðMPÞ sin ^FHM
¼ ;
sinðMHÞ sin ^MPH

which, using the above relations and the fact that sin h ¼ cosðhÞ, becomes,
cosð/M Þ sinð90 Þ
¼¼ :
cosðHJÞ sinðDkÞ

All quantities in this relation except cos(HJ) are known so cos(HJ), and hence
HJ, can be obtained. Then, ^F ¼ HJ is known and so is MH, which is equal to
90° − HJ.
Next, the Sine Theorem applied to D(PLF) yields

sinð^FÞ sinðPLÞ
¼
sinð^PLFÞ sinðPFÞ

and substituting the known values into this yields

sinð^FÞ sinð/M Þ
¼
sinð90 Þ sinðPFÞ

so that sin(PF), and hence PF, is known. Then, FN = / - PF is known and thus its
complement FZ is known.
Next, by the Rule of Four Quantities applied to D(FZI) and D(FHJ), we conclude
that
sinðFZÞ sinðFHÞ
¼
sinðZIÞ sinðHJÞ

which, on substituting in known values, becomes

sinðFZÞ sinð90 Þ
¼
sinðZIÞ sinðHJÞ

and, since all quantities except sin(ZI) are known we may determine ZI and, hence,
IQ = ZI. But, our principle (P1) applied to the horizons of M and Z implies that G is
a pole of KMZIQ, the altitude circle of Mecca, and hence that ^G = IQ, so ^G is
known.
Finally, the Sine Theorem applied to D(GFN) yields

sinð^GÞ sinðFNÞ
¼
sinð^FÞ sinðGNÞ
220 6 Spherical Trigonometry in the Islamic World

and therefore GN is known. But we have already remarked that GQ = 90°, so


NG = GQ is known. Since KS = NQ the qibla of Z is determined.
Al-Bīrūnī’s procedure is just one of the many solutions proposed to the problem
of finding the direction of prayer in Islam. Another solution involved composing
tables showing the qibla for a selection of localities. One such set of tables, with
2880 entries, was composed by the fourteenth century muwaqqit (time-keeper for a
mosque) Muḥammad al-Khalīlī, who we mentioned in the previous section. These
tables showed the direction of Mecca relative to the local horizon for each locality
of latitude 10°, 11°, …, 56° and 3312 (the latitude of Damascus) and for each
longitude east or west of Mecca by 1°, 2°, …, 60°. Solutions included methods of
approximation, descriptive geometry, solid geometry, trigonometry and the con-
struction of sets of tables ranging from a few dozen entries to many thousands. The
survival of this multitude of methods over hundreds of years suggests that the
history of ancient and medieval spherics is not one of steady intellectual ascent in
which superior innovations replace outdated methods. It is, rather, the story of the
development of a variety of techniques to the point where each is able to solve the
problems currently of interest. In this it seems typical of the history of mathematics.

Exercises

1. Show that if two points on the surface of the sphere are not diametrically
opposite then there is a unique great circle containing them, but that, when the
two points are diametrically opposite, there are infinitely many such great
circles.
2. Show that the intersection of a plane with a sphere is either a single point or a
circle on the sphere.
3. Show that any parallel circle is parallel to a unique great circle.
4. Show that any two great circles on a given sphere bisect each other.
_ _ _
5. Suppose in the figure of the complete quadrilateral that arcs AG, GD and AB are
_ _
quadrants, and that AE = 60° and DZ = 45°. Use Menelaos’s Theorem to
_
calculate AD.
6. Use Menelaos’s Theorem, and the fact that the angle e between the equator and
the ecliptic is approximately 2312 , to find d, the height of the sun above the
equator, on the day when it is at a point on the ecliptic 45° south (measured
along the ecliptic) of where the ecliptic and equator intersect. According to
Ptolemy’s Table of Chords, Crd(47°) = 47; 51. (Hint: In Fig. 2 take BA to be
the equator, BE the ecliptic, G the North Pole and Z the Sun.)
7. Let P be the vertex of a given cone and p and p′ parallel planes that do not
contain P. Then, p intersects the cone in a circle if and only if p′ does. Prove
this.
8. Show by a simple proof that stereographic projection maps any circle parallel to
the equator onto a circle.
Exercises 221

9. In Fig. 5 show that CD = (Cos(e)  NS) and that C′D′ = CD=[1 + Sin(e)]. Use
these results to construct the image, under stereographic projection, of the
parallel circle of latitude d.
10. In Fig, 5 take A΄B΄ as the X-axis and O as the origin. Show that if a great circle
whose diameter is XY is inclined to the equator at an angle of h then its
stereographic image is a circle whose diameter is tan(45° – h/2) tan(45° + h/2)
and whose center is at the point

        
 h 1  h  h
tan 45   tan 45  þ tan 45 þ ;h :
2 2 2 2

11. With reference to Fig. 5, calculate the center and diameter of a circle whose
_
diameter ZW is parallel to XY so that WY ¼ b.
12. Use the results of the above problems to construct the lines on the plate of an
astrolabe for your latitude that shows the equator and the two tropics, your
horizon, and almucantars for every 6°.
13. From the web, find the length of daylight for a particular day at your locality
and then calculate the times, according to your clock, at which the (seasonal)
hours of the night begin.
14. If you have access to an astrolabe try to calculate the length of the seasonal
hours of the day. (You will have to know the date and will have to use the back
of the astrolabe to find the position of the Sun in the ecliptic for the given date.)
15. If you have access to an astrolabe use it to find the declination of the Sun for a
given date. Describe the method you use.
16. Use the Rule of Four Quantities and the data in Exercise 6 to compute the
declination of the Sun.
17. Show that Menelaos’s Theorem implies the Rule of Four Quantities, which we
know implies the Sine Theorem. Show that the Sine Theorem implies
Menelaos’s Theorem and conclude that the three results are equivalent.
18. Solve Exercise 6 using the Sine Theorem.
19. Show that if one knows the qibla for a locality then the Sine Theorem may be
used to obtain the great circle distance between the locality and Mecca.
20. Prove (P1) and (P2) of Sect. 8 from the principle that X is on the horizon of Y if
and only if Y is on the horizon of X and the Basic Facts of Sect. 1.

Bibliography

Berggren, J.L. 1980. “A Comparison of Four Analemmas for Determining the Azimuth of the
Qibla”. Journal for the History of Arabic Science 4(1): 69–80.
Berggren, J.L. 1987. “Spherical Trigonometry in the Zīj of Kūshyār ibn Labbān”. In Festschrift: A
volume of studies of the history of science in the near east, Dedicated to E. S. Kennedy, ed.
222 6 Spherical Trigonometry in the Islamic World

D. A. King and G. Saliba. New York: New York Academy of Sciences. (An English translation
of the section in the zīj dealing with spherical trigonometry).
Bīrūnī, Abū al-Rayḥān. 1967. The determination of the coordinates of cities (transl: J. Ali). Beirut:
American University of Beirut.
Debarnot, M-Th. 1978. “Introduction du Triangle Polaire par Abū Naṣr b. ‘Irāq”. Journal for the
History of Arabic Science 2(1): 126–136.
Hilbert, D., and S. Cohn-Vossen. 1952. Geometry and the Imagination. New York: Chelsea.
Kennedy, E.S. 1973. A commentary upon Bīrūnī’s Kitāb Taḥdīd al-Amākin. Beirut: American
University of Beirut (Commentary on Bīrūnī’s The Determination, etc.).
King, D.A. 1973a. “Al-Khalīlī’s auxiliary tables for solving problems of spherical astronomy”.
Journal for the History of Astronomy 4: 99–110.
King, D.A. 1973. “Ibn Yūnus’ Very Useful Table for Reckoning Time from the Sun”. Archive for
History of Exact Sciences 10 (1973), 342–394 (The fundamental study of the Cairo Corpus of
tables for astronomical timekeeping.).
King, D.A. 1986. Article “Ḳibla”. In The Encyclopædia of Islam (2nd edition, Vol. V, p. 86).
Leyden: Brill.
King, D.A. 2004. In synchrony with the heavens: Studies in astronomical timekeeping and
instrumentation in medieval islamic civilization (Vol. One (2004) and Vol. 2 (2005)). Leyden:
Brill.
Lorch, R. 1980. Al-Khāzinī’s ‘Sphere that Rotates by Itself’. Journal for the History of Arabic
Science 4(2): 287–329.
Lorch, R. 2005. Al-Farghānī on the Astrolabe: Arabic text edited with translation and
commentary. Stuttgart: Franz Steiner Verlag.
MacAlister, Paul R., and Flolydia M. Etting (designers). The Astrolabe Kit with The Astrolabe:
some notes, etc. by R. S. Webster. Lake Bluff, IL: Paul MacAlister.
Pepper, J.V. 1968. “Harriot’s calculation of the meridional parts of logarithmic tangents”. Archive
for History of Exact Sciences 4(5): 359–413.
Van Brummelen, G. 2013. Heavenly mathematics: The forgotten art of spherical trigonometry.
Princeton: Princeton University Press.
Chapter 7
Number Theory and Combinatorics
in the Islamic World

1 Number Theory

Number theory has a rich ancient tradition, much of it being found in Books VII–IX
of Euclid’s Elements. Among the beautiful results in these three books, one finds a
proof that there are infinitely many prime numbers, and that if 2n − 1 is a prime
then 2n−1(2n − 1) is a perfect number, i.e., is equal to the sum of its proper divi-
sors.1 Moreover, in Book X, one finds a rule for generating squares of whole
numbers whose sum is also a square.
Not too long after Euclid, Eratosthenes developed his famous sieve for finding
the primes in a sequence of the first n integers, and, some centuries later,
Diophantos, among other investigations, solved the problem of finding two rational
numbers such that when either is added to the square of the other the result is the
square of a rational number. And in the same work, the Arithmetica,2 he sets forth
an algebraic method of finding rational solutions for indeterminate equations, i.e.,
equations such as x2 + y2 = z2 that have more than one solution. Both the
Arithmetica and the Elements were well known in medieval Islam.3
At some point between the time of Euclid and Diophantos (probably around
100 A.D), Nicomachos of Gerasa wrote his Introduction to Arithmetic in which he
discussed, among other topics, figured numbers (see below) and gave the fourth
perfect number, 8128.4
One of the earliest results concerning the number theory that was proved in
medieval Islam was that of Thābit b. Qurra who translated both Euclid and
Nicomachos and proved a sufficient condition for two whole numbers to be

1
The first two perfect numbers are 6 and 28.
2
At that time in the Greek world, “arithmetic” (arithmētikē) referred to what we call “the theory of
numbers.” Our “arithmetic” the Greeks called “logistic” (logistikē).
3
However, of the 13 books of Diophantus’s Arithmetica only six books survive in Greek and
another four those only in Arabic.
4
The earliest mention of the fifth perfect number, in the 15th century, is 33,550,336.

© Springer Science+Business Media New York 2016 223


J.L. Berggren, Episodes in the Mathematics of Medieval Islam,
DOI 10.1007/978-1-4939-3780-6_7
224 7 Number Theory and Combinatorics in the Islamic World

amicable. This notion, which seems to be an extension of the idea of perfect


number, went back to the ancient Greeks who called two numbers amicable if each
is equal to the sum of the proper divisors of the other. (The numbers 220 and 284
were standard ancient examples of such numbers.) Thābit’s theorem states that: If p,
q, and r are primes such that p = 3  2n − 1 − 1, q = 3  2n − 1, and r = 9 
22n − 1 − 1 then 2npq and 2nr are amicable numbers.

1.1 Representing Rational Numbers as Sums of Squares

We mentioned, above, that Diophantos gave a method for finding rational solutions,
x, y, and z, to the equation x2 + y2 = z2. In the thirteenth-century, in II, 1 of his
Algebra, Ibn al-Bannā’ deals with the problem of finding rational solutions to the
equation x2 + y2 = z2. This number theoretic material occurs in the context of
finding solutions to problems of three types: (1) dividing 10 into two parts satis-
fying certain conditions, (2) dividing money among a certain number of men so that
certain conditions are satisfied, and (3) a certain sum of money is increased and
decreased according to certain conditions. But, to prepare his reader to solve such
problems he states the following rules:
1. If a and b are any two numbers5 such that a/b = ¾ then a2 + b2 is a square of a
rational number.
2. If a is a square it may be expressed as the sum of two squares. His abbreviated
demonstration of this is: “This is because there exist two squares whose square
is a square. The given square is then decomposed according to their ratio.”
3. If a is not a square, then if there exist whole numbers x and y satisfying a =
x2 + y2 there also exists another, different, pair of numbers, w and z, such that
a = w 2 + z 2.
4. He concludes with the following procedure for deciding whether a whole
number can be expressed as the sum of two squares:
You may know whether it has two square parts by subtracting from it the first of the natural
squares, i.e. ‘one.’ And if the difference has a [whole number] root [then you have
expressed it as the sum of two squares. But if not, one subtracts the second square, which is
‘four’ and one examines the remainder And one proceeds step-by-step in this fashion.
If it is one of those numbers that cannot be expressed as the sum of two squares this will
become evident with whole numbers. For if it cannot be decomposed into two whole
number squares neither can it be decomposed into squares of fractions. Keep this in mind.

