THE THIRTEEN BOOKS OF EUCLID'S ELEMENTS * : "?^iV?-*' v ivv', J. ■* V L < -*V ^VJ * : *, JT o.UO!?-t' ytX/T*-«^t «wrt>ft«o U«rt ^ !■ 1 ■ •*•■ JLiyjur- ^ v-tJj ■ 6-a-^apttJ 67 ku lyL-»i &JJ - r • "» ft' ' , ' I ' "" , -j. *v * f-yTj^>»4j« Imly o l T*y^wff1.V «j(yjy-ftwo-, o u tu5^- Uj S-l (X^ftT» ta f^H itr^^V ' ^6t«Ctt) t-ajft-v-fay bdrftfip d^-roy £lr'li_Mty-e". iVtw^tF ^E-^ t^ 4 t-2;' ■an* / t 1 j» &rf4« fr bK^i^oji lufrvrpo'y (^t t-»« rAM tujt.jun/ . ior*n^lp Irvtr"rlni I I I i I 1 1 1 1 _t THE THIRTEEN BOOKS OF EUCLID'S ELEMENTS TRANSLATED FROM THE TEXT OF HEIBERG WITH INTRODUCTION AND COMMENTARY BY Sir THOMAS L. HEATH, K.C.B., K.C.V.O., F.R.S., SC.D. CAMS,, HON. D.SC. OXFORD HONORARY FELLOW (SOMETIME FELLOW) OF TRINITY COLLEGE CAMBRIDGE SECOND EDITION REVISED WITH ADDITIONS VOLUME I INTRODUCTION AND BOOKS I, H DOVER PUBLICATIONS, INC. NEW YORK Published in Canada by General Publishing Com- pany, Ltd., 30 Lesmill Road, Don Mills, Toronto, Ontario. Published in the United Kingdom by Constable and Company, Ltd., 10 Orange Street, London WC 2. This Dover edition, first published in 1956, is an unabridged and unaltered republication of the second edition. It is published through special ar- rangement with Cambridge University Press. Library of Congress Catalog Card Number: 56-4H6 Manufactured in the United States of America Dover Publications, Inc. 180 Varick Street New York, N.Y. 10014 PREFACE " ' I 'HERE never has been, and till we see it we never J. shall believe that there can be, a system of geometry worthy of the name, which has any material departures (we do not speak of corrections or extensions or developments) from the plan laid down by Euclid." De Morgan wrote thus in October 1 848 {Short supplementary remarks on the first six Books of Euclid's Elements in the Companion to the Almanac for 1 849) ; and I do not think that, if he had been living to-day, he would have seen reason to revise the opinion so deliberately pronounced sixty years ago. It is true that in the interval much valuable work has been done on the continent in the investigation of the first principles, including the formulation and classification of axioms or postulates which are necessary to make good the deficiencies of Euclid's own explicit postulates and axioms and to justify the further assumptions which he tacitly makes in certain propositions, content apparently to let their truth be inferred from observa- tion of the figures as drawn ; but, once the first principles are disposed of, the body of doctrine contained in the recent text- books of elementary geometry does not, and from the nature of the case cannot, show any substantial differences from that set forth in the Elements. In England it would seem that far less of scientific value has been done ; the efforts of a multitude of writers have rather been directed towards producing alter- natives for Euclid which shall be more suitable, that is to say, easier, for schoolboys. It is of course not surprising that, in vi PREFACE these days of short cuts, there should have arisen a movement to get rid of Euclid and to substitute a "royal road to geometry " ; the marvel is that a book which was not written for schoolboys but for grown men (as all internal evidence shows, and in particular the essentially theoretical character of the work and its aloofness from anything of the nature of "practical" geometry) should have held its own as a school- book for so long. And now that Euclid's proofs and arrange- ment are no longer required from candidates at examinations there has been a rush of competitors anxious to be first in the field with a new text- book on the more "practical" lines which now find so much favour. The natural desire of each teacher who writes such a text -book is to give prominence to some special nostrum which he has found successful with pupils. One result is, too often, a loss of a due sense of proportion ; and, in any case, it is inevitable that there should be great diversity of treatment. It was with reference to such a danger that Lardner wrote in 1846 : "Euclid once superseded, every teacher would esteem his own work the best, and every school would have its own class book. All that rigour and exactitude which have so long excited the admiration of men of science would be at an end. These very words would lose all definite meaning. Every school would have a different standard; matter of assumption in one being matter of demonstration in another ; until, at length, Geometry, in the ancient sense of the word, would be altogether frittered away or be only considered as a particular application of Arithmetic and Algebra." It is, perhaps, too early yet to prophesy what will be the ultimate outcome of the new order of things ; but it would at least seem possible that history will repeat itself and that, when chaos has come again in geometrical teaching, there will be a return to Euclid more or less complete for the purpose of standardising it once more. But the case for a new edition of Euclid is independent of any controversies as to how geometry shall be taught to schoolboys. Euclid's work will live long after all the text-books PREFACE vii of the present day are superseded and forgotten. It is one of the noblest monuments of antiquity ; no mathematician worthy of the name can afford not to know Euclid, the real Euclid as distinct from any revised or rewritten versions which will serve for schoolboys or engineers. And, to know Euclid, it is necessary to know his language, and, so far as it can be traced, the history of the *' elements " which he collected in his immortal work. This brings me to the raison d'itre of the present edition. A new translation from the Greek was necessary for two reasons. First, though some time has elapsed since the appearance of Heiberg's definitive text and prolegomena, published between 1883 and 1888, there has not been, so far as I know, any attempt to make a faithful translation from it into English even of the Books which are commonly read. And, secondly, the other Books, vn. to X. and xin., were not included by Simson and the editors who followed him, or apparently in any English translation since Williamson's (178 1 — 8), so that they are now practically inaccessible to English readers in any form. In the matter of notes, the edition of the first six Books in Greek and Latin with notes by Camerer and Ha'uber (Berlin, 1824 — 5) is a perfect mine of information. It would have been practically impossible to make the notes more exhaustive at the time when they were written. But the researches of the last thirty or forty years into the history of mathematics (I need only mention such names as those of Bretschneider, Hankel, Moritz Cantor, Hultsch, Paul Tannery, Zeuthen, Loria, and Heiberg) have put the whole subject upon a different plane. I have endeavoured in this edition to take account of all the main results of these researches up to the present date. Thus, so far as the geometrical Books are concerned, my notes are intended to form a sort of dictionary of the history of elementary geometry, arranged according to subjects ; while the notes on the arithmetical Books vii. — ix. and on Book x. follow the same plan. viii PREFACE I desire to express here my thanks to my brother, Dr R. S. Heath, Vice- Principal of Birmingham University, for suggestions on the proof sheets and, in particular, for the reference to the parallelism between Euclid's definition of proportion and Dedekind's theory of irrationals, to Mr R. D. Hicks for advice on a number of difficult points of translation, to Professor A. A. Bevan for help in the transliteration of Arabic names, and to the Curators and Librarian of the Bodleian Library for permission to reproduce, as frontispiece, a page from the famous Bodleian MS, of the Elements. Lastly, my best acknowledgments are due to the Syndics of the Cambridge University Press for their ready acceptance of the work, and for the zealous and efficient cooperation of their staff which has much lightened the labour of seeing the book through the Press. T. L H. Novem$er t 1908. PREFACE TO THE SECOND EDITION I LIKE to think that the exhaustion of the first edition of this work furnishes a new proof (if such were needed) that Euclid is far from being defunct or even dormant, and that, so long as mathematics is studied, mathematicians will find it necessary and worth while to come back again and again, for one purpose or another, to the twenty-two-centuries- old book which, notwithstanding its imperfections, remains the greatest elementary textbook in mathematics that the world is privileged to possess. The present edition has been carefully revised throughout, and a number of passages (sometimes whole pages) have been rewritten, with a view to bringing it up to date. Some not in- considerable additions have also been made, especially in the Excursuses to Volume I, which will, I hope, find interested readers. Since the date of the first edition little has happened in the domain of geometrical teaching which needs to be chronicled. Two distinct movements however call for notice. The first is a movement having for its object the mitigation of the difficulties (affecting in different ways students, teachers and examiners) which are found to arise from the multiplicity of the different textbooks and varying systems now in use for the teaching of elementary geometry. These difficulties have evoked a widespread desire among teachers for the establish- ment of an agreed sequence to be generally adopted in teaching the subject. One proposal to this end has already been made: but the chance of the acceptance of an agreed sequence has in the meantime been prejudiced by a second movement which has arisen in other quarters. x PREFACE TO THE SECOND EDITION I refer to the movement in favour of reviving, in a modified form, the proposal made by Wallis in 1663 to replace Euclid's Parallel- Postulate by a Postulate of Similarity (as to which see pp. 2 10 — 1 1 of Volume I of this work). The form of Postulate now suggested is an assumption that "Given one triangle, there can be constructed, on any arbitrary base, another triangle equiangular with (or similar to) the given triangle." It may perhaps be held that this assumption has the advantage of not referring, in the statement of it, to the fact that a straight line is of unlimited length ; but, on the other hand, as is well known, Saccheri showed ( 1 733) that it involves more than is necessary to enable Euclid's Postulate to be proved. In any case it would seem certain that a scheme based upon the proposed Postulate, if made scientifically sound, must be more difficult than the procedure now generally followed. This being so, and having regard to the facts ( 1 ) that the difference between the suggested Postulate and that of Euclid is in effect so slight and (2) that the historic interest of Euclid's Postulate is so great, I am of opinion that the proposal is very much to be deprecated. T. L. H. December 1925. CONTENTS VOLUME I. INTRODUCTION. Facsimile ofapageofthe Bodleian ms. of the Elements . Frontispiece This is a facsimile of a page (fol. 45 verso) of the famous Bodleian Ms. of the Element i, D'Orville 301 {formerly x. r inf. 1, 30), written in the year 888. The scholium in the margin, not very difficult to decipher, though some letters are almost rubbed out, is one of the scholia Vatican a given by Heiberg (Vol. v. p. 163} as ill. No- 15 : Aid tw xirrpov tOcQv q&k rj v f»rn$«tft aftav, tl Six* Tifiwauti* dWiJXaT * t4 y&p icfwrpor a&r&r 1j Sixrrofdn, bjLoltits xai i} et r$F tripe* Bt& rov Kfrrpov aCenjt 17 trip* ^ 5ta rav KfvTpov tlTj, Sri 06 3fva r4fit*erai 1j && tou Ktvrpoy. The rt before U in the last sentence should be omitted, PFVat. read ij without el. The marginal references lower down are of course to propositions quoted, (1) Sti t& a' rod y. " by III. i," and (aj &i ri 7' toC atow, u by 3 of the same/ 1 Chap. I. II. III. IV. V. VI. VII. VIII. IX. Euclid and the traditions about him Euclid's other works Greek commentators other than Proclus Proclus and his sources .... The Text The Scholia Euclid in Arabia ...... Principal translations and editions . § 1. On the nature of Elements g a. Elements anterior to Euclid's . § 3. First principles : Definitions, Postulates and Axioms ..... Theorems and Problems The formal divisions of a proposition Other technical terms The definitions §4- §5- §7- THE ELEMENTS. Definitions, Postulates, Common Notions Notes on Definitions etc. . Propositions .... Definitions .... Note on geometrical algebra Propositions .... Pythagoras and the Pythagoreans II. Popular names for Euclidean Propositions Greek Index to Vol. I English Index to Vol. I Book I. Book II. Excursus Excursus pack 1 7 19 *9 46 64 75 9i 114 116 117 124 129 132 143 1 S3 i55 241 370 37» 375 411 415 419 421 INTRODUCTION. CHAPTER I. EUCLID AND THE TRADITIONS ABOUT HIM. As in the case of the other great mathematicians of Greece, so in Euclid's case, we have only the most meagre particulars of the life and personality of the man. Most of what we have is contained in the passage of Proclus' summary relating to him, which is as follows 1 : "Not much younger than these (sc. Hermotimus of Colophon and Philippus of Med ma) is Euclid, who put together the Elements, collect- ing many of Eudoxus' theorems, perfecting many of Theaetetus', and also bringing to irrefragable demonstration the things which were only somewhat loosely proved by his predecessors. This man lived* in the time of the first Ptolemy. For Archimedes, who came imme- diately after the first (Ptolemy)', makes mention of Euclid: and, further, they say that Ptolemy once asked him if there was in geometry any shorter way than that of the elements, and he answered that there was no royal road to geometry*. He is then younger than the pupils of Plato but older than Eratosthenes and Archimedes ; for the latter were contemporary with one another, as Eratosthenes some- where says." This passage shows that even Proclus had no direct knowledge of Euclid's birthplace or of the date of his birth or death. He pro- ceeds by inference. Since Archimedes lived just after the first 1 Proclus, ed. Fried lein, p. 68, 6 — io, * The word ytyott must apparently mean " nourished," as Heiberg understands it (LitttrttrgathicktHeht Stttditn itber EtMid, 188,1, p. 16), not "was born," as Hankel took it ; otherwise part of Proclus' argument would lose its cogency. * So Heiberg understands ?n/iii\£iv n? *p&Ti# (sc. IlToXe^ttV). Friedlein's text has vol between Iwifi&Xuv And t<£ *p&ry \ and it is right to remark that another reading is ttX it Tif Trptirif (without trtpaXiiy'i which has been translated " in his first book," by which is understood On the Sphere and Cylinder I., where (i) in Prop- 3 »™ the words " let BC be made equal to D by tkt second (proposition) of the ftrit of Euclid's (books)," and (i) in Prop, fi the words " For these things are handed down in the Elements " (without the name of Euclid). Heiberg thinks the former passage is referred to, and that Proclus must therefore have had before him the words " by the second of the first of Euclid ": a fair proof that they are genuine, though in themselves they would be somewhat suspicious. 4 The same story is told in Stobaeus, Eel. (it. p. m8, 30, ed, Wachsmuthf about Alexander and Menaechmus. Alexander is represented as having asked Menaechmus to teach him geometry concisely, but he replied 1 "O king, through the country there are royal roads and road* for common citizen*, but in geometry there is one road for all." * INTRODUCTION [ch. i Ptolemy, and Archimedes mentions Euclid, while there is an anecdote about some Ptolemy and Euclid, therefore Euclid lived in the time of the first Ptolemy. We may infer then from Proclus that Euclid was intermediate between the first pupils of Plato and Archimedes. Now Plato died in 347/6, Archimedes lived 287-2 1 2, Eratosthenes c. 284-204 B.C. Thus Euclid must have flourished c. 300 b.c, which date agrees well with the fact that Ptolemy reigned from 306 to 283 B.C. It is most probable that Euclid received his mathematical training in Athens from the pupils of Plato; for most of the geometers who could have taught him were of that school, and it was in Athens that the older writers of elements, and the other mathematicians on whose works Euclid's Elements depend, had lived and taught. He may himself have been a Platonist, but this does not follow from the state- ments of Proclus on the subject, Proclus says namely that he was of the school of Plato and in close touch with that philosophy 1 . But this was only an attempt of a New Platonist to connect Euclid with his philosophy, as is clear from the next words in the same sentence, " for which reason also he set before himself, as the end of the whole Elements, the construction of the so-called Platonic figures." It is evident that it was only an idea of Proclus' own to infer that Euclid was a Platonist because his Elements end with the investigation of the five regular solids, since a later passage shows him hard put to it to reconcile the view that the construction of the five regular solids was the end and aim of the Elements with the obvious fact that they were intended to supply a foundation for the study of geometry in general, "to make perfect the understanding of the learner in regard to the whole of geometry*." To get out of the difficulty he says* that, if one should ask him what was the aim {o-kowos) of the treatise, he would reply by making a distinction between Euclid's intentions ( 1) as regards the subjects with which his investigations are concerned, (2) as regards the learner, and would say as regards ( 1 ) that " the whole of the geometer's argument is concerned with the cosmic figures." This latter statement is obviously incorrect It is true that Euclid's Elements end with the construction of the five regular solids ; but the pi an i metrical portion has no direct relation to them, and the arithmetical no relation at all ; the propositions about them are merely the conclusion of the stereo metrical division of the work. One thing is however certain, namely that Euclid taught, and founded a school, at Alexandria. This is clear from the remark of Pappus about Apollonius*: "he spent a very long time with the pupils of Euclid at Alexandria, and it was thus that he acquired such a scientific habit of thought," It is in the same passage that Pappus makes a remark which might, to an unwary reader, seem to throw some light on the 1 Proclus, p. 68, JO, ul tJ rpoatpim ti lUaTwiKii itrn cal r^i 0i\*& EvxXd&ov jukthjrait tr 'AX*(aj>?pr£p ch. i] EUCLID AND TRADITIONS ABOUT HIM 3 personality of Euclid. He is speaking about Apollonius' preface to the first book of his Conies, where he says that Euclid had not completely worked out the synthesis of the " three- and four-line locus," which in fact was not possible without some theorems first discovered by himself. Pappus says on this 1 : "Now Euclid — regarding Aristaeus as deserving credit for the discoveries he had already made in conies, and without anticipating him or wishing to construct anew the same system (such was his scrupulous fairness and his exemplary kindliness towards all who could advance mathematical science to however small an extent), being moreover in no wise con- tentious and, though exact, yet no braggart like the other [Apollonius] — wrote so much about the locus as was possible by means of the conies of Aristaeus, without claiming completeness for his demonstra- tions." It is however evident, when the passage is examined in its context, that Pappus is not following any tradition in giving this account of Euclid : he was offended by the terms of Apollonius' reference to Euclid, which seemed to him unjust, and he drew a fancy picture of Euclid in order to show Apollonius in a relatively unfavourable light. Another story is told of Euclid which one would like to believe true. According to Stobaeus*, "some one who had begun to read geometry with Euclid, when he had learnt the first theorem, asked Euclid, ' But what shall I get bylearning these things?' Euclid called his slave and said 'Give him threepence, since he must make gain out of what he learns.' " In the middle ages most translators and editors spoke of Euclid as Euclid of Megara. This description arose out of a confusion between our Euclid and the philosopher Euclid of Megara who lived about 400 B.C. The first trace of this confusion appears in Valerius Maximus (in the time of Tiberius) who says* that Plato, on being appealed to for a solution of the problem of doubling the cubical aitar, sent the inquirers to " Euclid the geometer." There is no doubt about the reading, although an early commentator on Valerius Maximus wanted to correct " Eucliden " into " Eudoxum? and this correction is clearly right. But, if Valerius Maximus took