Only in Rule 3 does he refer specifically to whole numbers, so one assumes that
when not further specified “number” refers to rational numbers in general.

Although this is true for any two rational numbers Ibn al-Bannā’ simply says “any two numbers.”
5
1 Number Theory 225

In summary, Rule 1 tells how to generate as many pairs of rational numbers as


one likes whose squares sum to a square of a rational number, and Ibn al-Bannā’
doubtless realized that it would apply to all pairs of rationals m and n whose ratio is
that of the sides of any right triangle with rational sides, p and q.6 In Rule 2, he hints
at how a given rational square can be written as the sum of two rational squares.7 In
Rule 3, he addresses the case of non-square rationals for which one has, somehow,
found an expression as a sum of squares. And he claims that in this case one can
find a different expression as well.
And Rule 4 gives an inductive procedure for finding when a whole number is a
sum of two squares.8 And, as a sort of coda to this rule he adds that if the procedure
shows that the number is not a sum of two whole number squares then neither is it
the sum of two rational squares. (One can only speculate how Ibn al-Bannā’
convinced himself of the truth of this fact. It is unlikely that he had a proof of this
latter interesting result, for, centuries later, Fermat stated the result but gave no
proof. In fact, it was not proved until the early twentieth century.)
As an application of these rules, consider his Problem 3 in the text following the
rules. This requires that one divides 10 into two squares. He first gives the integer
decomposition 10 = 1 + 9, an application of the inductive approach described in
Rule 4.
By Rule 3, there must exist another expression for 10 as a sum of two squares of
rational numbers, and these are clearly not integers. To find these two squares, one
multiplies 10 = 1 + 9 by some square, say 25, to obtain 250 = 25 + 225, a
decomposition into terms whose ratio is the same as the 1/9 ratio for 10. The
inductive approach of Rule 4 now yields another decomposition of 250 as a sum of
two squares, namely 250 = 81 + 169 = 92 + 132. One can now reverse the above
process of multiplying by the square, 25, to divide by that square to obtain
10 = 250/25 = 81/25 + 169/25 = (9/5)2 + (13/5)2, and this solves the problem.
In remarks following Problem 3 Ibn al-Bannā’ assures the reader that if one
chooses a square multiplier (such as 25) that does not work then some other square
multiplier will.
He also gives an algebraic approach and says that “If you want, you can set one
of the square roots you are looking for equal to a thing plus a side of one of the two
squares making up 10, so, say, if 10 = a2 + b2 then a = 1 + x. Then you set the
other square root equal to a some things less the side of the other square,

6
Briefly, the argument would be that since p, q and r are rational numbers such that p2 + q2 = r2,
and if m, n are rationals such that p/q = m/n then, on squaring both sides of the proportion and
adding 1 to the squares on each side, one sees that n2r2/q2 = m2 + n2, i.e. m2 + n2 is also the
square of a rational.
7
For example, because 9 + 16 = 25, if one wants to express 36 as a sum of two rational squares
one can multiply all terms by 36/25 and write (9/25)  36 + (16/25)  36 = 36. And each of the
summands on the left is a square of a rational number.
8
It was Fermat who discovered, and proved, that a whole number greater than 1 can be written as
the sum of two squares exactly when all primes congruent to 3 (mod 4) in its prime factorization
appear to an even power.
226 7 Number Theory and Combinatorics in the Islamic World

so b = nx − 3. Say n = 3.” Ibn al-Bannā’ then expands a2 + b2 = 10 to obtain


10x = 16. Substituting this value into the expressions for a and b one obtains
a = 2 3/5 and b = 1 + 4/5, and the sum of their squares is 10.
He also remarks, that, because of the choices one can make, the problem is
“indeterminate,” i.e., it has many solutions. And he points out that if, in the alge-
braic approach, one had chosen n = 2 instead of n = 3 one would have obtained the
decomposition 10 = 1 + 9 which, he says, “the algebraists call a vicious circle and
is of no interest.”
Finally, he points out that one would get the same result if one chose d = x − 3.
And if one chose d = rx − 3, where r < 1, one would get an impossibility. He ends
by enjoining the reader to “know this, think about it and understand the difference
between a vicious circle and an impossibility!”

1.2 Figured Numbers

The problem of summing series of whole numbers has long been a topic of
mathematical investigation and a number of ancient civilizations have found sums
of arithmetic series. Both Nicomachos, in what is now Jordan, and Āryabhaṭa, in
fifth century India, knew that the sum of the first n cubes is the square of the sum of
the first n whole numbers. And both of these wrote about the sums of the first
k whole numbers, k = 1, . . . , n, yielding the sequence 1, 3, 6, 10, etc. The
Pythagoreans represented those partial sums by pebbles or other markers arranged
to form triangles with an equal number of markers on each side and called them
triangular numbers. They also represented the numbers in other sequences by
polygons as well and called them square numbers, pentagonal numbers, etc.
(Fig. 1).
Thus, from the triangles they obtained 1, 3, 6, 10, . . .; from the squares 1, 4, 9,
16, . . . and from the pentagons the numbers 1, 5, 12, 22, . . . .
In general, such sequences of numbers are called figured numbers and
Nicomachos, in his Introduction to Arithmetic, explained how one could generate

Fig. 1
1 Number Theory 227

all figured numbers by starting with triangular numbers.9 Thus, two successive
triangular numbers create a square number, as one can see by dividing the 4 x 4
pattern above by a line running just above the main diagonal, into the two con-
secutive triangular numbers 10 and 6.
Such numbers arise naturally in summing the terms of an arithmetic series of
constant difference d, i.e., a series 1, 1 + d, 1 + 2d, . . . , 1 + nd, . . . . When d = 1,
2, and 3, one obtains (respectively) the arithmetic sequences

1, 2, 3, 4, etc.—the whole numbers


1, 3, 5, 7, etc.—the odd numbers
1, 4, 7, 10, etc.—
and the sums of these sequences produce triangular, square and pentagonal num-
bers, and so on.
1, 3, 6, 10, etc.
1, 4, 9, 16, etc.
1, 5, 12, 22, etc.
Thābit b. Qurra, who, as we mentioned above, translated Nicomachos’s
Introduction into Arabic, introduced readers in both the eastern and western parts of
medieval Islam to this work. Thus, it is no surprise that mathematical interest in
building sequences of integers on analogy with geometric figures continued with
the mathematicians in medieval Islam, and in the early eleventh century ‘Abd
al-Qāhir al-Baghdadī (d. 1037), stated rules for general terms of series of figured
numbers as well as for the general terms of the series whose terms are sums of the
figured numbers of a certain type. The tradition of study of this topic was still active
in the thirteenth century Maghrib when Ibn Mun‘im devoted the ninth ‘species’ of
his Fiqh al-ḥisāb (Laws of Calculation) to the following exposition of the topic.10
In this section, Ibn Mun‘im presents the following table showing, for integers
k = 1, . . . , 10 (listed in “Row of the Sides”) and each figured number with n sides,
from n = 3, . . . , 10 (listed in “Row of the triangles” to “Row of the Decagons”),
the number of points Fn(k) in the figured number with n sides and k points on each
side. For example, the number F5(4) in the column headed four in the row of the
pentagons, says that the pentagonal figure with four points on each side has 22
points in all.
Since, for n = 3, the triangular figure with k points on a side contains
k + (k − 1) + . . . + 2 + 1 points F3(k) = k + F3(k − 1). Ibn Mun‘im goes
on to state that an analogous result holds for Fn for n > 3, namely that
Fn(k) = Fn−1(k) + F3(k − 1). For example, if one looks at column 7 in the row of

9
Nicomachos, in his Introduction to Arithmetic, referred to these as ‘polygonal’ numbers. He also
stated—relative to our earlier reference to perfect numbers—that the fourth perfect number is
8,128.
10
See Djebbar, A. “L’analyse combinatoire au Maghreb.” Also the same author’s “Figurate
Numbers in the Mathematical Tradition of Andalusia and the Maghrib.” (Neither of these has been
published.) The present author also thanks Prof. Driss Lamrabet for sending him his doctoral
thesis, containing the edited Arabic text of Ibn Mun‘im’s Fiqh al-ḥisāb.
228 7 Number Theory and Combinatorics in the Islamic World

Row of sides 1 2 3 4 5 6 7 8 9 10
Row of triangles 1 3 6 10 15 21 28 36 45 55
Row of squares 1 4 9 16 25 36 49 64 81 100
Row of pentagons 1 5 12 22 35 51 70 92 117 145
Row of hexagons 1 6 15 28 45 66 91 120 153 190
Row of heptagons 1 7 18 34 55 81 112 148 189 235
Row of octagons 1 8 21 40 65 96 133 176 225 280
Row of enneagons 1 9 24 46 75 111 154 204 261 325
Row of decagons 1 10 27 52 85 126 175 232 297 370

Fig. 2

the squares one sees that F4(7) = 49 = 28 + 21 = F3(7) + F3(6), and, in the row of
the hexagons, F6(7) = 91 = 70 + 21 = F5(7) + F3(6). Thus, one can build up the
table recursively from an initial row of the sequence of whole numbers.
Ibn Mun‘im then points out that if one asks, for any species and any length of
side how many points that figure contains, there is a way of expressing the number
of points using only triangular numbers. For example: How many points are there
in the octagonal number of side 4? The answer, F8(4), is found in the table in Fig. 2.
And, as Ibn Mun‘im has already shown, it is generated by working one’s way up
the column from F8(4), always adding in the constant difference, F3(3), to obtain
F8(4) = F7(4) + F3(3) = F6(4) + 2F3(3) = F5(4) + 3F3(3) = F4(4) + 4F3(3) = F3(4) +
5F3(3). One now has only triangular numbers and the sum is 10 + 5  6, i.e., 40, in
agreement with what one finds in the table.
He also shows how the same idea can, with the aid of algebra, be used to solve
the converse problem, namely find the side of the octagonal number with 40 points.
His approach is to let the side be the unknown, in modern symbolism, x. Then, with
the same idea, one knows F8(x) = 40. So we know, by the previous paragraph, that
40 = F3(x) + 5F3(3). But F3(x) = x + (x − 1) + . . . + 2 + 1 which, ibn Mun‘im
2 2
tells his reader, is x2 þ 2x. To this one adds 5  F3(x − 1), which is 5  (x2  2x) and
obtains 3x2  2x ¼ 40. Ibn Mun‘im expresses this by what we would write as
x2 ¼ 23 x þ 13 13, which he solves by methods similar to those found in Propositions II,
5 & 6 of Euclid’s Elements.
(Ibn Mun‘im went beyond these figures into the third dimension by introducing
pyramidal numbers by studying the sums of figures of a given species. However, it
is not a topic we shall explore here.)

1.3 Magic Squares

To conclude our section on number theory, we consider the problem of magic


squares. These square arrays of consecutive whole numbers arranged so the sums of
the numbers in each of the rows, columns and two main diagonals are always the
1 Number Theory 229

same,11 were known in several ancient civilizations, and the 3  3 square of 9 cells
is found in Chinese and Indian writings prior to the Islamic era. Below is the earliest
magic square known, from a Chinese treatise written in the first century of our era.

2 9 4

7 5 3

6 1 8

However, available evidence supports the view that it is in the Arabic literature
that one finds the first systematic investigation of methods for constructing magic
squares of arbitrary order.12 Thābit b. Qurra, whom we have mentioned a number of
times, is the first known to have written on the topic, although his treatise has not
come to light. Others continued Thābit’s investigations and by the end of the tenth
century mathematicians in medieval Islam had worked out most of the techniques
for generating a number of different types of squares of different orders. One of the
most interesting treatises on these squares was written by a mathematician who we
have mentioned a number of times in our book, Abū al-Wafā’.13
But before we begin our account of his treatise, we state a few mathematical
preliminaries. The common value of the sums of the rows, columns and diagonals of
a magic square is called the magic constant so if—as is often the case—the numbers
in the square are taken to be the first n2 positive integers the magic constant must be n
(n2 + 1)/2. In the Chinese example given above, where n = 3, the magic constant is
15. In the case of an odd number, n, Abū al-Wafā’ expressed this formula for the
magic number as “we take the middle number and multiply it by the side of the
square.” So, in the example given above, the magic constant of the 3 x 3 magic
square is 15. (It is easy to see that there is no magic square of order 2.14)
Among the several types of magic squares is one square called a bordered
square, a magic square in which removing successive borders always leaves a
magic square, such as in the 5 x 5 magic square below, where the removal of the
border leaves a 3 x 3 magic square,15 as in the square below. (It is worth noting that

11
The consecutive whole numbers used need not begin with “1.” It was probably not until late in
the twelfth century that authors in the Islamic world began claiming magical powers for these
arrays when engraved on amulets, etc. Prior to that they had always been simply mathematically
interesting subjects of study and were referred to as ‘harmonious arrangements of numbers.’
12
The number of rows of a magic square is called its order.
13
Prof. Sesiano (1988) published an edition and French translation of Abū al-Wafā’’s treatise,
based on the text of the only known manuscript, which is in the Süleymaniye Library in Istanbul.
The author has used this publication and Prof. Sesiano’s insights into the work in the present
exposition.
14
This is because for any square array of four different whole numbers to be magic the sum of any
two of these integers must be the same. And that is impossible.
15
Since there is no magic square of order 2 there cannot be a bordered square of order 4.
230 7 Number Theory and Combinatorics in the Islamic World

bordered squares were the first for which investigators discovered a general method
of construction.16)

6 8 23 24 4

7 10 17 12 19

5 15 13 11 21

25 14 9 16 1

22 18 3 2 20

As the reader has seen earlier, Abū al-Wafā’ (940–997/8) was an excellent
mathematician and astronomer, but—apart from this—his treatise on magic squares
is of special interest because it is one of the two earliest surviving Arabic treatises
on the topic. (And the other one was written by a contemporary of his, ‘Alī
al-Anṭākī.) However, the latter contents himself with providing directions for
forming various kinds of magic squares while Abū al-Wafā’, in Sesiano’s words,
“presents as much as he can of the theory of the construction of magic squares,
beginning with examples of low order to discover the general rules.”
Abū al-Wafā’’s exposition discusses, first, squares of odd order and then even
order. We follow his exposition of the former type, beginning with squares of order
3. (He points out, as we did above, that the magic constant for this square, when the
cells are occupied by the integers 1, . . . , 9 is 15.)
He begins with the statement that the middle number in the sequence 1, . . . , 9 is
5 and that must be the number in the central cell. He gives no proof of this, although
he does point out that if we just arrange the integers in order, starting with the top
left cell and working to the bottom right then 5 will be in the central cell and the
diagonals will sum to the magic constant.17
Next, he claims that none of the four corner cells can contain an odd integer. There
are five odd integers in the interval from 1 to 9, one of which occupies the center. If there
were an odd integer in a corner, say the upper left corner,18 then the third element in the

16
See Sesiano, p. 121.
17
In a footnote Sesiano gives the following proof. If M is the magic constant for the square and we
denote by c the number in the central cell then c is the middle number in both the middle row and
middle column as well as in both diagonals. Together these contain all elements of the square,
other than c, once, and c four times. Hence, we may express the sum of the elements in the square
as 4(M − c) + c. But this sum is also equal to the sum of all the numbers in the square, i.e.
3M. Hence, 4(M − c) + c = 3M. Since M = 15 it follows that c = 5.
18
In the remainder of this argument we shall often appeal (explicitly or tacitly) to the reader’s sense
of symmetry. Abū al-Wafā’ makes no reference to this concept, but—fairly clearly—is depending
on it.
1 Number Theory 231

main diagonal would have to be odd.19 We now have three odd integers on the main
diagonal and there are only two odd integers left to put in the six empty cells.
If those two odd integers were both put above (or below) the main diagonal then,
since there are only even integers left, one row or column sum has to be even,
which is impossible. It is equally easy to see that neither one of the remaining odd
integers be placed above the main diagonal and the other below it.
Hence, the four even integers must occupy the four corner positions and so, by
symmetry, we may assume we have 2—5—8 in the main diagonal and 6—5—4 in
the off-diagonal. It is now trivial to place the remaining two odd numbers and, so, to
complete the magic square as in the Chinese example.
In light of what is to come, it is worth remarking here that if we add one and the
same number, n, to each of the nine cells of a 3 x 3 magic square we will still have
a magic square of order 3.
Now, Abū al-Wafā’ shows how to build up bordered squares of odd order from a
square of order 3, and he begins by stating the general identity, a2 − b2 = (a + b)
(a − b) and applying it to the case of a = n + 1 and b = n − 1 to obtain
(n + 1)2 − (n − 1)2 = 2(2n).20 In the case when b = 3 and a = 5, this implies the
difference of the squares of these numbers is equal to 2(3 + 5), namely 16. This
implies, in particular, that the border of a square of order 5 will have 16 cells.
To build a square of order 5 one starts with the 3 x 3 square based on 1, . . . , 9. Then
one takes the middle number in the series 1, . . . , 25, namely 13 and, in analogy to what
we did with the square of order 3 we would like to begin with 13 in the middle of the
5 x 5 square. Since number 5 occupies the middle place and we want 13 there, we
simply add 8 to each cell of the 3 x 3 square. The result is the magic square below.

10 17 12

15 13 11

14 9 16

Now, of the sequence 1, . . . , 25 one has used the nine middle numbers, 9, . . . , 17.
The border of this square will have 16 elements and, indeed, of the original sequence
there remain 16 elements not yet used. We arrange these in two sequences of 8
elements each, as follows:

1 2 3 4 5 6 7 8
25 24 23 22 21 20 19 18

19
This is because the sum of the other two elements is even and the sum of all three is 15.
20
As Abū al-Wafā’ puts it, the difference of the two (consecutive odd/even) squares is equal to
twice the sum of their sides.
232 7 Number Theory and Combinatorics in the Islamic World

We shall refer to the numbers in the top row as the “lesser” numbers and those in
the bottom row as the “greater.”
One notices that the two elements in each column sum to 26, and we refer to them as
“complementary.” Now, the magic constant of the order 3 square we constructed above
is 15 + 3  8, namely 39. And, as it happens, 39 + 26 = 65, which is the magic
constant for the 5  5 square. Hence, the bordering cells must be filled in so that
complementary numbers are opposite each other. So, if one places 1-2-3 in that order
down the left side of the 3  3 their complementary numbers, 25-24-23, must be
opposite them down the right hand side. And the same considerations hold for the
numbers in the four corners. The problem with the example we have just given is, of
course, that the sum just three elements in the rightmost column, namely 25 + 24 + 23
(=72), is already larger than the magic number, 65. So, one must be a little careful.
Abū al-Wafā’ now does as he did with the 3  3 square once he had placed 5 in
its middle square, i.e., he begins by filling in the two upper corner cells of the
5  5, and, following what worked with the 3  3 square, he uses the two numbers
that are on either side of the order of the square. In the 3  3 it was 2 and 4, so in
the case of 5  5 it is 4 and 6. The complements of these must go diagonally
opposite them, so he now has the partial square:

6 4

10 17 12

15 13 11

14 9 16

22 20

We begin by filling in the three cells of the first row, which must sum to 55 (i.e.,
65 − 10). When we examine the array of 16 numbers we find 23, 24, and 8, which
we put into the top three empty cells, and we put their complements opposite them
in the bottom row. We now have

6 8 23 24 4

10 17 12

15 13 11

14 9 16

22 18 3 2 20
1 Number Theory 233

Since the two numbers in the rightmost column add to 24 we must find three
numbers, among the six remaining, which add up to 41. We easily see that the only
choices are 19, 21, and 1, which we may distribute in any one of six ways in the
three empty cells on the right.
However we do it, we write their complements in the opposite cells, on the left,
and we come up with a magic square of order 5:

6 8 23 24 4

7 10 17 12 19

5 15 13 11 21

10 14 9 16 1

22 18 3 2 20

Apropos of his choice of 6 and 4, which led to a successful completion of the


square, Abū al-Wafā’ states explicitly that they were not chosen by chance. Indeed,
he names a number of possible choices, such as 6 and 7, that would not have
allowed one to complete the square and he lists a number of other pairs, namely
(1, 3), (3, 7), (5, 7), (2, 8), and (6, 8) that do allow completion of the square.
One notes that all of these numbers come from the eight lesser numbers. In fact,
there is no loss of generality in restricting oneself to these since, as Sesiano
remarks,21 the four corners contain two pairs of complements and so there must be
two consecutive corner squares that contain lesser numbers. By symmetry we may
assume they are the upper corners.
Abū al-Wafā’ gives no arguments to justify his conclusions and he probably
considered all 28 possible cases to find the six that yielded magic squares.22 To see
what is involved in eliminating a case let us take a pair not on his list of possible
pairs, say (1, 7). With this choice we would have the following as a partial square23:

21
Sesiano, p. 128.
22
It is worth noting that Abū al-Wafā’ wrote a book on the techniques of mental arithmetic used by
merchants and government officials. And, as the reader will see from the example we work out,
there is an easy procedure, with nothing involving more than additions and subtractions of small
numbers.
23
We do not indicate the numbers in the central 3 x 3 square, since they do not enter into the
following argument.
234 7 Number Theory and Combinatorics in the Islamic World

1 7

10 17 12

15 13 11

14 9 16

19 25
To fill in the top row, we must choose three integers from the remaining 12
(=16 − 4) possible entries. The largest of the lesser entries is 8. Since 57 − 8 = 49
and the sum of the two largest of the greater entries is 23 + 24 (=47) it must be that
all three entries come from the remaining greater numbers, i.e., 18, 20–24. But,
47 − 18 = 39 and any pair of integers between 20 and 24 sums to more than 40.
Hence, there is no way to complete the square.
On the other hand, if one chooses the pair (3, 7) for the top corners one quickly
finds that the first row can be filled in by 8—22—25, and completing the rest of the
square is easy.

2 Combinatorics

A number of ancient cultures were engaged in counting various combinations and


permutations of objects. In India writers counted the number of possible patterns of
long and short syllables in Sanskrit poetry and computed the number of combi-
nations of n things (e.g., six flavors) taken k at a time. In Greece in the second
century B.C. Hipparchos, calculated the number of possible “connections” in a
certain system of logic as being 103,049.24
Similar interests can be found in the early centuries of Islamic civilization. For
example, the eighth century alchemist, Jābir b. Ḥayyān, from Kufa in modern-day
Iraq, enumerated the number of possible combinations of seven degrees of each of
the four elements of ancient natural philosophy (heat, cold, wetness, and dryness).
Also, writers on music theory enumerated the ways in which three tones could be
arranged to create melodies.
In the ninth century Thābit b. Qurra, whose name has appeared frequently in this
book, enumerated the 36 ways in which the terms appearing in the ratios in

24
This is one of several results that Plutarch attributes to Hipparchos. No one knew quite what to
make of this particular number until 1994 when David Hough noticed that this number is what is
known as the tenth Schröder Number, named for the logician, E. Schröder. These numbers count
the number of ways in which a product of ten symbols can be bracketed. For example, the product
of a,b,c can be bracketed in exactly three ways—((a)(b)(c)), (((a)(b))(c)) and ((a)((b)(c)))—and the
product of four symbols can be bracketed in eleven ways.
2 Combinatorics 235

Menelaos’s Theorem could be permuted and still preserve the truth of the theorem.
(As a good mathematician, however, Thābit was not content to exhibit different
ways of writing the given rule. He also gave a neat argument to show that there
were no other ways.) Perhaps a bit more than a century later, al-Bīrūnī gave a
systematic enumeration of all possible solvable cases of spherical triangles.
Despite these achievements, there were no general rules that would apply to
different sorts of problems. And it was first in the Maghrib and al-Andalus, beginning
in the thirteenth century that we find writers creating a mathematical theory that
could solve a wide variety of combinatorial problems in the form of general rules
stating the number of combinations and permutations satisfying certain conditions.