Euclid the geometer for a contemporary of Plato, it could only be through confusing him with Euclid of Megara. The first specific reference to Euclid as Euclid of Megara belongs to the 14th century, occurring in the v^rofivtifuiTitrfial of Theodorus Metochita (d. 1332) who speaks of " Euclid of Megara, the Socratic philosopher, contemporary of Plato," as the author of treatises on plane and solid geometry, data, optics etc. : and a Paris MS. of the 14th century has " Euctidis philosophi Socratici liber ele mentor urn," The misunderstanding was general in the period from Campanus 1 translation (Venice 1482) to those of Tartaglia (Venice 1565) and Candalla (Paris 1566). But one Constantinus Lascaris (d. about 1493) had already made the proper 1 Pappus, VII. pp. €76, 55 — 678* 6. Hultsch, it is true, brackets the whole passage ppu 676, i j— 678, I j, but apparently on the ground of the diction only, * Stobaeus, I.e. * vill. n, eit. i. 4 INTRODUCTION [ch. i distinction by saying of our Euclid that "he was different from him of Megara of whom Laertius wrote, and who wrote dialogues "' ; and to Commandinus belongs the credit of being the first translator* to put the matter beyond doubt : " Let us then free a number of people from the error by which they have been induced to believe that our Euclid is the same as the philosopher of Megara " etc Another idea, that Euclid was born at Gela in Sicily, is due to tne same confusion, being based on Diogenes Laertius* description' of the philosopher Euclid as being "of Megara, or, according to some, of Gela, as Alexander says in the AiaBoxai-" In view of the poverty of Greek tradition on the subject even as early as the time of Proclus (410-485 A.D.), we must necessarily take cum gram the apparently circumstantial accounts of Euclid given by Arabian authors ; and indeed the origin of their stories can be explained as the result (1) of the Arabian tendency to romance, and (2) of misunderstandings. We read* that " Euclid, son of Naucrates, grandson of Zenarchus", called the author of geometry, a philosopher of somewhat ancient date, a Greek by nationality domiciled at Damascus, born at Tyre, most learned in the science of geometry, published a most excellent and most useful work entitled the foundation or elements of geometry, a subject in which no more general treatise existed before among the Greeks : nay, there was no one even of later date who did not walk in his footsteps and frankly profess his doctrine. Hence also Greek, Roman and Arabian geometers not a few, who undertook the task of illustrating this work, published commentaries, scholia, and notes upon it, and made an abridgment of the work itself. For this reason the Greek philosophers used to post up on the doors of their schools the well-known notice : ' Let no one come to our school, who has not first learned the elements of Euclid.' " The details at the beginning of this extract cannot be derived from Greek sources, for even Proclus did not know anything about Euclid's father, while it was not the Greek habit to record the names of grandfathers, as the Arabians commonly did. Damascus and Tyre were no doubt brought in to gratify a desire which the Arabians always showed to connect famous Greeks in some way or other with the East. Thus Nasiraddin, the translator of the Elements, who was of Tus in Khurasan, actually makes Euclid out to have been " Thusinus " also*. The readiness of the Arabians to run away with an idea is illustrated by the last words 1 Letter to Fernandus Acuna, printed in Maurolycus, Hitteria Sitiliat, fol. 11 r. (see Heiberg, EuMid-Studieriy pp. n — 3, 35). ■ Preface to translation (Pisauri, 1571). * Diog. L. 11. 106, p. ;8 ed. Cobet. * Casiri, Bibliotheca Arabifo-Hispana Escuriaiensis, 1. p. 339, Casirt's source is al- Qifti (d. 1 148). trie author of the Ta'rttk ai-ffu&amd, a collection of biographies of phi- losophers, mathematicians, astronomers etc ■ The Fi'Aris/ says "son of Naucrates, the son of Berenice f/) " (see Suter"s translation in Abhandiungtn tur Gach. d. Math. V], Heft, 1H91. p. 16). e The same predilection made the Arabs describe 'Pythagoras as a pupil of the wise Salomo, Hippftrcbus as the exponent of Chaldaean philosophy or as the Chaldaean, Archi- medes as an Egyptian etc. (Hij! Khalfa, Lexicon BiMiogrspkitvm, and Casiri). ch. i) EUCLID AND TRADITIONS ABOUT HIM 5 of the extract. Everyone knows the story of Plato's inscription over the porch of the Academy : " let no one unversed in geometry enter my doors " ; the Arab turned geometry into Euclid 's geometry, and told the story of Greek philosophers in general and ''their Academies," Equally remarkable are the Arabian accounts of the relation of Euclid and Apollonius 1 . According to them the Elements were originally written, not by Euclid, but by a man whose name was Apollonius, a carpenter, who wrote the work in 1 5 books or sections'. In the course of time some of the work was lost and the rest became disarranged, so that one of the kings at Alexandria who desired to study geometry and to master this treatise in particular first questioned about it certain learned men who visited him and then sent for Euclid who was at that time famous as a geometer, and asked him to revise and complete the work and reduce it to order. Euclid then re- wrote h in 13 books which were thereafter known by his name. (According to another version Euclid composed the 1 3 books out of commentaries which he had published on two books of Apollonius on conies and out of introductory matter added to the doctrine of the five regular solids.) To the thirteen books were added two more books, the work of others (though some attribute these also to Euclid) which contain several things not mentioned by Apollonius. According to another version Hypsicles, a pupil of Euclid at Alexandria, offered to the king and published Books xiv. and xv., it being also stated that Hypsicles had " discovered " the books, by which it appears to be suggested that Hypsicles had edited them from materials left by Euclid. We observe here the correct statement that Books XIV. and xv. were not written by Euclid, but along with it the incorrect informa- tion that Hypsicles, the author of Book XIV., wrote Book XV. also. The whole of the fable about Apollonius having preceded Euclid and having written the Elements appears to have been evolved out of the preface to Book xiv. by Hypsicles, and in this way ; the Book must in early times have been attributed to Euclid, and the inference based upon this assumption was left uncorrected afterwards when it was recognised that Hypsicles was the author. The preface is worth quoting : "Basiiides of Tyre, O Protarchus, when he came to Alexandria and met my father, spent the greater part of his sojourn with him on account of their common interest in mathematics. And once, when 1 The authorities for these statements quoted by Casiri and lliji Khalfa are al-Kindl '3 tract de inttituto tibri Eutlidis (al-Kindl died about 873) and a commentary by Qidliide ar-Ruml (<[. about 1440) on a book called Askk&l ai-ttf sis (fundamental propositions) by Ashraf Shanuaddln aS'Samaru&ndi (c. 1176) consisting of elucidations of 3$ propositions ■elected from the first books of Euclid. Naslraddin likewise says that Euclid cut out two of 15 books of elements then existing and published the rest under his own name. According to Qadtzade the king heard that there was a celebrated geometer named Euclid at Tyrt\ Naslr- atldin says that he sent for Euclid of Tus. * So says the Fihrist, Suter {<>p- cit, p. 49) thinks that the author of the Fihrist did not suppose Apollonius of Ptrga to be the writer of the Ettmmts, as later Arabian authorities did, but that he disUnguished another Apollonius whom he calls " a carpenter." Suter's argument is based on the fact that the Fihrist s article on Apollonius (of Perga) says nothing of the Ebmttus, aiiJ that it gives the three great mathematicians, Euclid, Archimedes and Apollonius, in the correct chronological order. 6 INTRODUCTION [ch. i examining the treatise written by Apollonius about the comparison between the dodecahedron and the icosahedron inscribed in the same sphere, (showing) what ratio they have to one another, they thought that Apollonius had not expounded this matter properly, and accordingly they emended the exposition, as I was able to learn from my father. And I myself, later, fell in with another book published by Apollonius, containing a demonstration relating to the subject, and I was greatly interested in the investigation of the problem. The book published by Apollonius is accessible to all— for it has a large circulation, having apparently been carefully written out later — but I decided to send you the comments which seem to me to be necessary, for you will through your proficiency in mathe- matics in general and in geometry in particular form an expert judgment on what I am about to say, and you will lend a kindly ear to my disquisition for the sake of your friendship to my father and your goodwil! to me." The idea that Apollonius preceded Euclid must evidently have been derived from the passage just quoted. It explains other things besides. Basil ides must have been confused with j3a<7t\ev$, and we have a probable explanation of the " Alexandrian king," and of the " learned men who visited " Alexandria. It is possible also that in the " Tyrian " of Hypsicles" preface we have the origin of the notion that Euclid was born in Tyre. These inferences argue, no doubt, very defective knowledge of Greek : but we could expect no better from those who took the Organon of Aristotle to be " instrumentum musicum pneumaticum," and who explained the name of Euclid, which they variously pronounced as U elides or /eludes, to be com- pounded of Ueli a key, and Dis a measure, or, as some say, geometry, so that [/elides is equivalent to the key of geometry '! Lastly the alternative version, given in brackets above, which says that Euclid made the Elements out of commentaries which he wrote on two books of Apollonius on conies and prolegomena added to the doctrine of the five solids, seems to have arisen, through a like confusion, out of a later passage 1 in Hypsicles' Book XIV. : " And this is expounded by Aristaeus in the book entitled 'Comparison of the five figures,' and by Apollonius in the second edition of his comparison of the dodecahedron with the icosahedron." The "doctrine of the five solids" in the Arabic must be the "Comparison of the five figures" in the passage of Hypsicles, for nowhere else have we any information about a work bearing this title, nor can the Arabians have had. The reference to the two books of Apollonius on conies will then be the result of mixing up the fact that Apollonius wrote a book on conies with the second edition of the other work mentioned by Hypsicles. We do not find elsewhere in Arabian authors any mention of a commentary by Euclid on Apollonius and Aristaeus: so that the story in the passage quoted is really no more than a variation of the fable that the Elements were the work of Apollonius. 1 Heibtrg*5 Euclid, vol. v. p. 6. CHAPTER II. EUCLID'S OTHER WORKS. In giving a list of the Euclidean treatises other than the Elements, I shall be brief: for fuller accounts of them, or speculations with regard to thern, reference should be made to the standard histories of mathematics 1 . I will take first the works which are mentioned by Greek authors, i. The Pseudaria. I mention this first because Proclus refers to it in the general remarks in praise of the Elements which he gives immediately after the mention of Euclid in his summary. He says'; " But, inasmuch as many things, while appearing to rest on truth and to follow from scientific principles, really tend to lead one astray from the principles and deceive the more superficial minds, he has handed down methods for the discriminative understanding of these things as well, by the use of which methods we shall be able to give beginners in this study practice in the discovery of paralogisms, and to avoid being misled. This treatise, by which he puts this machinery in our hands, he entitled (the book) of Pseudaria, enumerating in order their various kinds, exercising our intelligence in each case by theorems of all sorts, setting the true side by side with the false, and combining the refutation of error with practical illustration. This book then is by way of cathartic and exercise, while the Elements contain the irrefragable and complete guide to the actual scientific investigation of the subjects of geometry." The book is considered to be irreparably lost. We may conclude however from the connexion of it with the Elements and the reference to its usefulness for beginners that it did not go outside the domain of elementary geometry'. 1 See, for example, Lena, Le seiente esatte neW antica Grata, 1914, pp. 145- — 168; T. L. Heath, History of Greek Mathematics, 191 1, J. pp. \i\ — 446. Cf- Heiberg, Litttrar- gatkUhtliche Studim tibar Euklid, pp. 36 — 153; Euclidis opera omnia, ed. Heiberg and Menge, Vols. VI. — VLI1, ■ Proclus, p. 70, i — 18. 3 Heiberg points out that Alexander Aphrodisiensis appears to allude to the work in his commentary on Aristotle's Sophistiti Ehncki {fol. 25 b) : "Not only those {fheyx *) which do not start from the principles of the science under which the problem is classed... but also those which do start from the proper principles of the science but in some respect admit a paralogism, e.g. the Ptevobgraphemaia of Euclid." Tannery {Butt, ties jeiencei ma/A- et aifr. »• Serie, vi., 1881, [*" Partie, p. 147) conjectures that it may be from this treatise that the : commentator got his Information about the quadratures of the circle by Antiphon and 8 INTRODUCTION [ch. n 2. The Data. The Data (Se&oftiva) are included by Pappus in the Treasury of Analysis (x6tto<; avrikvo/ievoi;), and he describes their contents 1 . They are still concerned with elementary geometry, though forming part of the introduction to higher analysis. Their form is that of pro- positions proving that, if certain things in a figure are given (in magnitude, in species, etc.), something else is given. The subject- matter is much the same as that of the planimetrical books of the Elements, to which the Data are often supplementary. We shall see this later when we come to compare the propositions in the Elements which give us the means of solving the general quadratic equation with the corresponding propositions of the Data which give the solution. The Data may in fact be regarded as elementary exercises in analysis. It is not necessary to go more closely into the contents, as we have the full Greek text and the commentary by Marinus newly edited by Menge and therefore easily accessible 1 . 3. The book On divisions {of figures). This work (irepl Statpetre&v fiiffklov) is mentioned by Proclus*. In one place he is speaking of the conception or definition (\6yo70-i 14 1), to whom he also attributes the Likcrjudci (? judicis) super detimum Euciidis translated by Gherard of Cremona. ' Dc atperficicrum divisionibui liber Mackomcto Bagdadino adscripUa, nunc primam /oannis Da Londitunsis tt Federici Ctsmmandini Urbinatti opera in lutein cditus, Pisauri, 1570, afterwards included in Gregory's Euclid (Oxford, r7oj). ch. n] EUCLID'S OTHER WORKS 9 found it in Latin in a ms, which was then in his own possession but was about 20 years afterwards stolen or destroyed in an attack by a mob on his house at Mortlake'. Dee, in his preface addressed to Commandinus, says nothing of his having translated the book, but only remarks that the very illegible ms. had caused him much trouble and (in a later passage) speaks of " the actual, very ancient, copy from which I wrote out..." (in ipso unde descripsi vetustissimo exemplari). The Latin translation of this tract from the Arabic was probably made by Gherard of Cremona (1 1 14- 1 1 87), among the list of whose numerous translations a " liber divisionum " occurs. The Arabic original cannot have been a direct translation from Euclid, and probably was not even a direct adaptation of it ; it contains mistakes and un mathematical expressions, and moreover does not contain the propositions about the division of a circle alluded to by Proclus. Hence it can scarcely have contained more than a fragment of Euclid's work. But Woepcke found in a MS. at Paris a treatise in Arabic on the division of figures, which he translated and published in 1851*. It is expressly attributed to Euclid in the MS. and corresponds to the description of it by Proclus. Generally speaking, the divisions are divisions into figures of the same kind as the original figures, e.g. of triangles into triangles ; but there are also divisions into " unlike " figures, e.g. that of a triangle by a straight line parallel to the base. The missing propositions about the division of a circle are also here: " to divide into two equal parts a given figure bounded by an arc of a circle and two straight lines including a given angle " and " to draw in a given circle two parallel straight lines cutting off a certain part of the circle." Unfortunately the proofs are given of only four propositions (including the two last mentioned) out of 36, because the Arabic translator found them too easy and omitted them. To illustrate the character of the problems dealt with I need only take one more example : " To cut off a certain fraction from a (parallel-) trapezium by a straight line which passes through a given point lying inside or outside the trapezium but so that a straight line can be drawn through it cutting both the parallel sides of the trapezium," The genuineness of the treatise edited by Woepcke is attested by the facts that the four proofs which remain are elegant and depend on propositions in the Elements, and that there is a lemma with a true Greek ring: "to apply to a straight line a rectangle equal to the rectangle contained by AB, AC and deficient by a square." Moreover the treatise is no fragment, but finishes with the words "end of the treatise," and is a well-ordered and compact whole. Hence we may safely conclude that Woepcke's is not only Euclid's own work but the whole of it. A restoration of the work, with proofs, was attempted by Ofterdinger 3 , who however does not give Woepcke's props. 30, 31, 34. 35. 36- VVe have now a satisfactory restoration, with ample notes 1 R. C. Archibald, Euclid's Boot: on the Division of Figures with a restoration based on Wotpehts text and on the Praetita geometriae of Leonardo Pisano^ Cambridge, 19(5, pp. 4 — 9. ' Journal A natiqut, |8JI, p. 133 sqq. 3 L. F. Ofterdinger, Beitrtige tur Wiederherslettung dcr Sehrifi dot Eu&lides iibrr die Tkeilmtg der Figurtn, Ulm, 1BJ3. io INTRODUCTION [ch. n and an introduction, by R. C. Archibald, who used for the purpose Woepcke's text and a section of Leonardo of Pisa's Practica geometriae (1220) 1 . 