2.1 Enumerating Words of k Distinct Letters in an Alphabet


of n Letters

The triangular numbers, which we discussed above, appear also in Ibn al-Bannā’’s
Raising the Veil in his solution to the problem of calculating the number of k-letter
words (without repetitions) that could be formed from an alphabet of n letters. For
Ibn al-Bannā’, in this problem, “word” is simply a set of distinct letters and his task
is to count the number of ways of choosing a set of k distinct letters from an
alphabet of n letters. (As with any set, it is only the members and not the order in
which they are listed that make a difference.)
He begins by stating the general rule, which we state in modern terms as:
To find the number of words with k letters from an alphabet of n letters, multiply the
number of combinations preceding the desired combination [i.e. the number of words with
k − 1 letters] by the number preceding the given number [n] whose distance from it is equal
to the number of combinations sought [i.e. n − k + 1],25 and of that result one takes the part
given by the number of the sought combinations [i.e. the kth part].

To justify this rule, Ibn al-Bannā’ begins with k = 2 and simply says that the
number of pairs one can form from n letters is the sum of the numbers from 1 to
n − 1. He gives no argument for this, so evidently
 the reader is assumed to know
the result stated and that the sum is n  2 . He now turns to justifying the rule for
n1

the remaining cases.


In the case of combinations of three letters he argues that any choice of three of
the n letters, say {a, b, c}, contains three distinct pairs of two letters. Taking any
one of these three pairs and combining it with the remaining letter will produce the
same set of three letters. Thus, one takes the number of pairs of n letters, namely
2 , multiplies it by (n − 2), and divides the product by 3. The result is one
n  n1
form of the usual formula for C(n, 3).

25
For example, the distance of n − 1 from n is 2, the distance of n − 2 is 3, etc. So the count
begins with n and ends when, working downwards, one has counted to k.
236 7 Number Theory and Combinatorics in the Islamic World

The same sort of reasoning shows that the number of combinations of four letters
from an alphabet of n is the usual formula for C(n, 4), simply because any given
choice of four letters, say {a, b, c, d}, can arise from any of its triples, say {a, c, d},
by adding the remaining letter (in this case b). Hence it can arise from four different
triples, and, on dividing by 4, one then has C(n, 4).
At this point, Ibn al-Bannā’ feels, the reader will have seen the pattern
and restates the rule in a way such that one does not have to calculate all the
possible combinations of, say, from 2 to 10 letters to find the number of combi-
nations of 11 letters:
If a set of numbers is given, and we wish to know the number of combinations from these
having a given number of elements, we take the product of the numbers of the sequence of
integers starting from the greatest term, which is the number of elements of this set [from
which we are choosing the combinations], and take the number of terms equal to the
number of elements of the combination. Further, we put as a divisor of this product the
successive numbers of which the greatest term is the given number [of letters in each
combination] and which begins with one and with two. Then we cancel the common terms
between the first numbers and the second; and when we do that, all the second numbers are
cancelled; then you multiply the remaining first numbers together and you get the com-
binations from this set of numbers.

He gives the example of 28-choose-5 (28 being, as we said, the number of letters
in the Arabic alphabet) and arrives at the answer 5  24  13  9  7, which is
correct.26

2.2 Ibn Mun‘im on Counting Arabic Words of at Most Ten


Letters

One Muslim scholar who made some outstanding contributions to combinatorics


was Aḥmad b. Mun‘im, who was originally from al-Andalus but later worked in
Marrakesh and died there in 1228 A.D. (We have already mentioned his work on
number theory.)
In Chapter 11 of his Laws of Calculation (Fiqh al-ḥisāb) ibn Mun‘im poses, the
problem of counting all possible words in Arabic having at most ten-letters. The
Arabic alphabet has 28 letters and some very specific rules about what sort of words
are possible. (A trivial example of such rules, in the case of English, would be that
the letter q must be followed by the letter u.) Ibn Mun‘im says that the Arabic rules
of word formation, the number of letters in the alphabet, as well as his limit of ten
for the number of letters in a word, are simply details, specific to the language, and
the problem, and “our goal. . . is the description of a method by which it is possible
to enumerate words,” and if he has made certain assumptions specific to Arabic “it
is [only] to illustrate the procedure of the method we propose.”

26
However, taken literally his rule would lead to the (equal) expression 28  27  26  5. Perhaps
he felt the student might want to have as few of the successive factors as possible larger than 20.
2 Combinatorics 237

Fig. 3

He begins with a problem which appears to have little to do with his main
problem, and that is to calculate of the number of ways of making a tassel of
p (different) colors from silk threads of n different colors.
He begins by constructing a grid labeled as below, where the columns corre-
spond to the number, p, of colors available and the rows to the number, n, of
distinct colors in the tassel. Obviously, p  n, and if n = p then there is only one
possible tassel, which explains the “1”s in the diagonal of the square ascending
from the lower right corner (Fig. 3).
The rows are counted from bottom to top and the columns in a given row are
counted, beginning with the rightmost “1” of each row from right to the left. Since
the bottom row simply lists the colors (first, second, etc.) available we shall begin
the row count with the row above that, i.e., the row with all “1”s. (Thus, row 1 has
11 columns (entries 1, . . . , 1, 10), row 2 has 10 columns (entries 1, 2, 3, . . . , 9, 45),
etc.) And, in the following argument it will be easiest if we begin the column count
in row 1 with “1,” that in row 2 with 2, etc. So, in any given row, n, the entry in
column p (>n) is the number of new tassels, each having threads of n distinct colors,
you can make if you have p rather than only p − 1 colors.
The “1”s in the first row mean that if the tassels can have one thread then you
can only construct one more tassel each time you gain a color. And the 10 at the
extreme left in row 1 means that the total number of one-color tassels you can make
from ten colors is ten.
238 7 Number Theory and Combinatorics in the Islamic World

So, the ones in the first row are obvious enough, but, as for the entries in row
two, Ibn Mun‘im explains that if you want tassels with two colors you must have at
least two colors, with which you can make only one tassel. Hence the “1” in column
2 of row 2. But, if you have a third color available then you can make two more
tassels, each having that third color with one of the previous two colors—hence the
“2” in column 3 of row 2. Then, if you have a fourth color, you can make three new
tassels, each combining the fourth color with one of the 3 previous colors. And so
you fill in the second row right to left, as 1, 2, . . . , 9 and, at the extreme left of the
row, write the total number of choices of pairs from 10 colors by adding up the new
tassels gained as you increase the number of threads, the sum being 45.
Ibn Mun‘im takes next the case of tassels with three colors, so in the case of
tassels with three threads of distinct colors, starting with the minimum number of
colors possible, namely 3, yields only one possible tassel. If you add a fourth color
then each new 3-thread tassel must use the fourth color and will produce, if we
remove that color, a 2-thread tassel from the first three colors. Different 3-thread
tassels, each using the fourth color, must produce different 2-thread tassels using the
first three colors, and conversely. Hence, the number of new 3-thread tassels must
be just the total number of 2-thread tassels using three colors, i.e., the sum of the
numbers in the first two cells of the second row.
Ibn Mun‘im extends his detailed analysis of the generation of new tassels when
another color is added only up through the fourth row and then says, “One proceeds
in the same way for the construction of the fifth row beginning with the fourth in the
same way one did for the construction of the fourth row from the third.”
He then states the general rule to the effect that the number of tassels having
n threads chosen from p different colors, p  n, is the number in column p + 1 of
row n + 1, i.e., the sum of the entries in the first p columns of row n. In symbols, if
an+1, p+1 is the number in column p + 1 of row n + 1 then an+1,p+1 = an,p +
an,p−1 + . . . + an,n, from which it immediately follows that an,p = an,p−1 + an−1,p−1.
This is, of course, the rule that generates the entries of Pascal’s triangle of
binomial coefficients, which we saw in Chaps. 1 and 4. One sees, on reading the
columns up from the bottom, the familiar sequences 1—2—1, 1—3—3—1, etc.
And this cannot be much of a surprise, since one learns early on in mathematics that
the number of combinations of n distinct objects from a set of p distinct objects—
exactly Ibn Mun‘im’s problem—is the binomial coefficient C(p,n).
Ibn Mun‘im had clearly realized the many beautiful properties of this table since
he writes, “If you think about the particularities of this table and about that what
appears here as a surprising harmony it will show you extraordinary symmetries
whose discussion would take a long time.”
In Problem 2, Ibn Mun‘im sets the task of counting the number of permutations
of the letters of a word in which no letter is repeated. The answers, 1 and 2, for
words of one or two letters are obvious. For the case of three letters, he goes back to
the two possibilities for words of two letters, XY or YX. For each of the two words
the third letter, Z, can be placed in one of three places. The result is six possibilities.
Proceeding in this manner he shows that in the case of five letters there are 120
possibilities.
2 Combinatorics 239

He then states the general rule for calculating n! as:


If you have a word of which the number of letters is known, and of which no letter is
repeated, and you want to know the number of permutations of the letters of this word you
multiply one by two, and the result by three, then this result by four, then this result by five,
and so on, each result being multiplied by the next number in the sequence of whole
numbers, until one reaches the product by the number equal to the number of letters in the
word.

He then comes to Problem 3, namely: How many permutations are there of a


word with a known number of letters in which one, two or more letters are repeated
a known number of times?27 His rule, which he expresses without symbolism, is
that if the word has n letters and only one letter is repeated, say k times, you first
suppose the letters are all different, so by the above result there would be n!
permutations. Divide this number by the number of permutations of the letters of a
word having k distinct letters, namely k!, to obtain the answer n!/k!.
To justify this he says that if only one letter is repeated (say k times) in an n-letter
word then, to each position of the repeated letters there would correspond, if they
were different, k! permutations of the n-letter word. He then gives the general rule for
n-letter words in which k letters are repeated, each of the k letters being repeated
mi times, i = 1, . . . , k. The rule is to divide n! by the product m1!m2!. . . mk!.
The proof of this is, he says, analogous to the argument in the case of only one letter
being repeated.
Ibn Mun‘im now poses the fourth problem, which is to find out “the number of
configurations of a [single] word in which the number of letters is known, taking
account of the vowels and sukūns28 that succeed one another over the letters, but no
account of the permutations of the letters.”
In case the word has only one letter, say w, then there are only three possible
configurations (wa, wi, wu) since there are only three vowels in Arabic to go over
any consonant. (That only one of these three possibilities, wa (= and), is an Arabic
word (“and”) is of no concern to Ibn Mun‘im, since the other two certainly could be
words.
In the case of two letters there are, he tells us, 12 possible configurations because
there can be any one of three vowels above the first letter and any of three vowels or
a sukūn above the second letter, an example being lam (= not), which has the vowel
a over the first and a sukūn over the second.
In the case of three letters there are the 12 possibilities for the first two and 4 for
the third, i.e., 48 in all. However, a sukūn over each of the second and third is
impossible, and there are three cases for this, namely when the first letter has the
vowel, u, a, or i. So, for the case of three letters there are 45 possibilities.