4. The Porisms. It is not possible to give in this place any account of the con- troversies about the contents and significance of the three lost books of Porisms, or of the important attempts by Robert Simson and Chasles to restore the work. These may be said to form a whole literature, references to which will be found most abundantly given by Heiberg and Loria, the former of whom has treated the subject from the philological point of view, most exhaustively, while the latter, founding himself generally on Heiberg, has added useful details, from the mathematical side, relating to the attempted restora- tions, etc' It must suffice here to give an extract from the only original source of information about the nature and contents of the Porisms, namely Pappus*. In his general preface about the books composing the Treasury of Analysis (towot dvaXuofttvos) he says : "After the Tan gene ies (of Apollonius) come, in three books, the Porisms of Euclid, [in the view of many] a collection most ingeniously devised for the analysis of the more weighty problems, [and] although nature presents an unlimited number of such porisms' 1 , [they have added nothing to what was written originally by Euclid, except that some before my time have shown their want of taste by adding to a few (of the propositions) second proofs, each (proposition) admitting of a definite number of demonstrations, as we have shown, and Euclid having given one for each, namely that which is the most lucid. These porisms embody a theory subtle, natural, necessary, and of considerable generality, which is fascinating to those who can see and produce results]. " Now all the varieties of porisms belong, neither to theorems nor problems, but to a species occupying a sort of intermediate position [so that their enunciations can be formed like those of either theorems or problems], the result being that, of the great number of geometers, some regarded them as of the class of theorems, and others of pro- blems, looking only to the form of the proposition. But that the ancients knew better the difference between these three things is clear from the definitions. For they said that a theorem is that which is proposed with a view to the demonstration of the very thing proposed, a problem that which is thrown out with a view to the construction of the very thing proposed, and a porism that which is proposed with a view to the producing of the very thing proposed. [But this definition of the porism was changed by the more recent writers who could not produce everything, but used these elements 1 There is a remarkable similarity between the propositions of Wot octet's text and those of Leonardo, suggesting that Leonardo may have had before him a translation (perhaps by Gherard of Cremona) of the Arabic tract. * Heiberg, liu&lid-Studiin, pp. 56 — 79, and Loria, op. at., pp. 1 S3 — jfij. ' Pappus, ed. Hullsch, VII. pp. 648 — 660. I put in square brackets the words bracketed by Haltsch. ' I adopt Heiberg 1 s reading of a comma here instead of a full stop. ch. it] EUCLID'S OTHER WORKS n and proved only the fact that that which is sought really exists, but did not produce it 1 and were accordingly confuted by the definition and the whole doctrine. They based their definition on an incidental characteristic, thus : A porism is that which falls short of a locus- theorem in respect of its hypothesis'. Of this kind of porisms loci are a species, and they abound in the Treasury of Analysis ; but this species has been collected, named and handed down separately from the porisms, because it is more widely diffused than the other species]. But it has further become characteristic of porisms that, owing to their complication, the enunciations are put in a contracted form, much being by usage left to be understood; so that many geometers understand them only in a partial way and are ignorant of the more essential features of their contents, "[Now to comprehend a number of propositions in one enunciation is by no means easy in these porisms, because Euclid himself has not in fact given many of each species, but chosen, for examples, one or a few out of a great multitude*. But at the beginning of the first book he has given some propositions, to the number of ten, of one species, namely that more fruitful species consisting of loci.] Consequently, finding that these admitted of being comprehended in one enunciation, we have set it out thus: If, in a system of four straight lines' which cut each other two and two, three points on one straight line be given while the rest except one lie on different straight lines given in position, the remaining point also will lie on a straight line given in position*. 1 Heiberg points out that Props. 5—9 of Archimedes' treatise On Spirals are porisms in this sense. To take Prop. 5 as an example, DBF is a tangent to a circle with centre K- It is then possible, says Archimedes, to draw a straight line r> b F KffF t meeting the circumference in // and the tangent in F, such that FH;HJC^{*kBH):c, where c is the circumference of any circle. To prove this he assumes the following construction. E being any straight line greater than c, he says 1 let KG be parallel to DF, "and let the line GH equal to E be placed verging to the point B." Archimedes must of course nave known how to effect this construction, which requires conies. But that it is possible requires very little argument, for if we draw any straight line BHG meeting the circle in /if and KG in G, it is obvious that as G moves away from C, HG becomes greater and greater and may be made as great as we please. The " later writers " would no doubt have contented themselves with this considera- tion without actually tanstrtuting HG. 1 As Heiberg says, this translation is made certain by a preceding passage of Pappus (p. 648, 1 — 3) where he compares two enunciations, the latter of which " falls short of the former in hypothesis but goes beyond it in requirement" E.g. the first enunciation requiring us, given three circles, to draw a circle touching all three, the second may require us, given only row circles (one less datum), to draw a circle touching them and ef a given sat (an extra requirement). * I translate Heiberg 's reading with a full stop here followed by r/wi ipxi U tiu*t [rpb* &PXV* ($t&Oneror) Hultsch] TOU TpwTOU /Sl^X/ou,,., 1 The four straight lines are described in the text as (the sides) brriov 1 9 wapuwrtout i.e. aides of two' sotts of quadrilaterals which Simson tries to explain {see p. no of the Index Grtueitatis of Hultsch's edition of Pappus). * In other words (Chasles, p, 13; Loria, p. 356), if a triangle be so deformed that each of its sides turns about one of three points in a straight line, and two of its vertices lie on two straight lines given in position, the third vertex will also lie on a straight line. is INTRODUCTION [ch. n "This has only been enunciated of four straight lines, of which not more than two pass through the same point, but it is not known (to most people) that it is true of any assigned number of straight lines if enunciated thus : If any number of straight lines cut one another, not more than two (passing) through the same point, and all the points (of intersection situated) on one of them be given, and if each of those which are on another (of them) lie on a straight line given in position — or still more generally thus : if any number of straight lines cut one another, not more than two (passing) through the same point, and all the points (of intersection situated) on one of them be given, while of the other points of intersection in multitude equal to a triangular number a number corresponding to the side of this triangular number lie respectively on straight lines given in position, provided that of these latter points no three are at the angular points of a triangle («. having for sides three of the given straight lines) — each of the remaining points will He on a straight line given in position'. ■ It is probable that fhe writer of the Elements was not unaware of this but that he only set out the principle ; and he seems, in the case of all the porisms, to have laid down the principles and the seed only [of many important things], the kinds of which should be distinguished according to the differences, not of their hypotheses, but of the results and the things sought [All the hypotheses are different from one another because they are entirely special, but each of the results and things sought, being one and the same, follow from many different hypotheses.] "We must then in the first book distinguish the following kinds of things sought : "At the beginning of the book' is this proposition : I. ' If from two given points straight tines be drawn meeting on a straight line given in position, and one cut off from a straight line given in position (a segment measured} to a given point on it, the other will also cut off from another {straight line a segment) liaving to the first a given ratio' " Following on this (we have to prove) II. that such and such a point lies on a straight line given in position ; III. that the ratio of such and such a pair of straight lines is given ; " etc. etc, (up to xxix.). "The three books of the porisms contain 38 lemmas; of the theorems themselves there are 171." 1 Loria (p. as6, ». 3) gives the meaning of this as follows, pointing out that Sim son vu the discoverer of it : " If a complete n-lateral be deformed so that its sides respectively turn about n points on a straight line, and {» ~ 1) of iEs n (« - i)/i vertices move on as many straight lines, the other (» - [)(»- j)/i of its vertices likewise move on as many straight lines ; but it is necessary that it should be impossible to form with the {ft - 1) vertices any triangle having for sides the sides of the polygon." 1 Reading, with Heiberg, roB pipMov [tuS f Hulischl. ch. n] EUCLID'S OTHER WORKS 13 Pappus further gives lemmas to the Porisms (pp. 866 — 918, ed. Hultsch). With Pappus' account of Forisms must be compared the passages of Proclus on the same subject Proclus distinguishes two senses in which the word ir6pierfia is used. The first is that of corollary where something appears as an incidental result of a proposition, obtained without trouble or special seeking, a sort of bonus which the investi- gation has presented us with 1 . The other sense is that of Euclid's Porisms*, In this sense* "porism is the name given to things which are sought, but need some finding and are neither pure bringing into existence nor simple theoretic argument. For (to prove) that the angles at the base of isosceles triangles are equal is a matter of theoretic argument, and it is with reference to things existing that such knowledge is (obtained). But to bisect an angle, to construct a triangle, to cut off, or to place — all these things demand the making of something ; and to find the centre of a given circle, or to find the greatest common measure of two given commensurable magnitudes, or the like, is in some sort between theorems and problems. For in these cases there is no bringing into existence of the things sought, but finding of them, nor is the procedure purely theoretic. For it is necessary to bring that which is sought into view and exhibit it to the eye. Such are the porisms which Euclid wrote, and arranged in three books of Porisms. Proclus' definition thus agrees well enough with the first, " older," definition of Pappus. A porism occupies a place between a theorem and a problem ; it deals with something already existing, as a theorem does, but has to find it (e.g. the centre of a circle), and, as a certain operation is therefore necessary, it partakes to that extent of the nature of a problem, which requires us to construct or produce some- thing not previously existing. Thus, besides III. 1 of the Elements and X. 3, 4 mentioned by Proclus, the following propositions are real porisms: III. 25, VL 11— 13, vn, 33, 34, 36, 39, vm. 2, 4, x. 10, XIII. 18. Similarly in Archimedes On the Sphere and Cylinder I. 2 — 6 might be called porisms. The enunciation given by Pappus as comprehending ten of Euclid's propositions may not reproduce the form of Euclid's enunciations ; but, comparing the result to be proved, that certain points lie on straight lines given in position, with the class indicated by II. above, where the question is of such and such a point lying on a straight line given in position, and with other classes, e.g. (v.) that such and such a line is given in position, (VI.) that such and such a line verges to a given point, (XXVII.) that there exists a given point such that straight lines drawn from it to such and such (circles) will contain a triangle given in species, we may conclude that a usual form of a porism was " to prove that it is possible to find a point with such and such a property" 1 Produsjpp. 111, 14; 301, 11. ■ ibid. p. ill, 11. "The term porism is used of certain problems, like the Ptrrisms written by Euclid." ' ibid. pp. joi, 15 sqq. i 4 INTRODUCTION [ch. it or "a straight line on which lie all the points satisfying given conditions " etc. Sim son defined a porism thus : " Porisma est propositi o in qua proponitur demonstrate rem aliquant, vel pi u res datas esse, cui, vet quibus, ut et cuilibet ex rebus innumeris, non quidem datis, sed quae ad ea quae data sunt eandem habent re lat ion em, con venire ostendendum est affectionem quandam communem in propositione descriptam 1 ." From the above it is easy to understand Pappus' statement that loci constitute a large class of porisms. A locus is well defined by Simson thus : " Locus est propositio in qua propositum est datam esse demonstrare, vel invenire lineam aut superficiem cuius quodlibet punctum, vel superficiem in qua quaelibet linea data lege descripta, communem quandam ha bet proprietatem in propositione descriptam," Heiberg cites an excellent instance of a loats which is a porism, namely the following proposition quoted by Eutocius' from the Plane Loci of Apollonius : " Given two points in a plane, and a ratio between unequal straight lines, it b possible to draw, in the plane, a circle such that the straight lines drawn from the given points to meet on the circumference of the circle have (to one another) a ratio the same as the given ratio," A difficult point, however, arises on the passage of Pappus, which says that a porism is " that which, in respect of its hypothesis, falls short of a locus-theorem " (rovacov StrnptftuiTos:). Heiberg explains it by comparing the porism from Apollonius' Plane Loci just given with Pappus' enunciation of the same thing, to the effect that, if from two given points two straight lines be drawn meeting in a point, and these straight lines have to one another a given ratio, the point will lie on either a straight line or a circumference of a circle given in position. Heiberg observes that in this latter enunciation something is taken into the hypothesis which was not in the hypothesis of the enunciation of the porism, viz. " that the ratio of the straight lines is the same." I confess this does not seem to me satisfactory : for there is no real difference between the enunciations, and the supposed difference in hypothesis is very like playing with words. Chasles says : " Ce qui constitue le porisme est ce qui manque a I'hypotklse d'un tkioreme local (en d'autres termes, le porisme est inferieur, par l'hypothese, au th^oreme local; e'est-a-dire que quand quelques parties d'une pro- position locale n'ont pas dans l'enonce la determination qui ieur est propre, cette proposition cesse d'etre regardee comme un th^oreme et devient un porisme)." But the subject still seems to require further elucidation. While there is so much that is obscure, it seems certain (i) that the Porisms were distinctly part of higher geometry and not of elementary 1 This was thus expressed by Chasles : n Le porisme est une proposition dans laquelle on demande de demontrer qu'une chose ou plusieurs choses sont denrUts, qui, ainsi que Tone 3uelconque d'une infinite d'autres choses non donnees, jnais dont chacune est avec des choses onntes dans une meme relation, ont une certaine propnete commune, decrite dans 1ft pro- position." * Commentary on Apollonius" Conies (vol. tt. p. ISO, ed. Heiberg). ch. ii] EUCLID'S OTHER WORKS 15 geometry, (a) that they contained propositions belonging to the modern theory of transversals and to projective geometry. It should be remembered too that it was in the course of his researches on this subject that Chasles was led to the idea of anharmonk ratios. Lastly, allusion should be made to the theory of Zeuthen 1 on the subject of the porisms. He observes that the only porjsm of which Pappus gives the complete enunciation, " If from two given points straight lines be drawn meeting on a straight line given in position, and one cut off from a straight line given in position (a segment measured) towards a given point on it, the other will also cut off from another (straight line a segment) bearing to the first a given ratio," is also true if there be substituted for the first given straight line a conic regarded as the "locus with respect to four lines," and that this extended porism can be used for completing Apollonius' exposition of that locus. Zeuthen concludes that the Porisms were in part by- products of the theory of conies and in part auxiliary means for the study of conies, and that Euclid called them by the same name as that applied to corollaries because they were corollaries with respect to conies. But there appears to be no evidence to confirm this conjecture. S- The Surface4oci (tottoi -jrooi hirt^iaveia). The two books on this subject are mentioned by Pappus as part of the Treasury of Analysis 1 . As the other works in the list which were on plane subjects dealt only with straight lines, circles, and conic sections, it is a priori likely that among the loci in this treatise (loci which are surfaces) were included such loci as were cones, cylinders and spheres. Beyond this all is conjecture based on two lemmas given by Pappus in connexion with the treatise. (1) The first of these lemmas 1 and the figure attached to it are not satisfactory as they stand, but a possible restoration is indicated by Tannery*. If the latter is right, it suggests that one of the loci contained all the points on the elliptical parallel sections of a cylinder and was therefore an oblique circular cylinder. Other assumptions with regard to the conditions to which the lines in the figure may be subject would suggest that other loci dealt with were cones regarded as containing all points on particular elliptical parallel sections of the cones", (2) In the second lemma Pappus states and gives a complete proof of the focus-and-directrix property of a conic, viz. that the locus of a point whose distance front a given point is in a given ratio to its distance from a fixed line is a conic section, which is an ellipse, a parabola or a hyperbola according as the given ratio is less than, equal to, or greater than unity'. Two conjectures are possible as to the application of this theorem in Euclid's Surface-loci, {a) It may have been used to prove that the locus of a point whose distance from a given straight 1 Die Lekrcvwt den Kegilsiknitttn im Aliertttm, chapter VIII. * Pappus, vn. p. 636. " Hid- TO p- 1004. * Mullciin da seieneei math, eJ astro*., J* Sirie, VI. 1+9. 1 Further particulars will be found in The Works ef Archimedes, pp. litii — Ixiv, and in Zeuthen, Die Lthre von den KtgeUthttitun, p. 415 sqq. ' Pappus, Vtt. pp. j 006 — 1014, and Hultsdi's Appendix, pp. 1170 — 3. 1 6 INTRODUCTION [en. 11 line is in a given ratio to its distance from a given plane is a certain cone, (b) It may have been used to prove that the locus of a point whose distance from a given point is in a given ratio to its distance from a given plane is the surface formed by the revolution of a conic about its major or conjugate axis 1 . Thus Chasles may have been correct in his conjecture that the Surface-loci dealt with surfaces of revolution of the second degree and sections of the same". 6. The Conks. Pappus says of this lost work: "The four books of Euclid's Conies were completed by Apollonius, who added four more and gave us eight books of Conies'." It is probable that Euclid's work was lost even by Pappus' time, for he goes on to speak of "Aristaeus, who wrote the still extant five books of Solid Loci connected with the conies." Speaking of the relation of Euclid's work to that of Aristaeus on conies regarded as loci, Pappus says in a later passage (bracketed however by Huttsch) that Euclid, regarding Aristaeus as deserving credit for the discoveries he had already made in conies, did not (try to) anticipate him or construct anew the same system. We may no doubt conclude that the book by Aristaeus on solid loci preceded Euclid's on conies and was, at least in point of originality, more important Though both treatises dealt with the same subject-matter, the object and the point of view were different ; had they been the same, Euclid could scarcely have refrained, as Pappus says he did, from attempting to improve upon the earlier treatise. No doubt Euclid wrote on the general theory of conies as Apollonius did, but confined himself to those properties which were necessary for the analysis of the Solid Loci of Aristaeus. The Conks of Euclid were evidently superseded by the treatise of Apollonius. As regards the contents of Euclid's Conks, the most important source of our information is Archimedes, who frequently refers to propositions in conies as well known and not needing proof, adding in three cases that they are proved in the " elements of conies " or in "the conies," which expressions must clearly refer to the works of Aristaeus and Euclid 4 Euclid still used the old names for the conies (sections of a right- angled, acute 'angled, or obtuse-angled cone), but he was aware that an ellipse could be obtained by cutting a cone in any manner by a plane not parallel to the base (assuming the section to lie wholly between the apex of the cone and its base) and also by cutting a cylinder. This is expressly stated in a passage from the Pkaettontena of Euclid about to be mentioned". 7. The Pkaenometta. This is an astronomical work and is still extant. A much inter- 1 For further details see The Works of Arckimtd(s x pp. Ixiv, Lev, and Zeuthen, I. c. * Apcrfu hhtoriqtu, pp. 373 — 4. * Pappus, VII. p. 673. * For details of these propositions see my Apollonius of Perga, pp. xxxv, xxxvi. * Phattamtna, ed. Menge, p. 6: "if a cone or a cylinder be cut by a plane not parallel to the base, the section is a section of an acute-angled cone, which is like a shield \evftit)." ch. n] EUCLID'S OTHER WORKS 17 polated version appears in Gregory's Euclid. An earlier and better recension is however contained in the Ms. Vindobonensis philos. Gr. 103, though the end of the treatise, from the middle of prop. 16 to the last (18), is missing. The book, now edited by Menge 1 , consists of propositions in spheric geometry. Euclid based it on Autolycus' work Trepi Ktvov/ievTj? ecftatpas, but also, evidently, on an earlier text- book of Spkaerica of exclusively mathematical content. It has been conjectured that the latter textbook may have been due to Eudoxus'. 8. The Optics. This book needs no description, as it has been edited by Heiberg recently', both in its genuine form and in the recension by Theon. The Catoptrica published by Heiberg in the same volume is not genuine, and Heiberg suspects that in its present form it may be Theon's. It is not even certain that Euclid wrote Catoptrica. at all, as Proclus may easily have had Theon's work before him and inadvertently assigned it to Euclid 1 . 9. Besides the above-mentioned works, Euclid is said to have written the Elements of Music* (at tcara ftovatKrjv a"rotYety." We know that in the Neo-Platonic school the younger pupils learnt mathematics ; and it is clear that Proclus taught this subject, and that this was the origin of the commentary. Many passages show him as a master speaking to scholars. Thus "we Have illustrated 1 Zeller calls him "Der Gelehrtc, dem kern Feld damaligen Wissens verschlossen ist." * Van Pesch observes that in his commentaries on the Timaeus (pp. 671 — 2) he speaks as no real mathematician could have spoken. In the passage referred to the question is whether the sun occupies a middle place among the planets. Proclus rejects the view of Hipparchus and Ptolemy because "& Vtwpy6t" (sc. the Chaldean, says Zeller} thinks otherwise, "whom it is not lawful to disbelieve-" Martin says rather neatly, " Pour Proclus, leg Elements d'Euclide ont l 1 hen reuse chance de n'elre contredits ni par fes Oracles chaldalques, ibi par les speculations des pythagoriciens anciens et nouveaux ,.,,,." * Proclus, p. 84, 13. ' ibid. p. 429, tit, 6 ibid. p. 31, 20. ■ ibid. p. 31, «. 7 ibid. p. 37, J7 to 18, 7; cf. also p, »i, 15, pp. 46, +7. ch. iv] PROCLUS AND HIS SOURCES 31 and made plain all these things in the case of the first problem, but it is necessary that my hearers should make the same inquiry as regards the others as well '," and " I do not indicate these things as a merely incidental matter but as preparing us beforehand for the doctrine of the Timaeus 1 .* 1 Further, the pupils whom he was addressing were beginners in mathematics ; for in one place he says that he omits "for the present" to speak of the discoveries of those who employed the curves of Nicomedes and Hippias for trisecting an angle, and of those who used the Archimedean spiral for dividing an angle in any given ratio, because these things would be too difficult for beginners (Svff&ewpijTovs to« tla-ayoftivoi^)'. Again, if his pupils had not been beginners, it would not have been necessary for Proclus to explain what is meant by saying that sides subtend certain angles', the difference between adjacent and vertical angles' etc., or to exhort them, as he often does, to work out other particular cases for themselves, for practice {yvpvaaias ihteica)'. The commentary seems then to have been founded on Proclus' lectures to beginners in mathematics. But there are signs that it was revised and rt -edited for a larger public ; thus he gives notice in one place' "to those who shall come upon" his work (roh ivrev^o- fiivov;). There are also passages which could not have heen under- stood by the beginners to whom he lectured, e.g. passages about the cylindrical helix', conchoids and cissoids*. These passages may have been added in the revised edition, or, as van Pesch conjectures, the explanations given in the lectures may have been much fuller and more comprehensible to beginners, and they may haw; been shortened on revision. In his comments on the propositions of Euclid, Proclus generally proceeds in this way : first he gives explanations regarding Euclid's proofs, secondly he gives a few different cases, mainly for the sake of practice, and thirdly he addresses himself to refuting objections raised by cavillers to particular propositions. The latter class of note he deems necessary because of "sophistical cavils" and the attitude of the people who rejoiced in finding paralogisms and in causing annoyance to scientific men". His commentary does not seem to have been written for the purpose of correcting or improving Euclid. For there are very few passages of mathematical content in which Proclus can be supposed to be propounding anything of his own ; nearly all are taken from the works of others, mostly earlier commentators, so that, for the purpose of improving on or correcting Euclid, there was no need for his commentary at all. Indeed only in one place does he definitely bring forward anything of his own to get over a difficulty which he finds in Euclid"; this i, c where he tries to I Proclm, p. no, 1 9. ' il/id. p. 384, a. * ibid. p. 171, 11. * ibid, p. 338, 1*. * ibid.?. 19S, 14. * Cf. p. 114, is {on i. »). * ibid. p. 84, 9. ' ibid- p. 105. * ibid. p. 1 ti. " 'bid. p. 375, 9. II ibid. pp. 368—373. 32 INTRODUCTION [ch. iv prove the parallel-postulate, after first giving Ptolemy's attempt and then pointing out objections to it. On the other hand, there are a number of passages in which he extols Euclid; thrice 1 also he supports Euclid against Apollonius where the latter had given proofs which he considered better than Euclid's (I. 10, II, and 23). Allusion must be made to the debated question whether Proclus continued his commentaries beyond Book I, His intention to do so is clear from the following passages. Just after the words above quoted about the trisection etc. of an angle by means of certain curves he says, " For we may perhaps more appropriately examine these things on the third book, where the writer of the Elements bisects a given circumference*." Again, after saying that of all parallelograms which have the same periifieter the square is the greatest " and the rhomboid least of all," he adds : "But this we will prove in another place ; for it is more appropriate to the (discussion of the) hypotheses of the second book'." Lastly, when alluding (on I. 45) to the squaring of the circle, and to Archimedes' proposition that any circle is equal to the right-angled triangle in which the perpendicular is equal to the radius of the circle and the base to its perimeter, he adds, "But of this elsewhere* " ; this may imply, an intention to treat of the subject on Eucl. XII., though Heiberg doubts It*. But it is clear that, at the time when the commentary on Book I. was written, Proclus had not yet begun to write on the other Books and was uncertain whether he would be able to do so : for at the end he says", " For my part, if I should be able to discuss the other books' in the same manner, I should give thanks to the gods ; but, if other cares should draw me away, I beg those who are attracted by this subject to complete the exposition of the other books as well, following the same method, and addressing themselves throughout to the deeper and better defined questions involved " (to irpayfunetaiBe<..■!, I. came from Proclus, himself attached Proclus' name to the others- ch. iv] PROCLUS AND HIS SOURCES 33 MSS. as to suggest that the scholiast had further commentaries of Proclus which have vanished for us 1 ; (2; that there is no trace in the scholia of the notes which Proclus promised in the passages quoted above. Coming now to the question of the sources of Proclus, we may say that everything goes to show that his commentary is a compilation, though a compilation "in the better sense" of the term". He does not even give us to understand that we shall find in it much of his own ; " let us," he says, " now turn to the exposition of the theorems proved by Euclid, selecting the more subtle of the comments made on them by the ancient writers, and cutting down their interminable diffuse- ness...*": not a word about anything of his own. At the same time, he seems to imply that he will not necessarily on each occasion quote the source of each extract from an earlier commentary ; and, in fact, while he quotes the name of his authority in many places, especially where the subject is important, in many others, where it is equally certain that he is not giving anything of his own, he mentions no authority. Thus he quotes Heron by name six times ; but we now know, from the commentary of an-Nairizi, that a number of other passages, where he mentions no name, are taken from Heron, and among them the. not unimportant addition of an alternative proof to I. 19. Hence we can by no means conclude that, where no authority is mentioned, Proclus is giving notes of his own. The presumption is generally the other way ; and it is often possible to arrive at a con- clusion, either that a particular note is not Proclus' own, or that it is definitely attributable to someone else, by applying the ordinary principles of criticism. Thus, where the note shows an unmistakable affinity to another which Proclus definitely attributes to some com- mentator by name, especially when both contain some peculiar and distinctive idea, we cannot have much doubt in assigning both to the same commentator*. Again, van Pesch finds a criterion in the form of a note, where the explanation is so condensed as to be only just intelligible; the note is that in which a converse of I. 32 is proved" the proposition namely that a rectilineal figure which has all its in- terior angles together equal to two right angles is a triangle. It is not safe to attribute a passage to Proclus himself because he uses the first person in such expressions as " I say " or " I will prove " — for he was in the habit of putting into his own words the substance of notes borrowed from others — nor because, in speaking of an 1 While one class or scnuua (Schol. Vat.) have some better readings than our MSS. of Proclus h ave, and partly Fill up the gaps at 1. 36, 37 and I. 41 — 43, the other class (SchgL Vind.) derive from an inferior Proclus MS, which also had the same lacunae. a Knoche, UnUrsttchungcn iibcr dti Proklui Diattoihui Commtntar zu EttktidTi Eie- mentm {186?), p. 1 1. 3 Proclus, p. 30o, to — 13. 4 Instances of the application of this criterion will be found in the discussion of Proclus' indebtedness to the commentaries of Heron, Porphyry and Pappus, * Van Pesch attributes this converse and proof to Pappus, arguing from the fact that the proof is followed by a passage which, on comparison with Pappus' note on the postulate that all right angles are equal, he feels justified in assigning to Pappus, I doubt if the evidence is sufficient. 34 INTRODUCTION [ch. iv objection raised to a particular proposition, he uses such expressions as "perhaps someone may object" («rwe 8* &v tii>« it/trraUv...): for sometimes other words in the same passage, indicate that the objection had actually been taken by someone 1 . Speaking generally, we shall not be justified in concluding that Prod us is stating something new of his own unless he indicates this himself in express terms. As regards the form of Proclus' references to others by name, van Pesch notes that he very seldom mentions the particular zvork from which he is borrowing. If we leave out of account the references to Plato's dialogues, there are only the following references to books : the Bacchae of Philolaus 1 , the Symmikta of Porphyry*, Archimedes On the Sphere and Cylinder 1 , Apollonius On the cochlias*, a book by Eudemus on The Angle', a whole book of Posidonius directed against Zeno of the Epicurean sect', Carpus' Astronomy', Eudemus' History of Geometry*, and a tract by Ptolemy on the parallel-postulate 10 . Again, Proclus does not always indicate that he is quoting some- thing at second-hand. He often does so, e.g. he quotes Heron as the authority for a statement about Phi lip pus, Eudemus as attributing a certain theorem to Oenopides etc. ; but he says on i. 1 2 that " Oeno- pides first investigated this problem, thinking it useful for astronomy " when he cannot have had Oenopides' work before him. It has been said above that Proclus was in the habit of stating in his own words the substance of the things which he borrowed. We are prepared for this when we find him stating that he will select the best things from ancient commentaries and "cut short their intermin- able diffuseness," that he will "briefly describe" (oiWTo/ttif laTopqaai) the other proofs of 1. 20 given by Heron and Porphyry and also the proofs of I, 25 by Menelaus and Heron. But the best evidence is of course to be found in the passages where he quotes works still extant, e.g. those of Plato, Aristotle and Plotinus. Examination of these passages shows great divergences from the original; even where he purports to quote textually, using the expressions " Plato says," or " Plotinus says," he by no means quotes word for word". In fact, he seems to have had a positive distaste for quoting textually from other works. He cannot conquer this even when quoting from Euclid ; he says in his note on I. 22, " we will follow the words of the geometer " but fails, nevertheless, to reproduce the text of Euclid unchanged". We now come to the sources themselves from which Proclus drew I Van Pesch illustrates this by ;ui objection refuted in the note on I. 9, p. 1JJ, 11 sqq. After using the above expression to introduce the objection, Proclus uses further on (p. »73,ij) the term "they say" ijnelr). • Proclus, p. »a, 15. * ibid, p, j6, ij. * ibid. p. 71, 18. * ibid. p. 105, 3. ibid. p. j 35, 8. T ibid. p. aoo, 3. ■ ibid. p. j+i, 19. "' Hid. p. 351, tj. 10 ibid. p. 361, r5, II See the passages referred to by van Pesch (p. 70}. The most glaring case is a passage (p. 31, 10) where he quotes Plotinus, using the expression " Plotinus says......" Comparison with Plotinus. Hnntad. t. 3, 3, shows that very few words are those of Plotinus himself; the rest represent Plotinus' views in Proclus' own language. 11 Proclus, p. 330, 19 sqq ch. iv] PROCLUS AND HIS SOURCES 35 in writing his commentary. Three have already been disposed of, viz. Heron, Porphyry and Pappus, who had all written commentaries on the Elements 1 . We go on to Eudemus, the pupil of Aristotle, who, among other works, wrote a history of arithmetic, a history of astronomy, and a history of geometry. The importance of the last mentioned work is attested by the frequent use made of it by ancient writers. That there was no other history of geometry written after the time of Eudemus seems to be proved by the remark of Proclus in the course of his famous summary r " Those who compiled histories bring the development of this science up to this point. Not much younger than these is Euclid*. . .." The loss of Eudemus' history is one of the gravest which fate has inflicted upon us, for it cannot be doubted that Eudemus had before htm a number of the actual works of earlier geometers, which, as before observed, seem to have vanished completely when they were superseded by the treatises of Euclid, Archimedes and Apoilonius. As it is, we have to be thankful for the fragments from Eudemus which such writers as Proclus have preserved to us. I agree with van Pesch' that there is no sufficient reason for doubting that the work of Eudemus was accessible to Proclus at first hand. For the later writers Simplicius and Eutocius refer to it in terms such as leave no room for doubt that tftey had it before them. I have already quoted a passage from Simplicius' account of the lunes of Hippocrates to the effect that Eudemus must be considered the best authority since he lived nearer the times*. In the same place Simplicius says", " I will set out what Eudemus says word for word (tcwrh. \il-tv Xeyofifva), adding only a little explanation in the shape of reference to Euclid's Elements owing to the memorandum-like style of Eudemus (St a tov iriroji.viifi.aTticbv Tpoirov rov EvSjj/*ou) who sets out his explanations in the abbreviated form usual with ancient writers. Now in the second book of the history of geometry he writes as follows'." It is not possible to suppose that Simplicius would have written in this way about the style of Eudemus if he had merely been copying certain passages second-hand out of some other author and had not the original work itself to refer to. In like manner, Eutocius speaks of the paralogisms handed down in connexion with the attempts of Hippocrates and Antiphon to square; the circle", "with which I imagine that those are accurately acquainted who have examined (^reo-Ke^tei-ow) the geometrical history of Eudemus and know the Ceria Aristotelica." How could the contemporaries of Euto- cius have examined the work of Eudemus unless it was still extant in his time ? The passages in which Proclus quotes Eudemus by name as his authority are as follows : (l) On I. 26 he says that Eudemus in his history of geometry 1 See pp. 30 to a 7 above. * l*roclus, p. 68, 4—7. * De Prodi fontibui. pp. ft— Jg. * See above, p. 19. ' Simplicius, Jet. tit., ed. Dids, p. 60, 17. * Archimedes, ed. Heiberg, vol. ill. p. 11&. 36 INTRODUCTION [ch. iv referred this theorem to Thales, inasmuch as it was necessary to Thales 1 method of ascertaining the distance of ships from the shore 1 . (2) Eudemus attributed to Thales the discovery of Eucl. I. 15", and (3) to Oenopides the problem of I. 23*. (4) Eudemus referred the discovery of the theorem in 1. 32 to the Pythagoreans, and gave their proof of it, which Proclus reproduces*. (5) On I. 44 Proclus tells us* that Eudemus says that "these things are ancient, being discoveries of the Pythagorean muse, the application (irapafioXrf) of areas, their exceeding (vTrcpfioXij) and their falling short (lA-Xe^ec)," The next words about the appro- priation of these terms (parabola, hyperbola and ellipse) by later writers (i.e. Apollonius) to denote the conic sections are of course not due to Eudemus. Coming now to notes where Eudemus is not named by Proclus, we may fairly conjecture, with van Pesch, that Eudemus was really the authority for the statements (1) that Thales first proved that a circle is bisected by its diameter* (though the proof by reductio ad absurdum which follows in Proclus cannot be attributed to Thales 7 ), (z) that " Plato made over to Leodamas the analytical method, by means of which it is recorded (ltrTopi)Tftova.y°, (5) that the theorem that only three sorts of polygons can fill up the space round a point, viz. the equilateral triangle, the square and the regular hexagon, was Pythagorean". Eudemus may also be the authority for Proclus' description of the two methods, referred to Plato and Pythagoras respectively, of forming right-angled triangles in whole numbers". We cannot attribute to Eudemus the beginning of the note on I. 47 where Proclus says that "if we listen to those who like to recount ancient history, we may find some of them referring this theorem to Pythagoras and saying that he sacrificed an ox in honour of his discovery"." As such a sacrifice was contrary to the Pytha- gorean tenets, and Eudemus could not have been unaware of this, the story cannot rest on his authority. Moreover Proclus speaks as though he were not certain of the correctness of the tradition ; indeed, 1 Proclus, p. 331, 14 — 18. ■ ibid, p. 199, 3. * ibid. p. 333, 5. * Hid. p. 379, 1—16. * Hid. p. + iy, ij — 18. * ibid. p. 157, 10, 11. 