27
It is to be understood that different letters may each be repeated a different number of times.
28
The sukūn is a sign that looks like a very small ‘o’ and is written over a consonant to indicate that
no vowel follows that consonant. For example, the Arabic word for ‘key”, which is miftāḥ, would
have a sukūn over the ‘f’.’ Two consecutive consonants cannot each have a sukūn, and a word
cannot begin with a letter that has a sukūn over it.
240 7 Number Theory and Combinatorics in the Islamic World

In the case of four letters, one takes all possibilities for the first three letters, 45,
and multiplies it by the four possibilities for the last letter to obtain 180 possibilities.
Again, however, since a word cannot end in two sukūns one must subtract the
number of cases in which that happens, which is the number of possible configu-
rations for a two-letter word not ending in a sukūn, namely 3  3, i.e., 9. The
number of possible configurations for a given four-letter word is, thus, 171.
The case of four letters is the first in which the general rule becomes apparent so
Ibn Mun‘im works through the case of five letters as well, applies the same pro-
cedure for 6, and then states the general rule: “Subtract three from the number of
letters of the word and take the number of possible configurations for such a word,
taking account of vowels and sukūns. You multiply this by three and you keep this
as ‘the first result.’ Then you subtract one from the number of letters of the word
and you also multiply the number of configurations of the remainder by four. From
this product you subtract the first result. The remainder is equal to the number of
configurations of the letters of the word, taking account of vowels and sukūns, but
not the permutations of the letters of the word.”
Ibn Mun‘im’s rule is, thus, that if W(n) is the number of possible configurations
of a word of n letters then W(n) = 4 W(n − 1) − 3 W(n − 3). This rule, of course,
holds only for n > 3, and Ibn Mun‘im reminds the reader of the results for the first 3
cases. He ends his discussion of the problem with a table of 10 entries for words
with up to 10 letters, the last entry being 507,627 configurations of a ten-letter
word, taking account only of vowels and sukūns.
At the end of the discussion of this problem, he gives another rule for computing
W(n), which is W(n) = 3 W(n − 2) + 3 W(n − 1), a rule which holds for n = 3 as
well.
Having presented the table for W(n), n = 1. . . . , 10, he now connects the two,
seemingly disparate, problems: counting tassels and counting words. To see the
connection, he says, think of the letters of the alphabet as colors, so in the case of
the Arabic alphabet one has 28 colors. A word of n letters corresponds to a choice
of n threads of the 28 colors, each color being repeated as many times as the letter
corresponding to it in the word.
However, as is usual in mathematics, he begins with a simple case, here a
three-letter word in which no letter is repeated, which corresponds to a tassel with
three colors and three threads.29 In this case, he says, set the numbers 1, 2, 3 in a
line and take the products, so you get six. (This is, of course, the number of ways of
arranging three different letters into three-letter words.) Then, he says, multiply this
product by the number of configurations of a three-letter word according to vowels
and sukūns, which by the argument in Problem 4 is 12. The result, then, is that
given three distinct letters one can form 72 possible Arabic words.
This basic idea, clearly generalizes to an arbitrary number (n) of distinct letters,
so in general the number of words that can be formed from a given an n-letter word

29
There will be three threads because the word has three letters. There will be three colors because
no letter of the word repeats.
2 Combinatorics 241

by permutation and arrangements of vowels and sukūns is n!  (the number of


arrangements of vowels and sukūns).
To deal with the case of words in which certain letters are repeated a certain
number of times, he thinks of tassels in which threads of certain colors appear more
than once. He takes the case of tassels with eight threads of five colors; in which
each of three pairs of threads have a different color and each of the remaining two
threads have a different color. If we call the colors Ci, i = 1, . . . , 5, we may
represent this case by the following table:

C1 C2 C3 C4 C5
a a b b b

The letter ‘a’ represents single threads (in this case one thread of Color 1 and
another of Color 2. Each ‘b’ represents a pair of threads, in this case one of three
monocolored pairs. One now asks, how many permutations of the a’s and b’s would
result in different numbers of threads having a given color. For example, inter-
changing the two a’s or permuting the three b’s among themselves would not
change the appearance of the tassel. But interchanging one of the a’s with one of the
b’s would since it would result in the multiplicities of two different colors in the
tassel changing, although still preserving the condition the tassel must satisfy.
Thus, the enumeration in this case is accomplished by the rule stated in Problem
3. In this case, the rule gives 5!/(2!  3!), i.e., 10. And, when one keeps in mind the
correspondence between letters of a word and colors of a tassel one sees that there
are 10 possible permutations of the letters of an eight-letter word made up of five
different letters in which each of two letters appear only once and the other three
letters appear twice each.
In Problem 7, he takes the case of a 9-letter word of which five letters are
distinct, two of them not repeated, two each appearing twice and one appearing
three times. One creates the following table:

a b c d f
c d f
f

and underneath each column with the same number of letters one writes the same
letter, so:

a b c d f
c d f
f
a a d d r
242 7 Number Theory and Combinatorics in the Islamic World

One then has a combination30 of 5 letters of which two are repeated twice. These
may be permuted in 5!/(2!  2!), i.e., 30 different ways. And that is the answer to the
question posed in Problem 7.
In Problem 8, he now returns to the original problem, dealing with words of
anywhere from two to ten letters, and he lists all possibilities in which one or more
letters are repeated
Nine distinct letters and the tenth repeats one of these, or
Eight are distinct, the ninth and the tenth repeat just one letter among the eight or
two, or
Seven are distinct and the three remaining letters repeat one letter of the seven,
or two, or three, or
Six are distinct and four repeat just one letter, or even two. And if they repeat
two letters, this will either be once for one of them and three times for the second, or
twice for the one and twice for the other. Or [the four] repeat three letters and this
will be twice for one and once for each of the other two [of the three repeated]
letters. [Or, they repeat four letters, once each.]
And, for the sake of completeness, he concludes with the obvious:
If there is only one letter then all ten letters are the same.
In Problem 9, he begins with the first of the cases listed in Problem 8, namely: If
the tenth letter repeats one of the nine others, there will be, for the ten letters, nine
combinations, each [having] ten letters of which one letter is repeated twice.
In the case when the ninth and tenth repeat one and the same letter of the first
eight, there are obviously eight combinations, but if each repeats a different letter he
gives the answer first, 28 combinations, and then explains that to find this number
one goes back to the problem of the tassels. The case under discussion corresponds
to a tassel of six threads of different colors, and two pairs of threads—each of a
single different color. This is solved by the method of the tableaus and leads to the
following question: “How many permutations are there of the combination a a a a a
a b b?” the answer being 8!/(6!2!).
Ibn Mun‘im spares his reader the excruciating details of the calculations and
simply remarks, “In this way one determines the number of the species of com-
binations, the procedure having been already explained.”
With these preliminaries out of the way, Ibn Mun‘im returns to his main problem
and remarks that if all the letters of the word are distinct he has already solved the
problem. For the other cases, he will first discuss the case of 10-letter words, then
nine-letter words, and so on down to two-letter words.
In the first case then, one considers ten letter words made of nine distinct letters,
so one of the nine is repeated. Ibn Mun‘im reverts now to the problem of tassels
with ten threads of nine different colors chosen from threads of 28 colors.
The procedure is as follows:

30
Ibn Mun‘im uses this term to refer to an unordered list of letters, with repetitions allowed.
2 Combinatorics 243

First, given 28 colors of silk, we want to make tassels with ten threads of nine
colors, one color being repeated. How many such tassels are possible? Find this
number and keep it as the first number.
Now, given ten letters, nine distinct and one of them repeated twice, how many
combinations are possible for the ten letters. This is the second number; in this case 9.
Determine the number of permutations of a word of 10 letters, one of them
repeated twice. This is the third number.
Now, determine the number of configurations of a word with ten letters
according to vowels and sukūns. This is the fourth number. (Ibn Mun‘im points out
later that this number depends only on the number of letters in the word, so he
recommends saving it for use in all computations involving words of a fixed length,
“so you don’t get tired.”)
The product of these four numbers is the “number of words such that one cannot
pronounce a word of ten letters, of which one is repeated twice, without it being one
of them.”
He then goes through the same procedure in detail for both subcases of the case
of ten-letter words in which eight letters are distinct, and then closes saying that all
the other cases are the same.
The reader may have noticed that to get the “first number” in the procedure one
needs a table of the number of combinations of p things taken n at a time, for p = 28
and n = 1, . . . , 10 (in the case of the limits on word length Ibn Mun‘im has
imposed). It appears that Ibn Mun‘im initially omitted this and some other tables
due to lack of time. However, he tells us in a postscript to his treatise that when he
once again had time, but after the treatise had been copied and given to students, he
composed the necessary tables and included them with the work.
He works out a few examples for his readers, the first being the case of 10-letter
words with no repetitions. Since there are no repetitions the product will only
involve three factors. The first number, C10 28
, he calculates as equal to 13,123,110.
The second number, which in this case is 10!, he calculates as 3,628,800, and the
third, which counts the configurations of any 10 letter word from the point of view
of vowels and sukūns one can calculate as 2.41738  1019.

2.3 Ibn al-Majdī on Enumerating Polynomial Equations

Another counting problem that interested the Muslim authors was that of counting
the number of possible polynomial equations of a given degree in one variable.
Such a count was obviously important to anyone wanting to give a complete
discussion of solving equations of a given degree. Al-Khwarizmi, in the ninth
century, makes a point of saying that there are six different equations of degree at
most two and he is going to deal with all of them. Omar Khayyam, in his twelfth
century Algebra lists the 25 possible forms of equations of degree at most three.
244 7 Number Theory and Combinatorics in the Islamic World

Shihāb al-Dīn b. al-Majdī, who was responsible for regulating the times of
prayer at Cairo’s great al-Azhar mosque, died in 1447 A.D. He wrote a work called
Enveloping the Core, a commentary on the Summary of Ibn al-Bannā’, and in this
work he wanted to show that there was no limit to the number of polynomial
equations.31 To do so, he showed how to count the number of possible equations of
degree at most some given whole number. We, of course, have the so-called
canonical form of an equation of degree, say, four, namely ax4 + bx3 + cx2 +
dx + e = 0, but the canonical forms for equations of any degree in medieval Islam
all had to have positive coefficients so that what we could write as 5x4 − 2x3 −
8x2 + 2x − 7 = 0 would be written as 5x4 + 2x = 2x3 + 8x2 + 7.
The individual summands of these equations al-Majdī called “species” and he
counted the number of equations whose degree was less than or equal to a given
number by breaking the possible equations into cases, e.g., if there are five species
one case is when the sum of two species is equal to the sum of three other species
(the case listed above). Al-Majdī referred to the above case by what we shall denote
as (2, 3) and that would be different from the case (1, 2) or (1, 4).
For our above example of equations with at most five species, the possible cases
would correspond to the pairs (1, 1), (1, 2), (1, 3), (1, 4), (2, 2), and (2, 3).
He first gives the results of his enumeration of the various cases for equations of
degrees at most 2, 3, and 4. The totals of the various cases for each of these three
degrees, he tells his reader, are 6, 25, and 90. These numbers alone would have
convinced most of his readers that, as he claims, the numbers of equations of a
given degree grow without bound.
However, he then turns to stating general results. If there are n species then the
number of possible equations of type (1, 1) is n(n − 1)/2. To find the number of
equations of type (1, 2) multiply the number of type (1, 1) by n − 2.
For equations with four species the additional types are (1, 3) and (2, 2). In the
first case you take the product 13, which you remember, and then multiply the
number of trinomials by n − 3. The quotient of that product, divided by the number
you were to remember, is the number of quadrinomials of type (1, 3). For the other
type, (2, 2), multiply 22, remember it, multiply, as before, the number of trinomials
by n − 3. The quotient of that product, divided by the number you were to
remember is the number of type (2, 2).
Al-Majdī does not give proofs of his results, but one can understand his rea-
soning in the case of four species as follows: Any expression using four species
must come from a trinomial, all of which are of the type (1, 2). Now, a quadri-
nomial can come from adding the fourth species to the one sum appearing in the
trinomial, and, since there is only the one possible fourth species to add, the number
of quadrinomials of type (1, 3) is T, the number of trinomials. However, the three
terms in the sum could have come from a trinomial in any one of three ways, so we

31
I thank Prof. A. Djebbar for sending me the Arabic text of this part of the work and, also, Prof.
M. Bagheri for his help in understanding the treatise while we were waiting for a delayed flight
from Tehran to Isfahan.
2 Combinatorics 245

must divide T by 3 to get the number of quadrinomials of type (1, 3). As for those
of type (2, 2), they come from trinomials also, but by adding the fourth term to the
isolated term of the trinomial. Again, for any trinomial there is only one choice, but
the result could have come about in 22 different ways (two sums and two sum-
mands in each), so there are T/4 possible quadrinomials of type (2, 2).
So, his rule, in brief, is: If E4 is the number of equations formed from 4 distinct
species then E4 = B + T + T/3 + T/4, where B is the number of binomials formed
from the four species and T the number of trinomials.
Al-Majdī concludes with equations involving at most five species. If one of the
species is not present then one applies the rules given above for computing the
numbers of binomials, trinomials and quadrinomials that can be formed from four
species. Equations using all five species must come from quadrinomials, which are
either of the type (1, 3) or (2, 2). So, one multiplies the number of quadrinomials of
type (1, 3) by n − 4 and divides the result by 4. For those arising from type (2,2) we
obtain 2/3 of the second type of quadrinomials.
One has, expressing this rule as a formula, E5 = B + T + Q(1/3 + ¼) + Q(1, 3)/4 +
[Q(2, 2)2]/3, where Q(1, 3) and Q(2, 2) denote the numbers of quadrinomials of
types (1, 3) and (2, 2), respectively, and Q denotes the number of all quadrinomials.
The proof, which the author does not give, is similar to the case of quadrinomials.