7 Cantor [GcscA, d. Math. Is, p. ait) points out the connexion between the rrducth ad absurdum And the Analytical method said to have been discovered by Plato, Proclus gives the proof by rfduclio ad absurdum to meet an imaginary critic who desires a mathematical proof ; possibly Thales may have been satisfied with the argument in the same sentence which mentions Thales, "the cause o r the bisection being the unswerving course of the straight line through the centre." * Proclus, p. 31 1, 19 — 13. * ibid. p. 350, 10. " ibid. p. 383, ; — 10. '* Vrid. pp. 304, 11 — 303, 3. u ibid. pp. 418, 7 — 419, 9. a Hid. p. 416, 6 — 9. ch. iv] PROCLUS AND HIS SOURCES 37 so far as the story of the sacrifice is concerned, the same thing is told of Thales in connexion with his discovery that the angle in a semi- circle is a right angle 1 , and Plutarch is not certain whether the ox was sacrificed on the discovery of I. 47 or of the problem about application of areas*. Plutarch's doubt suggests that he knew of no evidence for the story beyond the vague allusion in the distich of Apollodorus " Logisticus " (the " calculator ") cited by Diogenes Laertius also'; and Proclus may have had in mind this couplet with the passages of Plutarch. We come now to the question of the famous historical summary given by Proclus*. No one appears to maintain that Eudemus is the author of even the early part of this summary in the form in which Proclus gives it. It is, as is well known, divided into two distinct parts, between which comes the remark, " Those who compiled histories* bring the development of this science up to this point. Not much younger than these is Euclid, who put together the Elements, collecting many of the theorems of Eudoxus, perfecting many others by Theaetetus, and bringing to irrefragable demonstration the things which had only been somewhat loosely proved by his pre- decessors." Since Euclid was later than Eudemus, it is impossible that Eudemus can have written this. Yet the style of the summary after this point does not show any such change from that of the former portion as to suggest different authorship. The author of the earlier portion recurs frequently to the question of the origin of the elements of geometry in a way in which no one would be likely to do who was not later than Euclid ; and it must be the same hand which in the second portion connects Euclid's Elements with the work of Eudoxus and Theaetetus'. If then the summary is the work of one author, and that author not Eudemus, who is it likely to have been ? Tannery answers that it is Geminus 1 ; but I think, with van Pesch, that he has failed to show why it should be Geminus rather than another. And certainly the extracts which we have from Geminus' work suggest that the sort of topics which it dealt with was quite different ; they seem rather to have been general questions of the content of mathematics, and even Tannery admits that historical details could only have come inci- dentally into the work'. Could the author have been Proclus himself? Circumstances 1 Diogenes Laertius, I. 14, p. 6, ed. Cobet. 1 Plutarch, nen posse tttaviter vivi secundum Epicurum, 1 1 ; Symp- VIM, 1. ■ Ding, Laert. vm. is, p. 107, ed. Cobet: 'livipta Hv0ay6pnt ri refHK\tit fGpcTo fp&fifia, Kfii 1$ Sertfi x\uvi)v ifyaye @w0 vv^v. See on this subject Tannery, La Giomttrii gr&cqu4 t p. 105. * Proclus, pp. 64 — 70. * The plural is well explained by Tannery, La Giomitrii gricqut t pp. 73, 74. No doubt the author of the summary tried to supplement Eudemus by means of any other histories which threw light on the subject. Thus e.g. the allusion {p. 64, 11) to the Nile recalls Herodotus. Cf. the depression in Proclus, p. 64, 19, rapi Tar roXkar iirripijTiu. * Tannery, La Gimiitrit grtequt, p. ;j. ' Hid. pp. 66 — 7 j. ' ibid. p. in. 38 INTRODUCTION [ch, iv which seem to suggest this possibility are (i) that, as already stated, the question of the origin of the Elements is kept prominent, (2) that there is no mention of Democritus, whom Eudemus would not be likely to have ignored, while a fol tower of Plato would be likely enough to do him the injustice, following the example of Plato who was an opponent of Democritus, never once mentions him, and is said to have wished to burn all his writings 1 , and (3) the allusion at the beginning to the "inspired Aristotle" (0 Baifiovux; 'Apto-ToTeXi;?)*, though this may easily have been inserted by Proclus in a quotation made by him from someone else. On the other hand there are considerations which suggest that Proclus himself was not the writer. (1) The style of the whole passage is rtot such as to point to him as the author. (2) If he wrote it, it is hardly conceivable that he would have passed over in silence the discovery of the analytical method, the invention of Plato to which he attached so much importance*. There is nothing improbable in the conjecture that Proclus quoted the summary from a compendium of Eudemus' history made by some later writer: but as yet the question has not been definitely settled. All that is certain is that the early part of the summary must have been made up from scattered notices found in the great work of Eudemus. Proclus refers to another work of Eudemus besides the history, viz. a book on The Angle [0t0Xlov wepl ya>vla<;)*. Tannery assumes that this must have been part of the history, and uses this assumption to confirm his idea that the history was arranged according to subjects, not according to chronological order*. The phraseology of Proclus however unmistakably suggests a separate work ; and that the history was chronologically arranged seems to be clearly indicated by the remark of Simplicius that Eudemus "also counted Hippocrates among the more ancient writers " {iv tok vd\atoT4poi<;) a . The passage of Simplicius about the lunes of Hippocrates throws considerable light on the style of Eudemus' history. Eudemus wrote in a memorandum-like or summary manner {jov vTrofivrifiaTiKov rpoirov tow EuS?jjM>ii)' when reproducing what he found in the ancient writers; sometimes it is clear that he left out altogether proofs or constructions of things by no means easy'. Gem in u 9. The discussions about the date and birthplace of Geminus form a whole literature, as to which 1 must refer the reader to Manitius and Tittel', Though the name looks like a Latin name (Gemlnus), Mani- 1 Diog. Laertius, Ix. 40, p. 337, ltd. Cobet. - Proclus, p. 64, 8. * Proclus, p. in, [9 sqq. j the passage is quoted above, p. 36. * ibid, p. 125, 8. i Tannery, La Giotnltrit grettjite, p. 26. * Simplicius, «d. Diels, p. 60, 23. ' ibid, p. - 60, 29, * Cf. Simplicius, p. 63, 19 sqq. ; p. 64. 15 sqo. ; also Usener's note *'dc supplcndis Hippocratis quas omisit Eudemus cons tructioni bus added to Diels' preface, pp. xxiii — xxvi. 'Manitius, Gemini ticmenia. astronotnitM (Tcubner, 1898), pp. 237- — 151; Tittel, ait. *' Geminos " in Pauly-Wissowa's Rtal-Eiuytiopiidif. dtr ttassischm Aitertutnswisiemehaft, vol. Vlt.. 1910. ch. tv] PROCLUS AND HIS SOURCES 39 tius concluded that, since it appears as VeftZvos in all Greek MSS. and as rf^Ktiio? in some inscriptions, it is Greek and possibly formed from 7e/A as 'Epyiuo? is from ipy and *AX.ef wot from aXef (cf. also 'I*tm'os, KpaTivos). Tittel is equally positive that it is Gemtnus and suggests that Te/iifo? is due to a false analogy with 'AXeftetx? etc. and Fe/tetpoc wrongly formed on the model of 'Avravuvo?, ' kypwrreiva. Geminus, a Stoic philosopher, born probably in the island of Rhodes, was the author of a comprehensive work on the classification of mathematics, and also wrote, about 73-67 B.C., a not !ess comprehensive commentary on the meteorological textbook of his teacher Pnsidonius of Rhodes. It is the former work in which we are specially interested here. Though Proclus made great use of it, he does not mention its title, unless we may suppose that, in the passage (p. 177, 24) where, after quoting from Geminus a classification of lines which never meet, he says, " these remarks I have selected from the tyXoKoKia of Geminus," (ptXoxaXla is a title or an alternative title. Pappus however quotes a work of Geminus "on the classification of the mathematics" (iv t$ wept Tt}<; t&v naffyfidrtov rafetus-) 1 , while Eutocius quotes from *' the sixth book of the doctrine of the mathematics" (tv ra> eVro) 1770 t&v fiaffiHidraiv QtwpiasV Tannery' pointed out that the former title corresponds well, enough to the long extract* which Proclus gives in his first prologue, and also to the fragments contained in the Anonymi variae collectiones published by Hultsch at the end of his edition of Heron'; but it does not suit most of the o'ther passages borrowed by Proclus, The correct title was therefore probably that given by Eutocius, The Doctrine, or Theory, of the Mathematics ; and Pappus probably refers to one particular portion of the work, say the first Book. If the sixth Book treated of conies, as we may conclude from Eutocius, there must have been more Books to follow, because Proclus has preserved us details about higher curves, which must have come later. If again Geminus finished his work and wrote with the same fulness about the other branches of mathematics as he did about geometry, there must have been a considerable number of Books altogether. At all events it seems to have been designed to give a complete view of the whole science of mathematics, and in fact to be a sort of encyclopaedia of the subject. I shall now indicate first the certain, and secondly the probable, obligations of Proclus to Geminus, in which task 1 have only to follow van Pesch, who has embodied the results of Tittel's similar inquiry also*. I shall only omit the passages as regards which a case for attributing them to Geminus does not seem to me to have been made out. First come the following passages which must be attributed to Geminus, because Proclus mentions his name: (1) (In the first prologue of Proclus') on the division of mathe- 1 Pappus, ed, Hultsch, p. toi6, 9. ' Apollonius, fid. Heiberg, vol. II. p. 170. * Tannery, La Giomttru grttiiut, pp. 18, 19. * Proclus, pp. 38, I — 41, 8. * Heron, ed. Hultsch, pp. 140, 16—149, li - •Van Pesch, Di Prgck fonlibus, pp. 97—113. The dissertation of Tittel is entitled Dt Gtmini Steici ttudiit maihrmatuis ( 1895}- ' Proclus, pp, 38, I — 41, 8, except the allusion in p. 41, S — 10, to Ctesibius and Heron and 4 o INTRODUCTION [ch. iv matical sciences into arithmetic, geometry, mechanics, astronomy, optics, geodesy, canonic (science of musical harmony), and logistic (apparently arithmetical problems); (3) (in the note on the definition of a straight line) on the classification of lines (including curves) as simple (straight or circular) and mixed, composite and incomposite, uniform (ou-oiofiepels;) and non-uniform (avoitotoptpeis), lines " about solids " and lines produced by cutting solids, including conic and spiric sections' ; (3) (in the note on the definition of a plane surface) on similar distinctions extended to surfaces and solids' ; (4) (in the note on the definition of parallels) on lines which do not meet (d* itid. pp. 144, 14—1+6, 11. ' » itid. pp. 151, 5— 134, 10. " itid. pp. 301, 11—301, 13. " itid. pp. 311, 4—313. 3' 4» INTRODUCTION [ch. iv Apollonius' Conies, and the curves invented by Nicomedes, Hippias and Perseus 1 ; (16) a passage on the parallel-postulate regarded as the converse of I. i7«. Of the authors to whom P roc! us was indebted in a less degree the most important is Apollon ius of Perga. Two passages allude to his Conies*, one to a work on irrationals 1 , and two to a treatise On the cocklias (apparently the cylindrical helix) by Apollonius*. But more important for our purpose are six references to Apollonius in connexion with elementary geometry, (i) He appears as the author of an attempt to explain the idea of a line (possessing length but no breadth) by reference to daily experience, e.g. when we tell someone to measure, merely, the length of a road or of a wall'; and doubtless the similar passage showing how we may in like manner get a notion of a surface (without depth) is his also'. (2) He gave a new general definition of an angle*. (3) He tried to prove certain axioms', and Proclus gives his attempt to prove Axiom I, word for word". Proclus further quotes : (4) Apollonius' solution of the problem in Eucl. I. 10, avoiding Euclid's use of r, 9", (5) his solution of the problem in I. 11, differing only slightly from Euclid's", and (6) his solution of the problem in 1. 23 1 *, Heiberg" conjectures that Apollonius departed from Euclid's method in these propositions because he objected to solving problems of a more general, by means of problems of a more particular, character. Proclus however considers all three solutions inferior to Euclid's ; and his remarks on Apollonius' handling of these ele- mentary matters generally suggest that he was nettled by criticisms of Euclid in the work containing the things which he quotes from Apollonius, just as we conclude that Pappus was offended by the remarks of Apollonius about Euclid's incomplete treatment of the " three- and four-line locus 1 *." If this was the case, Proclus can hardly have got his information about these things at second-hand ; and there seems to be no reason to doubt that he had the actual work of Apollonius before him. This work may have been the treatise mentioned by Marinus in the words "Apollonius in his general treatise" CAirokXaivios iv r$ KaBoKov irpar/fitLTeiq) 1 '. If the notice in the Fihrist" stating, on the authority of T ha bit b. Qurra, that I Prdclus, pp. 355, 10—356, 16. ' ibid. p. 364, 9—11 ; pp. 36+, 10—365, 4. * Hid. p. 71, ig; p. 356, 8, 6. * ibid, p. 74, 13, 14. * ibid, pp. 105, s, 6, 14, 15. s Hid. p. too, 5—19. ' ibid. p. 1 [4, 10 — 13. * ibid. p. 113, I J— 19 (cf. p. 114, 17. p. t»5, ij). * ibid. p. 183, 13, 14. 10 ibid. pp. 194, 13 — J95, 5. II ibid. pp. »79, 16 — 180, 4. " ibid. p. 181, 8 — 19. 18 ibid. pp. 335, 16—336, 5. » PkiMegas, vol. xliji. p. 489. 15 See above, pp. 1, 3. '* Marinus in Euclidis Data, ed. Menge, p. 134, 16. 17 FiArist, U. Suter, p. 19. ch. iv] PROCLUS AND HIS SOURCES 43 Apollonius wrote a tract on the parallel -postulate be correct, it may have been included in the same work. We may conclude generally that, in it, Apollonius tried to remodel the beginnings of geometry, reducing the number of axioms, appealing, in his definitions of lines, surfaces etc., more to experience than to abstract reason, and substituting for certain proofs others of a more general character. The probabilities are that, in quoting from the tract of Ptolemy in which he tried to prove the para I lei -postulate, Proclus had the actual work before him. For, after an allusion to it as "a certain book 1 " he gives two long extracts', and at the beginning of the second indicates the title of the tract, "in the (book) about the meeting of straight lines produced from (angles) less than two right angles," as he has very rarely done in other cases. Certain things from Posidonius are evidently quoted at second- hand, the authority being Geminus (e.g. the definitions of figure and parallels) ; but besides these we have quotations from a separate work which he wrote to controvert Zeno of Sidon, an Epicurean who had sought to destroy the whole of geometry*. We are told that Ze.10 had argued that, even if we admit the fundamental principles (tipxavT€lvY. It is not necessary to suppose that Proclus had the original work of Zeno before him, because Zeno's arguments may easily have been got from Posidonius' reply ; but he would appear to have quoted direct from the latter at all events. The work of Carpus tnec/tankus (a treatise on astronomy) quoted from by Proclus" must have been accessible to him at first-hand, because a portion of the extract from it about the relation of theorems and problems" is reproduced word for word. Moreover, if he were not using the book itself, Proclus would hardly be in a position to question whether the introduction of the subject of theorems and problems ' Proclus, p. 191, 13. ' ibid. pp. 363, 14— 363, IB; pp. 36s, 7—3671 *?■ ' ibid, p. loo, 1—3. * ibid. pp. 199, u— 200, 1. * ibid. pp. »i4, [8 — iif, 13 [ pp. 516, 10 — 118, M. 4 ibid. p. 3l6, 11. 7 ibid. p. ?]8, I. * ibid, pp. J41, 19 — »43, 11. ' Md, pp. 141, n — 143, 11. 44 INTRODUCTION [ch. iv was opportune in the place where it was I'ound («* piv Kara Kaipiv rj ^»i, Trapela0a> wpos to irapov) 1 . It is of course evident that Proclus had before him the original works of Piato, Aristotle, Archimedes and Plotinus, as well as the Ivfifiutrd of Porphyry and the works of his master Syrianus (d ^ftirepo^ Ka8fjyefta)v) % , from whom he quotes in his note on the definition of an angle. Tannery also points out that he must have had before him a group of works representing the Pythagorean tradition on its mystic, as distinct from its mathematical, side, from Philolaus downwards, and comprising the more or less apocryphal Upfc Xoyai of Pythagoras, the Oracles (\6yta), and Orphic verses'. Besides quotations from writers whom we can identify with more or less certainty, there are many other passages which are doubtless quoted from other commentators whose names we do not know. A list of such passages is given by van Pesch*, and there is no need to cite them here. Van Pesch also gives at the end of his work" a convenient list of the books which, as the result of his investigation, he deems to have been accessible to and directly used by Proclus, The list is worth giving here, on the same ground of convenience. It is as follows: Eudemus : history of geometry. Gem in us : the theory of the mathematical sciences. Heron : commentary on the Elements of Euclid. Porphyry: „ „ Pappus t „ „ „ Apollonius of Perga : a work relating to elementary geometry. Ptolemy : on the parallel-postulate. Posidonius : a book controverting Zeno of Si don. Carpus : astronomy. Syrianus j a discussion on the angle. Pythagorean philosophical tradition. Plato's works. Aristotle's works. Archimedes' works, Plotinus : Enneaaes. Lastly we come to the question what passages, if any, in the commentary of Proclus represent his own contributions to the subject As we have seen, the onus probandi must be held to rest upon him who shall maintain that a particular note is original on the part of Proclus, Hence it is not enough that it should be impossible to point to another writer as the probable source of a note ; we must have a positive reason for attributing it to Proclus, The criterion must there- fore be found either (i) in the general terms in which Proclus points out the deficiencies in previous commentaries and indicates the respects in which his own will differ from them, or (2) in specific expressions used by him in introducing particular notes which may 1 Proclus, p. 741, j 1, 11. ■ ibid. p. 113, 19, 8 Tannery, La GionUtrie grecque, pp. 15, 16, * Van Pesch, Dt Prvcli fetttitrus, p. 135. * ibid. p. 155, CH. iv] PROCLUS AND HIS SOURCES 45 indicate that he is giving his own views. Besides indicating that he paid more attention than his predecessors to questions requiring deeper study (to irpayuarettSSes) and " pursued clear distinctions ' {to evStaiperav (ieTa8ia>Kovra<:) x — by which he appears to imply that his predecessors had confused the different departments of their commentaries, viz. lemmas, cases, and objections (eyo-rrwret?)* — Proclus complains that the earlier commentators had failed to indicate the ultimate grounds or muses of propositions*. Although it is from Geminus that he borrowed a passage maintaining that it is one of the proper functions of geometry to inquire into causes (tijii alrtav koX to S*a Tt)*, yet it is not likely that Geminus dealt with Euclid's propositions one by one ; and consequently, when we find Proclus, on I. 8, 16, 17, 18, 32, and 47', endeavouring to explain causes, we have good reason to suppose that the explanations are his own. Again, his remarks on certain things which he quotes from Pappus can scarcely be due to anyone else, since Pappus is the latest of the commentators whose works he appears to have used. Under this head, come (1) his objections to certain new axioms introduced by Pappus*, (2) his conjecture as to how Pappus came to think of his alterna- tive proof of I. 5', (3) an addition to Pappus' remarks about the curvilineal angle which is equal to a right angle without being one*. The defence of Geminus against Carpus, who combated his view of theorems and problems, is also probably due to Proclus', as well as an observation on t. 38 to the effect that I. 35 — 38 are really compre- hended in VI. 1 as particular cases". Lastly, we can have no hesitation in attributing to Proclus himself (1) the criticism of Ptolemy's attempt to prove the parallel-postulate", and (2) the other attempted proof given in the same note" (on I. 29) and assuming as an axiom that " if from one point two straight lines forming an angle be produced ad infinitum the distance between them when so produced ad infinitum exceeds any finite magnitude (i.e. length)," an assumption which purports to be the equivalent of a statement in Aristotle 1 *. It is introduced by words in which the writer appears to claim originality for his proof: "To him who desires to see this proved (Ka-Ta&Kevatyfieiiov) let it be said by us (Xeyia-ff& trap thiwv)" etc." Moreover, Philoponus, in a note on Aristotle's Anal. pest. 1. to, says that " the geometer (Euclid) assumes this as an axiom, but it wants a great deal of proof, insomuch that both Ptolemy and Proclus wrote a whole book upon it"." 1 Proclui, p. 84, 1 j, p. 431, ff, j j. ■ Cf. ibid, p. *8g, u— ij ; p. 431, 13— ij. ibid. p. 431, 17. * ibid, p. 103, y—^$. Proclus, p. II pp. 316, T+- _ Proclus, p. 198, 5 — 15. 1 ibid, p. 150, 11—19. * ibid. p. 190, 9 — ^13. * See Proclus, p. 170, 3—14 (1. 8); pp. 309, 3—310, 8 (1. 16); pp. 310, 19—311, 33 •T)( PP' 3'<>. '+—318, » ('• '8); p. 384, 13— j 1 (1. 31) s pp. +16, 11— 417. 8 ft. 47). * ibid, p. 343, 11 — 19. * ibid. pp. 403, 6—406, 9. 11 ibid. p. 368, 1— 13. u ibid. pp. 371, 11—373, »■ u Aristotle, de caila, 1. 