Bibliography

Djebbar, A. 1981. “Ensignment et recherche mathématiques dans le Maghreb des XIIIe – XIVe
siècles (étude partielle).” Publications mathématiques d’Orsay, no. 81-02.
Lamrabet, D., ed. 2005. Fiqh al-Ḥisāb (The science of arithmetic). By Aḥmad Ibn-Mun‘im., Ph.D.
Thesis, University of Rabat: Faculté des sciences de l’éducation.
Sesiano, J. 1988. “Le traité d’ Abū’l-Wafā’ sur les carrés magiques”. Zeitschrift für Geschichte der
arabisch-islamischen Wissenschaften 12: 121–244.
Sesiano, J. 1982. Books IV to VII of Diophantus’ Arithmetica in the Arabic Translation Attributed
to Qusṭā ibn Lūqā. New York, Heidelberg, Berlin: Springer.
Index

A Aḥmadi ibn Mun‘im Fiqh al-hisāb, 227


Aaboe, A., 164 al-‛Azīz, 180
‛Abbāsids, 2 al-Andalus, 2, 7, 100, 111, 235, 236
‘Abd al-Qāhir al-Baghdadī, 227. See also Abū al-Andalūsī, Ṣā‘id, 4
Mansur al-Baghdādī, ‛Abd al-Laṭīf, 213
‘Abd al-Raḥmān, 2, 4 al-Baghdādī, Abū Manṣūr, 46, 74
‘Abd al-Raḥmān II, 4 al-Baghdādī, Abū Manṣūr Arithmetic, 74
‘Abd al-Raḥmān III, 4, 7 al-Battanī, 176
Abū al–‘Abbās al-Farghānī, 202 al-Bayhaqī, 14, 88, 147
Abū al-‛Abbās Ma’mūn 10, 12 al-Bīrūnī, Abū a1-Rayḥān, 7, 11
Abū al-Ḥasan al-Marrakūshī/Science of al-Bīrūnī Chronology, 11
Timekeeping , 164 al-Bīrūnī Coordinates of Cities, 9, 173, 216
Abū al-Jūd, 88, 119 al-Bīrūnī Gems, 12
Abū al-Wafā’, 11, 62, 107–111, 123, 166, 168, al-Bīrūnī India, 12, 173
183, 208, 209, 229–232 al-Bīrūnī Keys to Astronomy, 208
Abū al-Wafā’ Cube and Fourth Roots, 62 al-Bīrūnī Mas‛ūdic Canon, 178
Abū al-Wafā’ Geometry Needed by Craftsmen, al-Bīrūnī Pharmacology, 12
107 Aleppo, 106
Abū al-Wafā Zīj al-Majisṭī, 208 Alexander the Great, 121
Abū Bakr al-Ḥaṣṣār, 38, 143 Alexandria, 5, 9, 92, 104, 155, 198
Abū Kāmil, 9, 130–133, 135, 153 al-Farabī, 106, 107
Abū Kāmil Algebra 108 ff., 130 al-Farabī Music, 106
Abū Naṣr, 177, 208, 211 al-Farabī Spiritual Crafts, 106
Abū Sahl al-Kūhī, 79, 86–89, 91 al-Fazārī, 2, 123
Abū Sahl al-Kūhī Complete Compass, 88, 91 Algebra, 8, 14, 23, 57, 71, 82, 121, 123, 124,
Abū Sahl al-Kūhī Regular Pentagon, 79, 109, 130, 133, 134, 143, 145, 147, 151, 152, 228
183 arithmetization, 133, 134
Abū Sahl al-Kūhī Trisection of Angle, 92, 94 Egyptian, 7, 9, 130, 132
Abū Ṭāhir, 14, 147 symbolism, 146
Addition, 8, 32, 34, 166 ‛Umar’s view of, 151
Addition, sexagesimal, 50 Algeria, 151
Adelard of Bath, 164 Algorithm, 8, 22, 34, 36, 43, 52, 57, 61, 62, 68,
‛Aḍud al-Daula, 88 74, 123, 143, 185, 187
Afghanistan, 2, 12, 13, 25 Algorithm, Euclidean, 123
Aghlabids/ Ibrāhīm II, 4 al-Ḥajjāj ibn Maṭar, 6
Aḥmad al-Qatrawānī, 142 al-Hakam II, 113
Aḥmad ibn al-Qunfūdh, 145 al-Ḥākim, 7, 180
Aḥmad ibn Mūsā, 6 Alhazen. See Ibn al-Haytham

© Springer Science+Business Media New York 2016 247


J.L. Berggren, Episodes in the Mathematics of Medieval Islam,
DOI 10.1007/978-1-4939-3780-6
248 Index

Alhazen’s Problem, 99, 111 Altitude (of a star), 193, 199


‘Alī al-Anṭākī (magic squares), 230 Altitude circles, 192
‛Alī b. ‛Isā Use of Astrolabe, 205 al-Uqlīdisī, 43, 44, 46
‘Alī b. Muḥammad al-Qalaṣādī: Removing the al-Uqlīdisī Book of Chapters, 8
Veil, 40 al-Walīd, 75
Alidade, 199, 206 al-Wāthiq, 8
al-Jazarī, 5 Āmū Daryā, 8
al-Jazīra, 5 ‛Amr ibn al- ‛Aṣ, 1
al-Karādīsī, 96 Analogue computer, 208
al-Karajī, 25, 66, 130, 133–135, 137, 138, 143 Analogy, 44, 46, 160, 227, 231
al-Karkhī, 25 Analysis, 16, 63, 88, 91, 134, 147, 148, 185,
al-Kāshī calculation of Sin(l°), 23, 177, 182, 211, 238
184 Angle, bisection, 81, 101, 102, 104, 108, 109
al-Kāshī calculation of π, 21, 22 Angles, spherical, 209, 210
al-Kāshī Calculators’ Key, 9, 22, 23, 62, 71, Antipodal points, 217
106 Apollonios, 6, 83–86, 92, 94, 99
al-Kāshī, Jamshīd, 7, 8, 17, 45, 55, 57, 183 Apollonios Conics, 83, 87, 91, 99, 111, 148
al-Kāshī Planetary Equatorium, 20 Apollonios On Vergings, 92, 94, 111
al-Khalīlī, 3, 177, 215, 220 Approximation, 21, 60, 71, 76, 111, 182, 184,
al-Khāzin, 147 191, 206, 220
al-Khāzinī, 12, 192 Arab armies, 1
al-Khāzinī Sphere that Moves, 192 Arabic, 1, 2, 4, 6, 8–10, 12, 16, 19, 20, 23, 25,
al-Khwārizmī Algèbra, 3, 8, 9, 124, 130, 152, 26, 30–34, 43, 44, 48, 49, 52, 53, 82, 92,
243 99, 104, 112, 118, 122–124, 135, 143, 145,
al-Khwārizmī Astronomical Tables, 164 146, 155, 160–162, 164, 192, 199, 200,
al-Khwārizmī Hindu calculation, 8, 32 205, 212, 216, 227, 229, 230, 236, 239, 240
al-Khwārizmī Image of the Earth, 10 Arabic alphabet, 48, 236, 240
al-Khwārizmī, Muḥammad ibn Mūsā, 8, 30, 72 Arabic language, 2, 4, 5, 25, 26
al-Kindī, 27, 208 Arabic names, 5, 20, 25, 26, 143
al-Ma’mūn, 3, 6, 7, 9, 23, 82, 169, 173, 205, Arabic words, 1, 2, 26, 33, 34, 40, 124, 141,
208 145, 192, 236, 239, 240
Almagest, 18, 47, 59, 79, 112, 155–157, 166, Arabic writing, 23, 25, 44, 52, 53, 99, 112
182, 192 Arabs, 2, 48, 57
al-Māhānī, 147 Aral Sea, 8
al-Manṣūr, 2, 33, 123, 160 Arc of visibility, 194, 215
al-Miṣrī, Najm al-Din, 215 Archimedes, 79, 82, 86, 87, 92, 111, 117, 121,
Almucantars, 199, 200 147
Al-Mu’taman Ibn Hūd Book of Completion, Archimedes Heptagon in a Circle, 111
100, 111 Archimedes Lemmas, 6
al-Muṭī‛, 75 Archimedes Measurement of Circle, 6, 116
al-Nasāwī, 51 Archimedes Quadrature of Parabola
al-Nayrīzī, 177 Archimedes Sphere and Cylinder, 82, 147
Alphabet, Arabic, 26, 48, 135, 236, 240 Arcsine, 164
Alphabet, Greek, 48 Aristotle, 16, 106
al-Qabīṣī, Abū Ṣaqr, 187 Arithmetic, 6–8, 14, 22, 23, 25, 31–34, 37, 38,
al-Samaw’al, 9, 46, 134–138, 146, 153 46, 56, 57, 62, 69, 69, 71, 76, 130, 133,
al-Samaw’al The Shining, 134, 135 134, 138, 145, 152, 164, 183, 185, 226
al-Sijzī, 'Abd al-Jalīl, 7, 92, 94, 107 Arithmetic decimal, 7, 9, 33, 51, 134
al-Ṭabarī, 8 Arithmetic series, 226, 227
Altitude, 10, 114, 115, 161, 199, 205, 206, 212, Arithmetic, astronomers’, 48, 57
214, 215, 217, 219 Arithmetic, Babylonian, 9, 25, 47, 121
Index 249

Arithmetic, Hindu, 9, 32, 43, 51 Cardano, G. Ars Magna, 152


Arithmetic, Indian, 138 Cartography, projections, 9
Arithmetic, sexagesimal, 34, 35, 43, 46, 47, Caspian Sea, 33, 88, 94
50–55 Centers of gravity, 88
Armillary sphere, 192, 193 Ceva’s Theorem, 117
Āryabhaṭa, 226 Charts in mathematics, 6
Aṣbagh b. al-Samḥ, 143 China, 1, 4, 160
Ascendant, 205 Chords in a circle, 156, 158, 159
Ascension, oblique, 196, 211 Christians, 2, 5, 32, 216
Ascension, right, 196 Cipher, 8, 29, 32, 47, 48, 55
Astrolabe, 17, 20, 79, 165, 175, 192, 198–200, Cipherization, 48, 49
202, 204–206, 208, 212 Ciphers, Hindu, 33, 50
Astrology, 13, 43, 180 Circle segment
Athens, 198 its area, 117
Autolykos, 189 its arrow, 116
Auxiliary functions, 176, 177, 185 length of its arc, 116
Auxiliary tables, 32, 177, 179, 211, 212, 215 Circle
Avicenna. See Ibn Sīnā area of, 115
Axis of cone, 84, 84, 96, 98, 203, 204 circumference, 111, 11, 168, 175, 203
Axis of parabola, 82, 84, 85, 87, 96–98, 118, diameter of, 107, 116, 164, 203, 204, 221
148, 149 Classification of cubic equations, 147
Azimuth, 194, 200, 205, 212, 216 Classification of problems, 9
Coefficients, negative, 139
B Cohn-Vossen, S., 201
Baalbek, 6, 123 Combinatorics, enumeration, 234
Babylonians, 29, 35, 47, 121, 196 Common fractions
Baghdad, 2, 3, 6, 7, 11, 27, 74, 75, 82, 88, 106, different, 38
133, 134, 160, 177, 205, 208, 227 partitioned, 38
Bagheri, M., 244 related, 39–42, 43, 76
Balkh, 14 separated by a minus sign, 40
Bankipore, 117 simple, 38
Banū Mūsā, 4, 5, 27, 106, 202 Compass
Barmakids, 4 collapsible, 80
Baṣra, 10 complete, 88, 91
Berggren, J.L., 12 geometric, 105
Bhaskara I, 31, 122 rusty, 104, 106, 108, 109
Bible, 2 Complete quadrilateral, 191
Binomial coefficients, 57, 67, 140, 143, 153, Complex variables, 197
238 Conic sections, 14, 82–83, 86–89, 91–93, 95,
Black Sheep Turks, 18 99, 147, 150, 152
Brahmagupta, 123 Conic sections, construction, 81, 87–90,
Brahmasphuta-siddhanta, 123 92–95, 99
British Museum, 23 Connected (fractions), 42
Būyids, 75, 87, 208 Constantine (Algeria), 145
Byzantine armies, 1 Constructions
Byzantine books, 5, 46 euclidean, 83, 183
Byzantine knowledge of decimal fractions, 46 ruler and compass, 80
Byzantines, 23, 173 rusty compass, 106, 108
verging, 92, 111, 118
C Convex mirror, 99
Cairo, 24, 96, 177, 180, 211–213, 215, 244 Coordinate systems, spherical, 198
Calendar, Gregorian, 14 Cordova, 113, 143
Calendar, Muslim, 1, 11 Cosecant, 162
Calendars, 1, 11, 14, 27, 195 Cosine, 160, 163, 164, 166, 167, 172
250 Index

Cotangent, 161, 152 Elements IV, 156


Cubit, 175 Elements VI, 121, 141, 156, 228
Cycles in root extraction, 21, 46, 62 Elements VII–IX, 141, 223
Elements X, 133
D Elevates, 49, 54, 56
Dahistān, 11 Ephemerides, 179
Daily motion of sun, 155, 195 Equations, cubic, 14, 146, 147
Damascus, 2, 3, 43, 75, 106, 177, 205, 215 Equations, impossible, 126
Dār al-Ḥikma, 7 Equations, number of roots, 125, 126, 128,
Dār al-Kutub (Cairo), 24 129, 144, 145, 149, 150
Day-circle, 194 Equations, quadratic, 9, 92, 121–124, 126, 128,
Declination, 10, 194, 212, 215, 221 132, 133, 144, 152, 153
Deficient number, 134 Equations, rational coefficients, 133
Degree of longitude, 212, 215 Equator, 189, 194–196, 198, 199, 205–207,
Degrees, 19, 32, 35, 44, 49, 54, 146, 151, 165, 216, 220
175, 176, 179–181, 193, 196, 205, 213, Equatorium, 18, 20
217, 234, 244 Equinoxes, 205
Delambre, J.-B., 212 Eratosthenes, 9, 173, 223
Demonstration, 16, 38, 71, 80, 126, 131, 140, Euclid, 6, 16, 43, 79, 80, 82, 83, 99, 108–110,
151, 224 114, 121, 126–129, 133, 159, 168, 182,
Demonstrations, Greek, 80 190, 199, 208, 223
Demonstrations, in algebra, 71, 131, 151 definition of proportion, 16
Descartes, R., 91 difficulties, 14, 16
Descendant (of zodiacal sign), 205 on conics, 6, 83, 87, 148
Differential calculus, 185 parallel postulate, 16
Dimensions of cosmos, 17 Euclid Data, 6
Dinars, 213 Euclid Elements, 6, 15, 19, 80, 87, 126, 127,
Diocles, 99 135, 141, 156, 199, 228
Diophantos, 6, 121, 123, 134, 223, 224 Euclid Optics, 6, 99
Diophantos Arithmetica, 121, 123, 223 Euphrates, 5, 32
Dirhām, 74 Eutocios, 6
Discriminant, 126 Excess, 73, 85, 134, 135
Divisibility, tests for, 41 Exponents, 135, 136, 143, 144
Division
of line segment, 89 F
of numbers, 41 Factorization of whole numbers, 40, 46, 47, 56,
of polynomials, 134, 137 164
Diyar bakr, 5 False position, 132, 133
Djebbar, A., 41, 113, 143, 144, 227, 244 Fāṭima, 180
Double cone, 83, 84, 95 Fatimids, 180
Doubling, 41, 43, 58 Fermat, P., 225
Dust board, 34, 43, 50, 57, 59, 63, 137 Fibonacci, 130
Fifth root, 22, 57, 61, 62, 65, 68, 71
E Fils, 74
Easter date, 32 Flow chart, 62
Eclipse of moon, 11, 17, 194 Folk astronomy, 214
Ecliptic, 189, 194–196, 199, 205, 207, 211, Folkerts, M., 31
212, 215, 220, 221 Fourth proportional, 149
Egypt, 1, 5, 104, 146, 180, 202, 213 Fractions, common. See ‘common fractions’
Elements I, 15 Fractions, decimal, 7, 9, 21, 22, 31, 43, 44, 46,
Elements II, III, 98, 102, 114, 116, 121, 122, 56, 68
141, 158, 168, 172, 228 Fractions, estimating, 76
Index 251

Fractions, leveling, 51 Heron Metrica, 111


Fractions, sexagesimal, 74, 156, 168 Heuristics, 136
Fractions, “Turkish”, 46 Hilāl al-Ḥimsī, 6
Fractions, unit, 38, 73 Hilbert, D., 201
France, 1 Hind, 173
Fulus, 35, 36, 38, 44, 74 Hindu astronomy, 32, 33
Hindu mathematics, 31, 123
G Hindus, 8, 31, 32, 123
Gabriel, angel, 1 Hipparchos, 155, 160, 197, 234
Gabriel, translator, 4 Hippocrates of Chios, 92
Galen, 5 Hitti, Phillip, 1
Galois, E., 87 Hogendijk, J.P., 185
Gauss, C.F., 32 Horizon, 95, 161, 175, 189, 191, 195,
Generality, 125, 171, 233 198–200, 206, 207, 212, 217
Generator of cone, 84 Horner’s method, 63
Geography, mathematical, 173, 189, 216 Horoscopes, 8, 205
Geometry, 2, 15, 16, 31, 57, 79, 80, 82, 88, 91, Hour angle, 194, 212
92, 109, 112, 113, 122, 129, 133, 148, 156, Hour lines, 207
189, 192, 208 Hours, equinoctial, 205
descriptive, 155, 220 Hours, seasonal, 198, 205
Greek sources, 82 Houses, astrological, 180, 205
moving, 92 Ḥunayn b. Isḥāq, 6
solid, 220 Hunger, H., 46
Ghazna, 12, 173 Hypatia, 198
Ghuzz, 11 Hyperbola, 84, 86, 87, 91, 95, 98, 150
Gibbon, E., 2 Hyperbola, construction, 94, 95
Gibraltar, 2 Hyperbola, symptom, 83, 86, 98, 117
Gifts (in legacies), 152 Hypotenuse of reversed shadow, 162
Gnomon, 20, 127, 128, 131, 161 Hypotenuse of shadow, 162
Graduated ring, 11, 192
Great circle, 99, 173, 189–192, 194, 195, I
197–199, 207, 208, 210, 217 Ibn Abī Ḥajala, 180
Greek language, 2, 5, 12, 32 Ibn al-Bannā’, Algebra, 144
Greeks, 31, 32, 47, 48, 80, 92, 95, 110, 122, Ibn al-Bannā’, Raising the Veil, 39, 235
127, 189, 192, 194, 196, 201, 223 Ibn al-Bannā’, Summary Account, 39, 40
Guinea-Bissau, 11 Ibn al-Haytham, 15, 16, 83, 99, 100, 111
Gunbad, 13 Ibn al-Haytham/Optics, 100
Gutas, D., 4 Ibn al-Majdī Enveloping the Core, 244
Ibn al-Nadīm, 6, 83
H Ibn al-Shāṭir, 3
Ḥabash al-Ḥāsib, 176, 208 Ibn Khaldūn, 39, 76
Half-angle formulae, 176 Ibn Khallikān, 180
Halving, 34, 35, 41, 43, 44, 116, 152 Ibn Mun'im, Laws of Calculation, 40, 227, 236
Harran. See Diyar bakr Ibn Sīnā, 5
Harriot, Thomas, 201 Ibn Yūnus, 177, 180–180, 211–213, 215, 217
Harūn al-Rashīd, 3, 6, 82 Ibn Yūnus Ḥākimī Zīj, 211
Haywood, J.A., 26, 48 Ibn Yūnus Sine Tables, 180–182
Hebrews, 48 Ibn Yūnus Very Useful Tables, 212, 213, 215
Height of mountain, 173, 174, 187 Ibraham b. Sinān, 24, 82, 89, 95, 97, 98, 106,
Height of sun, 220 202
Heirs, natural, 71, 152 Ibrāhīm ibn Sīnān Analysis and Synthesis, 96
Heptagon, 6, 82, 87, 89, 91, 106, 111 Ibrāhīm ibn Sīnān Drawing Conic
Heptagon, construction, 87, 89, 90, 91 Sections, 88
Heron, 111, 121 Ibrāhīm ibn Sīnān On Sundials, 96
252 Index

Identities Koran. See Qur’ān


algebraic, 61, 66, 130, 132, 142, 152, 231 Kunya, 5
trigonometric, 185 Kūshyār ibn Labbān, 33, 138, 176
Ihsan, 180 Hindu Reckoning, 40, 50
‛ilm al-farā’iḍ, 71, 76
‛ilm al-miqāt, 212 L
‛ilm al-waṣāyā, 152 Laqab, 5
‛Imād al-Daula, 75 Latin, 8, 16, 33, 49, 69, 99, 124, 163, 164, 192,
India, 1, 12, 23, 31, 155, 173, 175, 179, 199, 200, 208
226, 234 Latitude, 10, 11, 96, 173, 177, 189, 196, 199,
Indian sources, 43 200, 211, 215, 217, 220, 221
Indians, 2, 12, 26, 31, 43, 62, 123, 138, 155, Latitude measurement, 10, 96, 173, 177, 189,
160, 177, 229 196, 199, 200, 211, 214–217, 219–221
Inheritance, 9, 10, 37, 71, 73, 152 Latus rectum, 85
Interpolation, 59, 60, 71, 76, 156, 165, Latus transversum, 85
175–182, 185, 211 Law of exponents, 135, 136
linear, 178, 179 Leap years, 14
second-order, 180–182 Legacies, 73, 152
Iran, 14, 17, 23, 25, 135 Length of daylight, 196, 205, 212, 214, 221
Iraq, 1, 25, 208, 213, 234 Line segment, division of, 81, 89, 118, 121,
Isfahan, 14–16, 105, 106, 147 122
Isḥāq ibn Ḥunayn, 4, 6 Long division, 41, 139
Islām, 1–3, 9, 11, 12, 22, 25, 27, 55, 75, 113, Longitude measurement, 10, 19, 20, 173, 195,
135, 145, 152, 160, 208, 215, 227, 229, 244 205, 212, 213, 216, 220, 216
Islamic art, 105 Lorch, R., 192, 202
Iteration, 185 LSin, 181
fixed point, 185, 194 Lunar eclipse, 17

J M
Jābir b. Ḥayyān, 234 Macron, 26
Jerusalem, 216 Madrasa, 18, 23, 105
Jews, 216 Maghrib, 2, 4, 9, 38, 41, 76, 134, 143, 145,
John of Seville, 69 198, 227, 235
Jurist, 10 Magic squares, 228–230, 229
Jurjān, 11 Maḥmūd of Ghazna, 12, 173
Māl, 125, 126, 128–130, 134–136, 142, 14,
K 145
Ka‛ba, 1, 216 Malikshah, 14
Kāshān, 17 Manuscript libraries
Kasir, D.S., 148 Dār al-kutub, 24, 96
Kāth, 11 Fāḍilīya, 23
Katz, V., 28, 29 European, 5, 16, 23, 25, 46, 151, 155
Keneshra, 32 Manuscripts, Arabic, 23, 25, 92, 135
Kennedy, E.S., 12, 18, 20, 28 Maragha, 135
Kepler, J., 87 Mark
Khālid al-Marwarrūdhī, 10 for decimal place, 9, 21, 69, 164, 187
Khāqānī tables, 18 for sexagesimal place, 56, 164, 182–184
Khurasān, 14 Marrakesh, 39, 40, 236
Khwārizm, 3, 7, 8–11, 32, 33, 69, 71, 73, 74, Martel, Charles, 1
124–126, 130, 131, 147, 152, 164, 243 Marw, 14
King, D.A., 52, 97, 180, 211, 212, 214, 217 Maḥmūd of Ghazna, 12, 173
Index 253