3 (171 b 18 — 30), " Produs, p. 371, 10. u Berlin Aristotle, voL IV. p. 114 a 9 — 11. CHAPTER V. THE TEXT 1 . It is well known that the title of Simson's edition of Euclid (first brought out in 'Latin and English in 1756) claims that, in it, "the errors by which Theon, or others, have long ago vitiated these books are corrected, and some of Euclid's demonstrations are restored " ; and readers of Simson's notes are familiar with the phrases used, where anything in the text does not seem to him satisfactory, to the effect that the demonstration has been spoiled, or things have been interpo- lated or omitted, by Theon "or some other unskilful editor." Now most of the MSS. of the Greek text prove by their titles that they proceed from the recension of the Elements by Theon ; they purport to be either " from the edition of Theon "(«tij? ©eWoe e'wSoVew?) or " from the lectures of Theon " (Avo avvovaw&v tov QsWoc). This was Theon of Alexandria {4th c. A.D.) who also wrote a commentary on Ptolemy, in which there occurs a passage of the greatest importance in this connexion*: "But that sectors in equal circles are to one another as the angles on which they stand has been proved by me in my edition of the Elements at the end of the sixth book." Thus Theon himself says that he edited the Elements and also that the second part of VI. 33, found in nearly all the MSS., is his addition. This passage is the key to the whole question of Th eon's changes in the text of Euclid ; for, when Peyrard found in the Vatican the MS. 190 which contained neither the words from the titles of the other MSS. quoted above nor the interpolated second part of VI. 33, he was justified in concluding, as he did, that in the Vatican MS. we have an edition more ancient than Theon 's. It is also clear that the copyist of P, or rather of its archetype, had before him the two recensions and systematically gave the preference to the earlier one ; for at xiii. 6 in P the first hand has added a note in the margin : " This theorem is not given in most copies of the new edition, but is found in those of the old." Thus we are more fortunate than Simson, since our judgment of Theon's recension can be formed on the basis, not of mere conjecture, but of the documentary evidence afforded by a comparison of the Vatican MS. just mentioned with what we may conveniently call, after Heiberg, the Theonine MSS. 1 The material for the whole of this chapter is taken from Hei berg's edition of the Elements, introduction to vol. v., and from the same scholar's Littcrargcscktihlltihc Siudien iibcr Euilid, p. tJ4sqq. and ParaHpomena zit EuMidm Her-mcs, XXXVltt., too}. 9 1. p. J 01 ed. Ililmj = |i. 50 ed. Basel. ch. v] THE TEXT 47 The MSS. used for Hei berg's edition of the Elements are the following : (i) P = Vatican MS. numbered 190, 4to, in two volumes (doubt- less one originally) ; 10th c. This is the MS. which Peyrard was able to use ; it was sent from Rome to Paris for his use and bears the stamp of the Paris Imperial Library on the last page. It is well and carefully written. There are corrections some of which are by the original hand, but generally in paler ink, others, still pretty old, by several different hands, or by one hand with different ink in different places (P m. 2), and others again by the latest hand (P m. rec). It contains, first, the Elements I. — xin. with scholia, then Marinus' commentary on the Data (without the name of the author), followed by the Data itself and scholia, then the Elements XIV., XV. (so called), and lastly three books and a part of a fourth of a commentary by Theon eis rot)* irpoxeipow Kavovat IlToXe- fiaiou. The other MSS. are " Theon ine." (2) F = MS. xxvili, 3, in the Laurentian Library at Florence, 4to; 10th c. This MS. is written in a beautiful and scholarly hand and contains the Elements I. — XV., the Optics and the Phaenotnena, but is not well preserved. Not only is the original writing renewed in many places, where it had become faint, by a later hand of the 1 6th c, but the same hand has filled certain smaller lacunae by gumming on to torn pages new pieces of parchment, and has replaced bodily certain portions of the MS., which had doubtless become illegible, by fresh leaves. The larger gaps so made good extend from Eucl. Vlt. 1 2 to IX. 1 5, and from xil. 3 to the end ; so that, besides the conclusion of the Elements, the Optics and Phaenomena are also in the later hand, and we cannot even tell what in addition to the Elements 1. — xin. the original MS. contained. Hei berg denotes the later hand by and observes that, while in restoring wo;ds which had become faint and filling up minor lacunae the writer used no other MS., yet in the two larger restorations he used the Laurentian MS. XXVIII, 6, belonging to the 13th — 14th c. The latter MS. (which Heiberg denotes by f) was copied from the Viennese MS. (V) to be described below. (3) B m Bodleian MS., D'Orville X. 1 inf. 2, 30, 4to ; A.D. 888. This MS. contains the Elements I. — XV. with many scholia. Leaves 15 — 118 contain 1. 14 (from about the middle of the proposition) to the end of Book VI., and leaves 123 — 387 (wrongly numbered 397) Books VII, — XV, in one and the same elegant hand (9th c). The leaves preceding leaf 15 seem to have been lost at some time, leaves 6 to 14 (containing Elem. L to the place in I. 14 above referred to) being carelessly written by a later hand on thick and common parch- ment (13th c). On leaves 2 to 4 and 122 are certain notes in the hand of Arethas, who also wrote a two-line epigram on leaf 5, the greater part of the scholia in uncial letters, a few notes and corrections, and two sentences on the last leaf, the first of which states that the MS. was written by one Stephen cleric us in the year of the world 6397 48 INTRODUCTION [ch. v (= 888 a.D,), while the second records Arethas' own acquisition of it. Arethas lived from, say, 865 to 939 a.D. He was Archbishop of Caesarea and wrote a commentary on the Apocalypse. The portions of his library which survive are of the greatest interest to palaeography on account of his exact notes of dates, names of copyists, prices of parchment etc. It is to him also that we owe the famous Plato MS. from Fatmos (Cod. Clarkianus) which was written for him in November 895 ! . (4) V - Viennese MS. Philos. Gr. No. 103 ; probably 12th c. This MS. contains 292 leaves, Eucl, Elements I. — XV. occupying leaves 1 to 254, after which come the Optics (to leaf 271), the P%aenomena (mutilated at the end) from leaf 372 to leaf 282, and lastly scholia, on leaves 283 to 292, also imperfect at the end. The different material used for different parts and the varieties of handwriting make it necessary for Heiberg to discuss this ms. at some length'. The handwriting on leaves 1 to 183 (Book 1. to the middle of X. 105) and on leaves 203 to 234 (from XI 31, towards the end of the proposition, to XIII. 7, a few lines down) is the same; between leaves 184 and 202 there are two varieties of handwriting, that of leaves 184 to 189 and that of leaves 200 (verso) to 202 being the same. Leaf 235 begins in the same handwriting, changes first gradually into that of leaves 184 to 189 and then (verso) into a third more rapid cursive writing which is the same as that of the greater part of the scholia, and also as that of leaves 243 and 282, although, as these leaves are of different material, the look of the writing and of the ink seems altered. There are corrections both by the first and a second hand, and scholia by many hands. On the whole, in spite of the apparent diversity of handwriting in the MS., it is probable that the whole of it was written at about the same time, and it may (allowing for changes of material, ink etc) even have been written by the same man. It is at least certain that, when the Laurentian ms. xxviii, 6 was copied from it, the whole MS, was in the condition in which it is now, except as regards the later scholia and leaves 283 to 292 which are not in the Laurentian MS., that MS. coming to an end where the Phaenomena breaks off abruptly in V, Hence Heiberg attributes the whole MS. to the izthc. But it was apparently in two volumes originally, the first con- sisting of leaves 1 to 183 ; and it is certain that it was not all copied at the same time or from one and the same original. For leaves 1 84 to 202 were evidently copied from two MSS, different both from one another and from that from which the rest was copied. Leaves 184 to the middle of leaf 189 (recto) must have been copied from a MS. similar to P, as is proved by similarity of readings, though not from F itself. The rest, up to leaf 202, were copied from the Bologna MS. (b) to be mentioned below. It seems clear that the content of leaves 1 84 to 202 was supplied from other MSS. because there was a lacuna in the original from which the rest of V was copied. 1 See Pauty-Wissowa, Ktal- Encyclopedic dtr class. AUertitiHSWitunickaft, vol. [I., 1896, dberg, vol. v. pp. xxts — xxxiii. * **fc ch. v] THE TEXT 49 Heiberg sums up his conclusions thus. The copyist of V first copied leaves i to 183 from an original in which two quaterntones were missing (covering from the middle of Eucl. X. 105 to near the end of XI. 31). Noticing the lacuna he put aside one quatemio of the parchment used up to that point. Then he copied onwards from the end of the lacuna in the original to the end of the Pkaenomena. After this he looked about him for another ms. from which to fill up the lacuna; finding one, he copied from it as far as the middle of leaf 189 (recto). Then, noticing that the MS. from which he was copying was of a different class, he had recourse to yet another MS. from which he copied up to leaf 302, At the same time, finding that the lacuna was longer than he had reckoned for, he had to use twelve more leaves of a different parchment in addition to the quatemio which he had put aside. The whole MS. at first formed two volumes (the first containing leaves 1 to 1 83 and the second leaves 1 84 to 282) ; then, after the last leaf had perished, the two volumes were made into one to which two more quaterniones were also added. A few leaves of the latter of these two have since perished, (5) b = MS. numbered 18 — 19 in the Communal Library at Bologna, in two volumes, 4to ; nth c. This MS. has scholia in the margin written both by the first hand and by two or three later hands ; some are written by the latest hand, Theodorus Cabasilas (a descendant apparently of Nicolaus Cabasilas, 14th c.) who owned the MS. at one time. It contains (a) in 14 quater- niones the definitions and the enunciations (without proofs) of the Elements I. — XIII, and of the Data, {b) in the remainder of the volumes the Proem to Geometry (published among the Variae Collectiones in Hultsch's edition of Heron, pp. 252, 24 to 274, 14) followed by the Elements I. — XIII. (part of XIII. 18 to the end being missing), and then by part of the Data (from the last three words of the enunciation of Prop. 38 to the end of the penultimate clause in Prop. 87, ed. Menge). From xi. 36 inclusive to the end of xn. this MS. appears to represent an entirely different recension. Heiberg is compelled to give this portion of b separately in an appendix. He conjectures that it is due to a Byzantine mathematician who thought Euclid's proofs too long and tiresome and consequently contented himself with indicating the course followed 1 . At the same time this Byzantine must have had an excellent MS. before him, probably of the ante-Theonine variety of which the Vatican Ms. 190 (P) is the sole representative. (6) p = Paris MS. 2466, 41.0; 12th c This manuscript is written in two hands, the finer hand occupying leaves 1 to 53 (recto), and a more careless hand leaves 53 (verso) to 64, which are of the same parchment as the earlier leaves, and leaves 65 to 239, which are of a thinner and rougher parchment showing traces of writing of the 8th — 9th c (a Greek version of the Old Testament). The ms, contains the Elements 1. — xm. and some scholia after Books XL, XII. and xm. 1 ZtUs)v toD kvk\ov V€ptif>epeiav also found in practically all the MSS. Thus Heiberg's assumption that both expressions are interpolations is now conSrmed by this oldest of all sources. 2. The Oxyrhynchus Papyri \. p. 58, No. xxix. of the 3rd or 4U1 c. This fragment contains the enunciation of Eucl. II. 5 (with figure, apparently without letters, immediately following, and not, as usual in our MSS., at the end of the proof) and before it the part of a word irept#xpp£ belonging to II. 4 (with room for -v

' cmep ehu 1 [dt r]A rod EtixXtloav rrotxtia Tpo\at*fia*6tti*& i* w6r Hp&cXou wopdiijv *al rar' /rt- rdp^r. Cf. p. 33, note 8, Above. ■ Heiberg, Paralipemtna ;u liuklid in Htrma, XXXV MI., 190 J, pp. 46 — 74, ifil — aoi, 311—356. 1 Described by Heiberg in Overrigt ever dtt kngi, dansks Vtdtnskabtrittt Sthkabs Forhandiingtr, rooo, p. 161. em v] THE TEXT 51 Sei|o( and a stroke to mark the end), showing that the fragment had not the Porism which appears in all the Theonine MSS. and (in a later hand) in P, and thereby confirming Hetberg's assumption that the Porism was due to Theon. 3, A fragment in Fayum towns and their papyri, p. 96, No. IX. of 2nd or 3rd c. This contains I. 39 and I. 41 following one another and almost complete, showing that I. 40 was wanting, whereas it is found in all the MSS. and is recognised by Prod us. Moreover the text of the beginning of I. 39 is better than ours, since it has no double Biopwfios but omits the first (" I say that they are also in the same parallels ") and has " and" instead of "for let AD be joined " in the next sentence. It is clear that I. 40 was interpolated by someone who thought there ought to be a proposition following I. 39 and related to it as I. 38 is related to I. 37 and I. 36 to I. 35, although Euclid nowhere uses 1. 40, and therefore was not likely to include it The same interpolator failed to realise that the words " let AD be joined" were part of the eK^ca-is or setting-out, and took them for the *oTocr*einj or " construc- tion " which generally follows the 8(op«r^>? or " particular statement " of the conclusion to be proved, and consequently thought it necessary to insert a &topurfi,6 for fl ( V for Z in my figure) because he saw that the perpendicular from K to B4> would fall on <£ itself, so that , tl coincide. But, if the substitution is made, it should be proved that 4>, £1 coincide. Euclid can hardly have failed to notice the fact, but it may be that he deliberately ignored it as unnecessary for his purpose, because he did not .want to lengthen his proposition by giving the proof. I I. Emendations intended to improve the form or diction of Euclid. Some of these emendations of Theon affect passages of appreciable length. Heiberg notes about ten such passages ; the longest is in Eucl. xif. 4 where a whole page of Heiberg's text is affected and Theon's version is put in the Appendix. The kind of alteration may be illustrated by that in ix. r 5 where Euclid uses successively the propositions VII, 24, 25, quoting the enunciation of the former but not of the latter ; Theon does exactly the reverse. In a few of the cases here quoted by Heiberg, Theon shortened the original somewhat. But, as a rule, the emendations affect only a few words in each sentence. Sometimes they are considerable enough to alter the con- formation of the sentence, sometimes they are trifling alterations "more magistellorum ineptorum" and unworthy of Theon. Generally speaking, they were prompted by a desire to change anything which was out of the common in expression or in form, in order to reduce the language to one and the same standard or norm. Thus Theon changed the order of words, substituted one word for another where the latter was used in a sense unusual with Euclid (e.g. i-ireihijirep, " since," for on. in the sense of " because "), or one expression for another in like circumstances (e.g. where, finding "that which was enjoined would be done" in a theorem, VII. 31, and deeming the phrase more appropriate to a problem, he substituted for it " that which is sought would be manifest"; probably also and for similar reasons he made certain variations between the two expressions usual at the end of propositions otrep eSei Setl-ai and oirep (Bet iroifjerat, quod erat demonstrandum and quod erat faciendum). Sometimes his alterations show carelessness in the use of technical terms, as when he uses aTTTtaSai (to meet) for t4>d-n-T£. at the end of propositions. This is often the case at the end of porisms, where, in omitting the words, Theon seems to have deliberately departed from Euclid's practice. The MS. P seems to show clearly that, where Euclid put a porism at the end of a proposition, he omitted the Q.E.D. at the end of the proposition but inserted it at the end of the porism, as if he regarded the latter as being actually a part of the proposition itself. As in the Theonine MSS. the Q.ED, is generally omitted, the omission would seem to have been due to Theon. Sometimes in these cases the Q.E.D. is interpolated at the end of the proposition. Heiberg summed up the discussion of Theon 's edition by the remark that Theon evidently took no pains to discover and restore from MSS. the actual words which Euclid had written, but aimed much more at removing difficulties that might be feit by learners in studying the book. His edition is therefore not to be compared with the editions of the Alexandrine grammarians, but rather with the work done by Eutocius in editing 1 Apollonius and with an interpolated recension of some of the works of Archimedes by a certain Byzantine, Theon occupying a position midway between these two editors, being superior to the latter in mathematical knowledge but behind Eutocius in industry (these views now require to be some- what modified, as above stated). But however little Theon's object may be approved by those of us who would rather know the ipsissima verba of Euclid, there is no doubt that his work was approved by his pupils at Alexandria for whom it was written ; and his edition was almost exclusively used by later Greeks, with the result that the more ancient text is only preserved to us in one MS. As the result of the above investigation, we may feel satisfied that, where P and the Theonine MSS. agree, they give us (except in a few accidental instances) Euclid as he was read by the Greeks of the 4th c. But even at that time the text had been passed from hand to hand through more than six centuries, so that it is certain that it had already suffered changes, due partly to the fault of copyists and partly to the interpolations of mathematicians. Some errors of copyists escaped Theon and were corrected in some MSS. by later hands. Others appear in all our MSS. and, as they cannot have arisen accidentally in all, we must put them down to a common source more ancient than Theon. A somewhat serious instance is to be found in III. % ; and the use of aTrrka&w for eaTrr4 in the sense of " touch " may also be mentioned, the proper distinction between the words having been ignored as it was by Theon also. But there are a number of imperfections in the ante-Theonine text which it would be unsafe to put down to the errors of copyists, those namely where the good MSS. agree and it is not possible to see any motive that a copyist could have had for altering a correct reading. In these cases it is possible that the imperfections are due to a certain degree of carelessness on the part of Euclid himself; for it 58 INTRODUCTION [ch. v is not possible " Euclidem ab omni naevo vindicare," to use the words of Saccheri", and consequently Slmson is not right in attributing to Theon and other editors all the things in Euclid to which mathe- matical objection can be taken. Thus, when Euclid speaks of " the ratio compounded of the sides" for "the ratio compounded of the ratios of the sides," there is no reason for doubting that Euclid himself is responsible for the more slip-shod expression. Again, in the Books XI. — Xin. relating to solid geometry there are blots neither few nor altogether unimportant which can only be attributed to Euclid himself; and there is the less reason for hesitation in so attributing them because solid geometry was then being treated in a thoroughly systematic manner for the first time. Sometimes the conclusion (ovfvn-ipaa-pa) of a proposition does not correspond exactly to the enunciation, often it is cut short with the words teal t« i^rjf "and the rest" (especially From Book X. onwards), and very often in Books viil., IX. it is omitted. Where all the MSS. agree, there is no ground for hesitating to attribute the abbreviation or omission to Euclid; though, of course, where one or more mss. have the longer form, it must be retained because this is one of the cases where a copyist has a temptation to abbreviate. Where the true reading is preserved in one of the Theonine MSS. alone, Heiberg attributes the wrong reading to a mistake which arose before Theon's time, and the right reading of the single MS. to a successful correction. We now come to the most important question of the Interpolations