Maximum of product, 182 Transversal Figure, 156


Mean-value theorem, 185 Nestorians, 32
Measurement, 6, 57, 113, 114, 117, 173, 175, Neugebauer, O., 164
204 Nicomachus Introduction to Arithmetic, 223,
Mecca, 1, 3, 160, 177, 213, 216, 217, 219–221 226, 227
Medīna, 1, 75 Nilometer, 202
Medīna al-Salām. See Baghdad Nisba, 5, 143
Menaechmos, 87 Nishapur, 14
Menelaos, 6, 190, 220, 221, 235 Niẓām al-Mulk, 14
Menelaos’ Theorem, 191, 220, 224, 235 North Africa, 1
Menelaus Spherics, 112 Numbers
Meridian, 3, 9, 10, 11, 20, 21, 173, 175, 189, absolute, 134, 151
192, 193, 194, 197, 201, 206, 214, 216, 217 amicable, 224
Meridian degree, 11 figured, 223, 226–228
Merinid Dynasty, 39 irrational, 122, 133
Meshhed, 23 negative, 47, 123, 134, 179
Mesopotamia, 5, 47, 121 perfect, 223, 224, 227
Minutes, 19, 32, 35, 36, 44, 50, 51, 54, 55, 74, rational, 121, 133, 223–225
175, 179, 181, 186, 196, 211, 213 Schroeder, 234
Miqyās, 96, 161 signed, 123
Mirrors, 41, 79, 87, 97, 99, 118 simple, 124
Models, 18, 79, 155, 156, 160, 192, 196, 197, whole, 31, 34, 40, 43, 44, 46, 47, 56, 59,
208 130, 138, 164, 223, 224, 225, 226, 239,
Moon, visibility of crescent, 160 244
Morocco, 25 Numeration
Mosque alphabetic, 48
Isfahan, 15, 16, 105, 106, 147, 244 decimal, 33, 43, 51
Umayyad, 2–4, 75, 113, 215 decimal positional, 8, 31, 43
Mosul, 10 East Arabic, 2, 227
Mu‛izz al-Dawla, 75 Hindu, 9, 32, 33, 199
Mughals, 23 Hindu-Arabic, 31, 34
Muḥammad, 1, 2, 5, 7, 8, 23, 32, 38, 40, 74, 75, sexagesimal, 34, 35, 43, 46, 50, 123
113, 124, 151, 177, 180, 204, 215, 220 Numerical methods, 178, 207
Muḥammad al-Ḥaṣṣār Demonstration and
Reminder, 38–40, 143 O
Muḥammad ibn ‘Abdūn, 113–117 Obliquity of ecliptic, 194, 215
Multiplication, 9, 34, 36, 37, 40, 46, 51, 52, 53, Observations of sun, 156, 161, 192, 211
54, 61, 68, 77, 135, 137, 138, 144 Observatory
Multiplication tables, 52 Baghdad, 3, 33, 82, 88, 177
Multiplication, gelosia method, 55 Isfahan, 14, 16
Muqaṭṭam Hills, Cairo, 24, 96, 177, 180, 181, Samarqand, 17, 21, 57
211–215, 244 Omar Khayyam. See ‛Umar al-Khayyāmī
Muslim architecture, 22 Orders in algebra, 9, 121, 124, 128, 130
Muṣṭafa Sidqī, 24 Ordinates, 84, 98
Mutawakkil, 202 Oxus River. See Āmū Daryā
Muwaqqit, 3, 96, 220. See also Timekeeper
P
N Pakistan, 2
Nādir Shāh, 23 Palestine, 5
Nahmad, H.M., 26, 48 Paper, 23, 27, 34, 43, 57, 80, 137, 201
Nandana, 174, 175 Pappus of Alexandria, 92, 104
Napier, John, 46 Pappus’ Mathematical Collection, 100, 104
Naṣīr al-Dīn al-Ṭūsī Parabola
Ilkhani Zij, 211 chord, 84
254 Index

Parabola (cont.) conformal, 201


construction, 97, 117 globular, 12
diameter, 149 Pronunciation of Arabic letters, 49, 145
symptom, 83 Prosthaphairesis, 212
Parallel circles, 189, 191 Pseudo-Ṭūsī, 16
Parallel postulate, 16 Ptolemais, 198
Parallelopipeds, 148 Ptolemy
Parallels of latitude, 189 Almagest, 18, 47, 59, 79, 155–157, 166,
Parameter of conics, 85 182, 192
Part, 9, 11, 14, 21, 23, 31, 34, 35, 38, 40–43, Geography, 8, 173
47, 53–57, 68, 69, 71, 79, 82, 83, 85, 88, Optics, 99, 100, 215
94, 97, 110, 124, 130, 134, 135, 144, 147, Planispherium, 197, 202
151, 152, 155, 164, 165, 178, 185, 194, Pythagoreans, 226
199, 204, 205, 208–210, 214 Pythagorean theorem, 114, 156, 167
Pascal Traitè du Triangle, 66
Pascal’s triangle, 66, 67, 140, 238 Q
Patronage, 180 qāḍī. See Jurist
Banū Mūsā, 106 Qāḍīzādeh Rūmī, 22
Būyid, 88 Qairawān, 9
Fakhr al-Mulk, 133 Qibla, 216, 217, 220
Shāh Jahān, 23 Qibla tables, 220
Ulūgh Beg, 18, 211 Quadrant, 165, 210, 217
Pell’s equation, 123 Quantities, algebraic, 130
Pen, 34, 43, 57 Qur’ān, 5, 19, 26
Pentagon, construction, 79 Qusṭā b. Lūqā
Pepper, J.V., 201 On the Sphere that Moves by Itself, 191
Perpendicular, construction, 81, 86, 97, 107, Sphere with a Frame, 192
114, 149, 160
Persia, 1 R
Persian, 2, 3, 12, 17, 23, 26, 88, 160 Radius, 9, 80, 99, 110, 116, 156, 160, 161,
Phoenecians, 48 163–165, 168, 176, 189
Plane problems, 119, 147 Radius of earth, 175
Plato, 4, 89 Raising fractions, 38
Pointer of sundial, 95 Rashed, R., 46, 134
Points on Arabic letters, 49, 145 Ratio, 9, 16, 73, 79, 82, 89, 100–104, 111–113,
Poles on sphere, 192 121, 147, 160, 163, 171, 175, 224, 225
Postage stamps, 4 Real number, 16
Prayer Rectangles, areas of, 141
direction, 177, 214, 216 Reduction by a number, 41
times, 2, 96, 176, 212, 214, 215, 244 Refraction, 99, 215
Precision Regular nonagons, 91–95
in division, 56 Regular polygons, 82, 92
in trigonometric tables, 175 Regular polyhedra, 79
Problems Renaissance, 96, 151
curvilinear, 92 Rete, 198, 200
indeterminate, 226 Rhodes, 155, 197
solid, 57, 92 Right cone, 83
Projection Right triangles, spherical, 210
equidistant, 12 Rising times, 195–196
azimuthal, 12 Rome, 16, 190
circle preserving, 200 Roots, algebraic, 14
Index 255

Roots, cube, 32, 62, 146, 152 Sind. See Pakistan


Roots, fifth, 22, 61, 62, 65, 69 Sine, 154, 159–161, 163–167, 170, 175, 178,
Roots, square, 34, 57–61, 65, 125, 130, 134, 179, 181, 182, 217, 219
145, 146, 225 Sine Law, 170
Rosen, F., 125 Sine Law, for spherical triangles, 209, 217
Rosenfeld, B., 185 Sine, al-Bīrūnī’s concept of, 175, 177
Rosenthal, F., 5, 76 Sines, addition theorem, 166–169
Rule of four quantities, 208–210, 219 Sinjār, 10
Rule of signs, 131, 134 Size of cosmos, 17
Ruler in geometry, 92 Size of Earth, 9
Ruler of astrolabe, 205 Solar longitude, 212, 213
Russia, 25 Solutions, universal, 215
Solving right triangles, s.a., 172
S Solving right triangles, s.s., 172
Ṣā‘id al-Andalūsī, 4 Souissi, M., 39, 144
Sabians, 5 Spain, 1, 2, 25, 104, 113, 160, 199
Saccheri, G., 16 Sphere, celestial, 189, 192–195, 198
Saidan, A.S., 74 Spherical astronomy, 176, 177, 179, 211–215
Saliba, G., 2 Spherics, 189, 192, 195, 212, 220
Samarqand, 7, 14, 17–21, 57, 147, Spica, 205
183, 185 Spiral, 92
Samt, 192, 194 Square, construction, 230
Sanad b. ‛Alī, 173 Squares, sums of, 224–226
Sanskrit, 12, 33, 123, 160, 234 Stade, 9
Sanskrit scientific works, 3 Star, 5, 192–194, 198, 199, 205, 206, 215
Saphea, 199 Star catalog, 21
Saragossa, 100 Star map, 20, 198, 200, 205, 206
Secant, 101, 162 Star, its arc of visibility, 194
Seconds, 49–51, 54, 74, 211 Star, time since rising, 194, 215
Section, “golden”, 110 Stereographic projection, 197–205
Sedillot, J.-J., 165 Stevin, S. The Tenth, 46
Sedillot, L.-A., 165 Straightedge, 80, 87, 88, 91, 94, 106, 118
Seljuks, 14 Subtraction, 8, 32–35, 40, 50, 123, 126, 134,
Semitic languages, 25, 48 138, 139, 233
Sesiano, J., 229, 230, 233 Sulṭān Iskandar, 18
Severus Sebokht, 32, 199 Sun, 10, 19, 20, 95, 155, 156, 161, 179, 192,
Seville, 143 194–196, 199, 205–207, 212–215, 220, 221
Sextant, 21 Sundials, 79, 87, 95–97
Shadow lengths, 161 Surya Siddhanta, 160
Shāh Jahān, 23 Synesios, 198
shahāda, 75 Synthesis, 9, 89, 96
Shares, 71, 73, 152 Syria, 1, 4, 102
Shifting in algebra, 138 Syriac, 2, 12, 199
Shifting in arithmetic, 36
Shirāz, 92, 94 T
Side (in algebra), 124 Tabarsī, 23
Signs of zodiac, 20, 195 Table of chords, 155–159, 186, 197, 220
Simultaneous discovery, 209 Tables
Sin (60°), 183 multiplication, 52
Sin (12°), 183 spherical trigonometry, 215
Sin (3°), 183–185 timekeeping, 215
Sin (6°), 183 trigonometric, 175, 181, 185
Sin (72°), 183 Tajikistān, 25
Sin (l°), 23, 176, 177, 183–184 Tamurlane, 17, 18
256 Index

Tangent, 81, 84, 98, 99, 101–102, 104, 119, ‛Umar al-Khayyāmī Difficulties in Euclid, 16
161, 162, 166, 175–176, 179, 185, 203, 204 ‛Umar al-Khayyāmī Rub‛āyāt, 13
Thābit ibn Qurra, 5, 15, 122 Umayyads, 2
Theon of Alexandria, 198, 199 Unequal hours, 205
Thing, in algebra, 134 United States, 25
Thirds, 49, 51, 55, 56, 74, 211 Unknowns, 122–124, 144, 151
Thousand and One Nights, 3 Urgench, 8
Tilings, 104 Uzbekistān, 8, 25
Time of day, 198, 205, 207
Time since sunrise, 212 V
Timekeeper. See muwaqqit van der Waerden, B.L., 31
Timekeeping, 164, 176, 211–215 Ventilators, 213, 214
Toomer, G.J., 99, 157 Venus, 180, 192
Toth, I., 15 Verging constructions, 88, 92, 94
Tours, France, 1 Versed sine, 163, 164
Tower of Winds, 198 Vertex
Transfer of lengths, 80 of cone, 83
Translations of parabola, 97, 119, 148
from Greek, 5, 6 of hyperbola, 91
Latin, 8, 16, 33 Vitruvius, 198
Transliteration, 26, 53 Vogel, K., 46
Transverse side, 85, 91, 94, 98
Triangles W
area, 114, 117 Wallis, J., 16
spherical, 191 Wealth (algebra), 74, 111, 124
Trigonometry, 107, 155, 156, 160, 164, 166, Wensink, H., 28
173, 176, 177, 185, 189, 191, 208, 209, Winds (in Cairo), 198, 213, 214
211, 215, 216, 220 Words, enumeration, 241
spherical, 7, 189, 191, 208, 209, 216 World map, 10
Trisection of angle, 92
Tritton, A.S., 26 Y
Tropic of Cancer, 199, 200, 206 Yaḥyā ibn Aktham, 10
Tropic of Capricorn, 195, 199–200, 206
Tunis, 40, 146, 214 Z
Turks, 11, 14 zakāt, 74, 77
Ṭūs, 14 Zenith, 192, 193, 196, 200, 217
Twilight, 213, 214 Zero, 31–34, 41, 47, 48, 68, 124, 126, 134, 137
zīj, 160
U Zīj al-Majisṭī. See Abū a1-Wafā’
U.S.S.R., 25 Zīj al-Sindhind, 2
Ulūgh Beg, 17–19, 22, 23, 176, 211 Zīj, al-Ḥākimī. See Ibn Yūnus
‛Umar al-Khayyāmī, 7, 13, 15, 124, 146 Zīj, Ilkhanī. See Naṣīr al-Dīn al-Ṭūsī
‛Umar al-Khayyāmī Algebra, 146 Zodiac, 49, 195, 200, 206

You might also like