introduced before Theon's time. I. Alternative proofs or additional cases. It is not in itself probable that Euclid would have given two proofs of the same proposition ; and the doubt as to the genuineness of the alternatives is increased when we consider the character of some of them and the way in which they are introduced. First of all, we have those of VI. 20 and XII. 1 7 introduced by " we shall prove this otherwise more readily (irpox^pOTepov)" or that of X. 90 " it is possible to prove more shortly (trvvTopcoTepov)." Now it is impossible to suppose that Euclid would have given one proof as that definitely accepted by him and then added another with the express comment that the latter has certain advantages over the former. Had he con- sidered the two proofs and come to this conclusion, he would have inserted the latter in the received text instead of the former. These alternative proofs must therefore have been interpolated. The same argument applies to alternatives introduced with the words "or even thiis " (tj Kal ovtbw), " or even otherwise " (1} teal a\ka>ftmv) is either placed round or taken away"; 11. No, 13, also on the gnomon; IV. No. 2 stating that Book IV. was the discovery of the Pythagoreans ; V. No. 1 attributing the content of Book v. to Eudoxus ; x. No. 1 with its allusion to the discovery of incommensurability by the Pytha- goreans and to Apollonius' work on irrationals; x. No. 62 definitely attributing X. 9 to Theaetetus; XIII. No. I about the "Platonic" figures, which attributes the cube, the pyramid, and the dodecahedron to the Pythagoreans, and the octahedron and icosahedron to Theaetetus. Sometimes the scholia are useful in connexion with the settlement of the text, (1) directly, e.g. III. No. 16 on the interpolation of the word "within" (eWot) in the enunciation of 111, 6, and x. No. 1 alluding to the discussion by "Theon and some others" of irrational "surfaces" and "solids," as well as "lines," from which we may 1 Heiberg, Om Schelitrne til Eutlids Elemenitr, Kjebenhavn, 1888. The tract is written in Danish, but, fortunately for those who do not read Danish easily, the author has appended (pp. 70 — 78) a resume in French. ch. vi] THE SCHOLIA 65 conclude that the scholium at the end of Book x, is not genuine ; (2) indirectly in that they sometimes throw light on the connexion of certain MSS. Lastly, they have their historical importance as enabling us to judge of the state of mathematical science at the times when they were written. Before passing to the classification of the scholia, Heiberg remarks that we must separate from them a number of additions in the nature of scholia which are found in the text of our MSS. but which can, in one way or another, be proved to be spurious. As they are found both in P and in the Theonine MSS., they must have been in the MSS. anterior to Theon (4th c). But they are, in great part, only found in the margin of P and the Theonine MSS. ; in V they are half in the text and half in the margin. This can hardly be explained except on the supposition that these additions were originally (in the MSS. before Theory) in the margin, and that Theon kept them there in his edition, but that they afterwards found their way gradually into the text of P as well as of the Theonine MSS., or were omitted altogether, while particular MSS. have in certain places preserved the old arrange- ment Of such spurious additions Heiberg enumerates the following: the axiom about equals subtracted from unequals, the last lines of the porism to vi. 8, second porisms to v. 19 and to vi. 20, the porism to hi. 31, vi. Def. 5, various additions in Book X., the analyses and syntheses of XI 1 1. 1 — Si an ^ the proposition XIII. 6. The two first classes of scholia distinguished by Heiberg are denoted by the convenient abbreviations "Schol. Vat." and "Schol. Vind." I. Schol, Vat. It is first necessary^ to set out the letters by which Heiberg denotes certain collections of scholia. P = Scholia in P written by the first hand. B = Scholia in B by a hand of the same date as the MS. itself, generally that of Aretha s. F = Scholia in F by the first hand. Vat = Scholia of the Vatican MS. 204 of the 10th c, which has these scholia on leaves 198 — 205 (the end is missing) as an independent collection. It does not contain the text of the Elements. V c = Scholia found on leaves 283 — 292 of V and written in the same hand as that part of the MS. itself which begins at leaf 233. Vat 192 = a Vatican MS. of the 14th c. which contains, after (l) the Elements I. — XIII. (without scholia), (2) the Data with scholia, (3) Marinus on the Data, the Schol. Vat as an independent collection and in their entirety, beginning with 1. No. 88 and ending with xm. No. 44. The Schol. Vat., the most ancient and important collection of scholia, comprise those which are found in PBF Vat. and, from VII, 12 to IX. 15, in PB Vat. only, since in that portion of the Elements F was restored by a later hand without scholia; they also include 1, 66 INTRODUCTION [ch. vi No. 88 which only happens to be erased in F, and IX. Nos. 28, 29 which may be left out because F, here has a different text In F and Vat. the collection ends with Book x. ; but it must also include Schol. FB of Books xi. — xill., since these are found along with Schol. Vat. to Books I. — X. in several MSS. (of which Vat. 192 is one) as a separate collection. The Schol. Vat. to Books X. — XIII. are also found in the collection V c (where, curiously enough, xill. Nos, 43, 44 are at the beginning). The Schol. Vat. accordingly include Schol. PBV C Vat, 192, and doubtless also those which are found in two of these sources. The total number of scholia classified by Heiberg as Schol. Vat. is 138. As regards the contents of Schol. Vat. Heiberg has the following observations. The thirteen scholia to Book I. are extracts made from Proclus by a writer thoroughly conversant with the subject, and cleverly recast (with some additions). Their author does not seem to have had the two lacunae which our text of Proclus has (at the end of the note on 1. 36 and the beginning of the next note, and at the beginning of the note on I. 43), for the scholia I. Nos. 125 and 137 seem to fill the gaps appropriately, at least in part. In some passages he had better readings than our MSS. have. The rest of Schol. Vat. (on Books II. — xill.) are essentially of the same character as those on Book 1., containing prolegomena, remarks on the object of the propositions, critical remarks on the text, converses, lemmas ; they are, in general, exact and true to tradition. The reason of the resemblance between them and Proclus appears to be due to the fact that they have their origin in the commentary of Pappus, of which we know that Proclus also made use. In support of the view that Pappus is the source, Heiberg places some of the Schol, Vat. to Book X. side by side with passages from the com- mentary of Pappus in the Arabic translation discovered by Woepcke 1 ; he also refers to the striking confirmation afforded by the fact that XII, No. 2 contains the solution of the problem of inscribing in a given circle a polygon similar to a polygon inscribed in another circle, which problem Eutocius says' that Pappus gave in his commentary on the Elements. But, on the other hand, Schol. Vat. contain some things which cannot have come from Pappus, e.g. the allusion in X. No. 1 to Theon and irrational surfaces and solids, Theon being later than Pappus ; in. No. 10 about porisms is more like Proclus' treatment of the subject than Pappus', though one expression recalls that of Pappus about forming (c^Tj/MtTifeo-ftit) the enunciations of porisms like those of either theorems or problems. The Schol, Vat. give us important indications as regards the text of the EUmtntt as Pappus had it. In particular, they show that he could not have had in his text certain of the lemmas in Book X. For example, three of these are identical with what we find in Schol, 1 Om Scholitnu til Euklids EUmtnltr, pp. 11, it: cf. Euilid-StHtiien, pp. 170, 171; Woepcke, MJmoira prtstnt. & ?Ac«d. dts Seitnnt, 18 56, XIV. p. 6j8sqq. * Archimedes, ed. Heiberg, in. p. a8, 19—15. CM. VI] THE SCHOLIA 67 Vat (the lemma to X. 17 = Schol. X. No. 106, and the lemmas to X. 54, 60 come in Schol. X. No. 328) ; and it is not possible to suppose. that these lemmas, if they were already in the text, would also be given as scholia. Of these three lemmas, that before X. 60 has already been condemned for other reasons ; the other two, un- objectionable in themselves, must be rejected on the ground now stated. There were four others against which Heiberg found nothing to urge when writing his prolegomena to Vol. v., viz. the lemmas before X. 42, X. 14, X. 2Z and X. 33. Of these, the lemma to X. 22 is not reconcilable with Schol, x. No. 161, which takes up the assumption in the text of Eucl. X. 22 as if no lemma had gone before. The lemma to X. 42, which, on account of the words introducing it (see p. 60 above), Heiberg at first hesitated to regard as an inter- polation, is identical with Schol. X. No. 27a It is true that in Schol. x. No. 269 we find the words "this lemma has been proved before (£p tow ipTrpoaStv), but it shall also be proved now for convenience' sake (rov eroinov evexa,)" and it is possible to suppose that " before " may mean in Euclid's text before x. 42 ; but a proof in that place would surely have been as " convenient " as could be desired, and it is therefore more probable that the proof had been given by Pappus in some earlier place. (It may be added that the lemma to X. 14, which is identical with the lemma to XI. 23, con- demned on other grounds, is for that reason open to suspicion.) Heiberg's conclusion is that all the lemmas are spurious, and that most or alt of them have found their way into the text from Pappus' commentary, though at a time anterior to Theon's edition, since they are found in all our MSS. This enables us to fix a date for these interpolations, namely the first half of the 4th c. Of course Pappus had not in his text the interpolations which, from the fact of their appearing only in some of our MSS., are seen to be later than those above-mentioned. Such are the lemmas which are found in the text of V only after X. 29 and X. 31 respectively and are given in Heiberg's Appendix to Book X. (numbered 10 and 11). On the other hand it appears from Woepcke's tract 1 that Pappus already had x, 115 in his text : though it does not follow from this that the proposition is genuine but only that interpolations began very early. Theon interpolated a proposition (or lemma) between X. 12 and X. 13 (No, S in Heiberg's Appendix). Schol. Vat. has the same thing (X. No. 125). The writer of the scholia therefore did not find this lemma in the text. Schol. Vat IX. Nos. 28, 29 show that neither did he find in his text the alterations which Theon made in Eucl. IX. 19; the scholia in fact only agree with the text of P, not with Theon's. This suggests that Schol. Vat. were written for use with a MS. of the ante-Theonine recension such as F is. This probability is further confirmed by a certain independence which P shows in several places when compared with the Theon ine MSS. Not only has P better readings in some passages, but more substantial divergences; and, 1 Woepcke, op, rft. p. 702. 68 INTRODUCTION [eft vi in particular, the absence in P of three notes of a historical character which are added, wholly or partly from Prod us, in the Theonine MSS. attests an independent and more primitive point of view in P. In view of the distinctive character of P, it is possible that some of the scholia found in it in the first hand, but not in the other sources of Schol. Vat., also belong to that collection ; and several circumstances confirm this. Schol. XIII. No. 45, found in P only, which relates to a passage in Eucl. XIII. 13, shows that certain words in the text, though older than Theon, are interpolated ; and, as the scholium is itself older than Theon, is headed "third lemma," and follows a "second lemma" relating to a passage in the text im- mediately preceding, which "second lemma" belongs to Schol. Vat. and is taken from Pappus, the "third" in all probability came from Pappus also. The same is true of Schol. XII. No. 72 and xm. No. 69, which are respectively identical with the propositions vulgo XI. 38 (Heiberg, A pp. to Book xi., No, 3) and XII I. 6; for both of these interpolations are older than Theon. Moreover most of the scholia which P in the first hand alone has are of the same character as Schol, Vat Thus VII. No. 7 and XIII. No. I introducing Books VII. and xm. respectively are of the same historical character as several of Schol Vat ; that vil. No. 7 appears in the text of P at the beginning of Book VII. constitutes no difficulty. There are a number of converses, remarks on the relation of propositions to one another, explanations such as XII. No. 89 in which it is remarked that , fl in Euclid's figure to xil. 17 {Z, V in my figure) are really the same point but that this makes no difference in the proof. Two other Schol. P on XII. 17 are connected by their headings with XII. No. 72 mentioned above, xi. No. 10 (P) is only another form of xi. No. 1 1 (B) ; and B often, alone with P, has preserved Schol. Vat On the whole Heiberg considers some 40 scholia found in P alone to belong to Schol. Vat. The history of Schol. Vat. appears to have been, in its main outlines, the following. They were put together after 500 A.D., since they contain extracts from Proclus, to which we ought not to assign a date too near to that of Proclus' work itself; and they must at least be earlier than the latter half of the 9th c, in which B was written. As there must evidently have been several intermediate links between the archetype and B, we must assign them rather to the first half of the period between the two dates, and it is not improbable that they were a new product of the great development of mathematical studies at the end of the 6th c. (Isidorus of Miletus). The author extracted what he found of interest in the commentary of Proclus on Book I. and in that of Pappus on the rest of the work, and put these extracts in the margin of a MS. of the class of P. As there are no scholia to I. 1 — 22, the first leaves of the archetype or of one of the earliest copies must have been lost at an early date, and it was from that mutilated copy that partly P and partly a MS. of the Theonine class were taken, the scholia being put in the margin in both. Then the collection spread through the Theonine MSS., gradually losing some ch. vi] THE SCHOLIA 69 scholia which could not be read or understood, or which were accidentally or deliberately omitted. Next it was extracted from one of these MSS. and made into a separate work which has been preserved, in part, in its entirety (Vat. 192 etc.) and, in part, divided into sections, so that ihe scholia to Books X. — xni. were detached (V c ). It had the same fate in the mss, which kept the original arrangement (in the margin), and in consequence there are some MSS. where the scholia to the stereometric Books are missing, those Books having come to be less read in the period of decadence. It is from one of these MSS. that the collection was extracted as a separate work such as we find it in Vat. ( roth c). II. The second great division of the scholia is Schol. Vind. This title is taken from the Viennese MS. (V), and the letters used by Heiberg to indicate the sources here in question are as follows. V* = scholia in V written by the same hand that copied the MS. itself from fol. 235 onward. q = scholia of the Paris MS. 2344 (q) written by the first hand. 1 = scholia of the Florence ms. Laurent xxvin, 2 written in the 13th — 14th c, mostly in the first hand, but partly in two later hands. V b = scholia in V written by the same hand as the first part (leaves 1 — 1S3) of the MS. itself; V" wrote his scholia after V". q 1 = scholia of the Paris MS. (q) found here and there in another hand of early date. Schol. Vind. include scholia found in V m q. 1 is nearly related to q ; and in fact the three Mss. which, so far as Euclid's text is con- cerned, show no direct interdependence, are. as regards their scholia, derived from one original. Heiberg proves this by reference to the readings of the three in two passages (found in Schol. I, No. 109 and X. No, 39 respectively). The common source must have contained, besides the scholia found in the three MSS. V a ql, those also which are contained in two of them, for it is more unlikely that two of the three should contain common interpolations than that a particular scholium should drop out of one of them. Besides V" and q, the scholia V b and q 1 must equally be referred to Schol. Vind., since the greater part of their scholia are found in 1. There is a lacuna in q from Eucl. VIII. 25 to IX. 14, so that for this portion of the Elements Schol. Vind. are represented by VI only, Heiberg gives about 450 numbers in all as belonging to this collection. Schol. Vind. did not all come from one source; this is shown by differences of substance, e.g. between X. Nos. 36 and 39, and by differences of time of writing : e.g. vi. No. 52 refers at the beginning to No. 55 with the words "as the scholium has it" and is therefore later than that scholium ; X. No. 247 is also later than x. No. 246. The scholia to Book I. are here also extracts from Proclus, but more copious and more verbatim than in Schol. Vat. The author has not always understood Proclus; and he had a text as bad as that of our MSS., with the same lacunae. The scholia to the other 7 o INTRODUCTION [ch. VI Books are partly drawn (i) from Schol. Vat., the MSS. representing Schol. Vind, and Schol. Vat. in these cases showing nearly all possible combinations; but there is no certain trace in Schol. Vind. of the scholia peculiar to P. The author used a copy of Schol. Vat. in the form in which they were attached to the Theonine text ; thus Schol. Vind. correspond to BF Vat., where these diverge from P, and especially closely to B. Besides Schol. Vat., the editors of Schol. Vind. used {2) other old collections 0/ scholia of which we find traces in B and F; Schol. Vind. have also some scholia common with b. The scholia which Schoi. Vind. have in common with BF come from two different sources, and were apparently afterwards introduced into the other MSS. ; one result of this is that several scholia are reproduced twice. But, besides the scholia derived from these sources, Schol. Vind. contain a large number of others of late date, characterised by in- correct language or by triviality of content (there are many examples in numbers, citations of propositions used, absurd diroplai, and the like). Unlike Schol. Vat, these scholia often quote words from Euclid as a heading (in one case a heading is inserted in Schol. Vind. where a scholium without the heading is quoted from Schol. Vat, see V. No. 14). The explanations given often presuppose very little know- ledge on the part of the reader and frequently contain obscurities and gross errors. Schol. Vind. were collected for use with a MS. of the Theonine class; this follows from the fact that they contain a note on the proposition vulgo VII. 22 interpolated by Theon (given in Heiberg's App. to Vol. II. p. 430), Since the scholium to vn. 39 given in V and p in the text after the title of Book VIII. quotes the proposition as VII. 39, it follows that this scholium must have been written before the interpolation of the two propositions vulgo VII. 20, 22 ; Schol. Vind. contain (vn. No. 80) the first sentence of it, but without the heading referring to VII. 39. Schol. VII. No. 97 quotes VII. 33 as VII. 34, so that the proposition vulgo vn. 22 may have stood in the scholiast's text but not the later interpolation vulgo vn. 20 (later because only found in B in the margin by the first hand). Of course the scholiast had also the interpolations earlier than Theon. For the date of the collection we have a lower limit in the date (12th c.) of MSS. in which the scholia appear. That it was not much earlier than the 12th c. is indicated (1) by the poverty of its contents, (2) by the quality of the ms. of Proclus which was used in the compilation of it (the Munich MS. used by Friedlein with which the scholiast's excerpts are essentially in agreement belongs to the I ith — 12th c), (3) by the fact that Schol. Vind. appear only in MSS. of the 12th c. and no trace of them is found in our MSS. belonging to the 9th — 10th c. in which Schol. Vat. are found. The collection may therefore probably be assigned to the 1 ith c. Perhaps it may be in part due to Psellus who lived towards the end of that century : for in a Florence MS. (Magliabecch. XI, 53 of the 15 th c.) containing a mathematical compendium intended for use in the reading of Aristotle ch. vi] THE SCHOLIA 71 the scholia i. Nos. 40 and 49 appear with the name of Psellus attached. Schoi. Vind. are not found without the admixture of foreign elements in any of our three sources. In 1 there are only very few such in the first hand. In q there are several new scholia in the first hand, for the most part due to the copyist himself. The collection of scholia on Book x. in q (Heiberg's q=) is also in the first hand ; it is not original, and it may perhaps be due to Psellus (Maglb. has some definitions of Book x. with a heading "scholia of... Michael Psellus on the definitions of Euclid's 10th Element" and Schol. X. No. 9), whose name must have been attached to it in the common source of Maglb. and q ; to a great extent it consists of extracts from Schol. Vind. taken from the same source as VI. The scholia q 1 (in an ancient hand in q), confined to Book II., partly belong to Schol, Vind. and partly correspond to b 1 (Bologna MS.), q* and q b are in one hand (Theodorus Antiochita), the nearest to the first hand of q ; they are doubtless due to an early possessor of the MS. of whom we know nothing more. V* has, besides Schol. Vind., a number of scholia which also appear in other MSS., one in BFb, some others in P, and some in v (Codex Vat. IO38, 13th c.) ; these scholia were taken from a source in which many abbreviations were used, as they were often misunderstood by V 1 . Other scholia in V" which are not found in the older sources — some appearing in V* alone— are also not original, as is proved by mistakes or corruptions which they contain ; some others may be due to the copyist himself. V b seldom has scholia common with the other older sources ; for the most part they either appear in V b alone or only in the later sources as v or F* (later scholia in F), some being original, others not. In Book X. V b has three series of numerical examples, ( 1 ) with Greek numerals, (2) alternatives added later, also mostly with Greek numerals, (3) with Arabic numerals. The last class were probably the work of the copyist himself. These examples (cf. p. 74 below) show the facility with which the Byzantines made calculations at the date of the MS. (12th c). They prove also that the use of the Arabic numerals (in the East- Arabian form) was thoroughly established in the 1 2th c. ; they were actually known to the Byzantines a century earlier, since they appear, in the first hand, in an Escurial MS. of the 1 1 th c. Of collections in other hands in V distinguished by Heiberg (see preface to Vol. v.), V 1 has very few scholia which are found in other sources, the greater part being original ; V ! , V s are the work of the copyist himself; V* are so in part only, and contain several scholia from Schol. Vat. and other sources. V* and V J are later than 13th — 14th c, since they are not found in f (cod. Laurent XXVHl, 6) which was copied from V and contains, besides V" V b , the greater part of V 1 and vi. No. 20 of V (in the text). In P there are, besides P* (a quite late hand, probably one of the old Scriptores Grace i at the Vatican), two late hands (P 1 ), one of which has some new and independent scholia, while the other has 7* INTRODUCTION [ch. vi added the greater part of Schol. Vind., partly in the margin and partly on pieces of leaves stitched on. Our sources for Schol. Vat. also contain other elements. In P there were introduced a certain number of extracts from Proclus, to supplement Schol. Vat. to Book I. ; they are all written with a different ink from that used for the oldest part of the MS., and the text is inferior. There are additions in the other sources of Schol. Vat. (F and B) which point to a common source for FB and which are nearly all found in other mss., and, in particular, in Schol. Vind,, which also used the same source ; that they are not assignable to Schol. Vat. results only from their not being found in Vat. Of other additions in F, some are peculiar to F and some common to it and b; but they are not original. F s (scholia in a later hand in F) contains three original scholia ; the rest come from V. B contains, besides scholia common to it and F, b or other sources, several scholia which seem to have been put together by Arethas, who wrote at least a part of them with his own hand. Heiberg has satis6ed himself, by a closer study of b, that the scholia which he denotes by b, ji and b 1 are by one hand ; they are mostly to be found in other sources as well, though some are original. By the same hand (Theodoras Cabasilas, 15th c.) are also the scholia denoted by b", B', b* and B ! . These scholia come in great part from Schol. Vind., and in making these extracts Theodoras probably used one of our sources, 1, mistakes in which often correspond to those of Theodoras. To one scholium is attached the name of Demetrius (who must be Demetrius Cydonius, a friend of Nicolaus Cabasilas, 14th c); but it could not have been written by him, since it appears in B antl Schol. Vind. Nor are all the scholia which bear the name of Theodoras due to Theodoras himself, though some are so. As B' (a late hand in B) contains several of the original scholia of b*, B* must have used b itself as his source, and, as all the scholia in B* are in b, the latter is also the source of the scholia in B 8 which are found in other MSS. B and b were therefore, in the 15th c, in the hands of the same person ; this explains, too, the fact that b in a late hand has some scholia which can only come from B. We arrive then at the conclusion that Theodoras Cabasilas, in the 15th c, owned both the MSS. B and b, and that he transferred to B scholia which he had before written in b, either independently or after other sources, and inversely transferred some scholia from B to b. Further, B' are earlier than Theodoras Cabasilas, who certainly himself wrote B* as well as b' and b 8 . An author's name is also attached to the scholia VI. No. 6 and X. No. 223, which are attributed to Maximus Planudes (end of 13th c) along with scholia on I. 31, x. 14 and X. 18 found in 1 in a quite late hand and published on pp. 46, 47 of Heiberg's dissertation. These seem to have been taken from lectures of Planudes on the Elements by a pupil who used 1 as his copy. There are also in 1 two other Byzantine scholia, written by a late hand, and bearing the names Ioannes and Pediasimus respectively ; ch. vi] THE SCHOLIA 73 these must in like manner have been written by a pupil after lectures of Ioannes Pediasimus (first half of 14th c), and this pupil must also have used 1. Before these scholia were edited by Heiberg, very few of them had been published in the original Greek. The Basel editio princeps has a few (v. No. 1, VI. Nos. 3, 4 and some in Book X.) which are taken, some from the Paris MS. (Paris. Gr, 2343) used by Grynaeus, others probably from the Venice MS. (Marc. 301) also used by him; one published by Heiberg, not in his edition of Euclid but in his paper on the scholia, may also be from Venet. 301, but appears also in Paris. Gr. 2342. The scholia in the Basel edition passed into the Oxford edition in the text, and were also given by August in the Appendix to his Vol. II. Several specimens of the two series of scholia (Vat. and Vind.) were published by C. Wachsmuth {Rhein, Mus. xvm, p. 132 sqq.) and by Knoche {Untersuchungen iiber die neu aufgefundenen Scholien des Proklus, Herford, 1 865). The scholia published in Latin were much more numerous. G. Valla {De expetendis et fugiendis rebus, 1 501) reproduced apparently some 200 of the scholia included in Heiberg's edition. Several of these he obtained from two Modena MSS. which at one time were in his possession (Mutin. Ill B, 4 and II E, 9, both of the 15th c.) ; but he must have used another source as well, containing extracts from other series of scholia, notably Schol. Vind. with which he has some 87 scholia in common. He has also several that are new. Commandinus included in his translation under the title "Scholia antiqua " the greater part of the Schol, Vat. which he certainly obtained from a MS. of the class of Vat. 192; on the whole he adhered closely to the Greek text. Besides these scholia Com- mandinus has the scholia and lemmas which he found in the Basel editio princeps, and also three other scholia not belonging to Schol. Vat., as well as one new scholium (to Xii. 13) not included in Heiberg's edition, which are distinguished by different type and were doubtless taken from the Greek MS. used by him along with the Basel edition. In Conrad Dasypodius' Lexicon matkematicum published in 1573 there is (on fol. 42—44) "Graecum scholion in definitiones Euclidis libri quinti elementorum append ids loco propter pagellas vacantes annexum." This contains four scholia, and part of two others, published in Heiberg's edition, with some variations of readings, and with some new matter added (for which see pp. 64 — 6 of Heiberg's pamphlet). The source of these scholia is revealed to us by another work of Dasypodius, haaci Monachi Scholia in Euclidis elementorum geometriae sex prions tibros per C. Dasypodium in latinum sermonem trans lata et in lucem edita (1579). This work contains, besides excerpts from Proclus on Book I. (in part closely related to Schol. Vind.), some 30 scholia included in Heiberg's edition, several new scholia, and the above-mentioned scholia to the definitions of Book v. published in Greek in 1573. After the scholia follow " Isaaci Monachi 74 INTRODUCTION [ch. vi prolegomena in Euclidis Elementorum geometriae libros" (two definitions of geometry) and " Varia miscellanea ad geometriae cogni- tionem necessaria ab Isaaco Monacho collecta " (mostly the same as pp. 252, 24 — 272, 27 in the Varieu Collectiones included in Hultsch's Heron) ; lastly, a note of Dasypodius to the reader says that these scholia were taken "ex clarissimi viri Joannis Sambuciantiquocodice manu propria Isaaci Monachi scrip to." Isaak Monachus is doubtless Isaak Argyrus, 14th c. ; and Dasypodius used a MS. in which, besides the passage in Hultsch's Variae CoIUctiotus, there were a number of scholia marked in the margin with the name of Isaak (cf. those in b under the name of Theodorus Cabasilas). Whether the new scholia are original cannot be decided until they are published in Greek ; but it is not improbable that they are at all events independent arrange- ments of older scholia. All but five of the others, and all but one of the Greek scholia to Book V., are taken from Schol. Vat. ; three of the excepted ones are from Schol, Vind., and the other three seem to come from F (where some words of them are illegible, but can be supplied by means of Mut. Ill B, 4, which has chese three scholia and generally shows a certain likeness to Isaak's scholia). Dasypodius also published in 1564 the arithmetical commentary of Barlaam the monk (14th c.) on Eucl. Book 11., which finds a place in Appendix IV. to the Scholia in Heiberg's edition. Hultsch has some remarks on the origin of the scholia 1 . He observes that the scholia to Book I. contain a considerable portion of Geminus' commentary on the definitions and are specially valuable because they contain extracts from Geminus only, whereas Proclus, though drawing mainly upon him, quotes from others as well. On the postulates and axioms the scholia give more than is found in Proclus. Hultsch conjectures that the scholium on Book V., No. 3, attributing the discovery of the theorems to Eudoxus but their arrangement to Euclid, represents the tradition going back to Geminus, and that the scholium XIII., No. 1, has the same origin. A word should be added about the numerical illustrations of Euclid's propositions in the scholia to Book x. They contain a large number of calculations with sexagesimal fractions'; the fractions go as far as fourth-sixtieths (i/6o*). Numbers expressed in these fractions are handled with skill and include some results of surprising accuracy* 1 Art. " Eukleides" in Pauly-Wissowa's KeaJ-Ettcyflopadit. 1 Hultsch has written upon these in Biblietluea Matktmatisa, v a , 1904, pp. 155 — 133. * Thus v'W) is given (allowing for a slight correction by means of the context) as 5 1 i 1 46" 10'", which gives for V3 the value t 43 Jj" 13"', being the same value as that given by Hipparchus in his Table of Chords, and correct to the seventh decimal place. Similarly J 8 is given as 1 49' 41" 10'" 10"", which is equivalent to \'i= IV 141 133s- Hultsch gives instances of the various operations, addition, subtraction, etc., carried out in these fractions, and shows how the extraction of the square root was effected. Cf. T- L. Heath, History ttj Creek Mathematics, t. , pp. 50 — 03- CHAPTER VII. EUCLID IN ARABIA, We are told by Hajl Khalfa' that the Caliph al-MansQr (754-775) sent a mission to the Byzantine Emperor as the result of which he obtained from him a copy of Euclid among other Greek books, and again that the Caliph al-Ma'mun (813-833) obtained manuscripts of Euclid, among others, from the Byzantines. The version of the Elements by al-Hajjaj b. Yusuf b. Matar is, if not the very first, at least one of the first books translated from the Greek into Arabic'. According to the Fikrist* it was translated by al-Hajjaj twice ; the first translation was known as " Haruni" (" for Harun"), the second bore the name "Ma'muni" ("for al-Ma'mun") and was the more trust- worthy. Six Books of the second of these versions survive in a Leiden MS. (Codex Leidensis 399, 1) now in part published by Besthorn and Heiberg*. In the preface to this ms. it is stated that, in the reign of Harun ar-Rashid (786-809), al-Hajjaj was commanded by Yahya b. Khalid b. Barmak to translate the book into Arabic. Then, when al-Ma'mun became Caliph, as he was devoted to learning, al-Hajjaj saw that he would secure the favour of al-Ma'mun "if he illustrated and expounded this book and reduced it to smaller dimensions. He accordingly left out the superfluities, filled up the gaps, corrected or removed the errors, until he had gone through the book and reduced it, when corrected and explained, to smaller dimensions, as in this copy, but without altering the substance, for the use of men endowed with ability and devoted to learning, the earlier edition being left in the hands of readers." The Fikrist goes on to say that the work was next translated by Ishaq b. Hunain, and that this translation was improved by Thabit b, Qurra. This Abu Ya'qub Ishaq b. Hunain b. Ishaq al-Tbadi (d. 910) was the son of the most famous of Arabic translators, Hunain b. Ishaq al-'lbadi (809-873), a Christian and physician to the Caliph al- Mutawakkil (847-861). There seems to be no doubt that Ishaq, who 1 ijxicon bibliegr. et tncyclop. ed. FliigeL lilt pp. 91 , pa. 1 Klamroth, Zeitschrift der Deittschen Morgenltindisehcn Gesellsihaft, XXXV. p, 303. 3 Fihrist (tr. Suter), p. 16. * Codex Leidensis 399, 1. Eueiidis Elementa ex interpretaiione at-Hadschdsehadschii eum cgmmmfariis al-Narizii* Elauniae, part C. i. 1893, part t. ii. 1897, part U- L [900, pari ][. ii. 1905, part III. i. 1910. 76 INTRODUCTION [ch. vii must have known Greek as well as his father, made his translation direct from the Greek. The revision must apparently have been the subject of an arrangement between Ishaq and Thabit, as the latter died in 901 or nine years before Ishaq, Thabit undoubtedly consulted Greek MSS. for the purposes of his revision. This is expressly stated in a marginal note to a Hebrew version of the Elements, made from Ishaq's, attributed to one of two scholars belonging to the same family, viz. either to Moses b. Tibbon (about 1 244- 1 274) or to Jakob b. Machir (who died soon after 1306) 1 . Moreover Thabit observes, on the pro- position which he gives as ix. 31, that he had not found this proposition and the one before it in the Greek but only in the Arabic ; from which statement Klamroth draws two conclusions, (1) that the Arabs had already begun to interest themselves in the authenticity of the text and (2) that Thabit did not alter the numbers of the propositions in Ishaq's translation'. The Fihrist also says that Yuhanna al-Qass (i.e. " the Priest ") had seen in the Greek copy in his possession the pro- position in Book I. which Thabit took credit for, and that this was confirmed by Nazlf, the physician, to whom Yuhanna had shown it This proposition may have been wanting in Ishaq, and Thabit may have added it, but without claiming it as his own discovery*. As a fact, t. 45 is missing in the translation by al-Hajjaj. The original version of Ishaq without the improvements by Thabit has probably not survived any more than the first of the two versions by al-Hajjaj ; the divergences between the MSS, are apparently due to the voluntary or involuntary changes of copyists, the former class varying according to the degree of mathematical knowledge possessed by the copyists and the extent to which they were influenced by considerations of practical utility for teaching purposes*. Two MSS. of the Ishaq-Thabit version exist in the Bodleian Library (No. 279 belonging to the year 1238, and No. 280 written in 1260-1) 11 ; Books I. — XIII. are in the Ishaq-Thabit version, the non- Euclidean Books XIV., XV. in the translation of Qusta b. Luqa at-Ba'labakki (d. about 912). The first of these MSS. (No. 279) is thafctX)) used by Klamroth for the purpose of his paper on the Arabian Euclid. The other MS. used by Klamroth is (K) Kjobenhavn LXXXI, undated but probably of the 13th c, containing Books v. — xv., Boiks V. — X. being in the Ishaq-Thabit version, Books XI. — XIII. purporting to be in al-Hajjaj's translation, and Books XIV, xv. in the version of Qusta b. Luqa. In not a few propositions K and O show not the slightest difference, and, even where the proofs show considerable differences, they are generally such that, by a careful comparison, it is possible to reconstruct the common archetype, so that it is fairly clear that we have in these cases, not two recensions of one translation, but arbitrarily altered and 1 Steinschneider, Ziitschrift fur Math. «. Physik, XXXI., hist. -lilt. Abtheilung, pp. 85, 86, 90. J Klamtoth, p. »79, ■ Steinschneider, p, 88. * Klamroth, p. 306. 1 These MSS. are described by Nicoll and Pusey, Catatogus tod. m$s. orient, bibl. Bca- hiattot, pt. u. 1835 (pp. 157—161). ch. vh] EUCLID IN ARABIA 77 shortened copies of one and the same recension 1 . The Bodleian MS. No. 280 contains a preface, translated by Nicoll, which cannot be by Thibit himself because it mentions Avkenna (980-1037) and other later authors. The MS. was written at Maraga in the year 1260-1 and has in the margin readings and emendations from the edition of Nasiraddln at-Tusi (shortly to be mentioned) who was living at Maraga at the time, is it possible that at-Tusl himself is the author of the preface*? Be this as it may, the preface is interesting because it throws light on the liberties which the Arabians allowed themselves to take with the text. After the observation that the book (in spite of the labours of many editors) is not free from errors, obscurities, redundancies, omissions etc., and is without certain definitions neces- sary for the proofs, it goes on to say that the man has not yet been found who could make it perfect, and next proceeds to explain (1) that Avicenna "cut out postulates and many Definitions" and attempted to clear up difficult and obscure passages, (2) that Abu'l Wafa al-Buzjanl (939-99?) "introduced unnecessary additions and left out many things of great importance and entirely necessary," inasmuch as he was too long in various places in Book VI. and too short in Book X. where he left out entirely the proofs of the apotomae, while he made an unsuccessful attempt to emend XII. 14, (3) that Abu Ja'far al-Khazin (d, between 961 and 971) arranged the postulates excellently but " disturbed the number and order of the propositions, reduced several propositions to one " etc. Next the preface describes the editor's own claims' and then ends with the sentences, " But we have kept to the order of the books and propositions in the work itself (i.e. Euclid's) except in the twelfth and thirteenth books. For we have dealt in Book xni, with the (solid) bodies and in Book XII. with the surfaces by themselves." After Thabit the Fihrist mentions Abu 'Uthman ad-Dimashql as having translated some Books of the Elements including Book X. (It is Abu 'Uthman's translation of Pappus' commentary on Book X. which Woepcke discovered at Paris.) The Fikrist adds also that " Nazif the physician told me that he had seen the tenth Book of Euclid in Greek, that it had 40 propositions more than the version in common circulation which had 109 propositions, and that he had determined to translate it into Arabic." But the third form of the Arabian Euclid actually accessible to us is the edition of AbQ Ja'far Muh. b. Muh. b. al-Hasan Nasiraddln at-Tusi (whom we shall call at-Tusi for short), born at Tus ( m Khurasan) in 1201 (d. 1274). This edition appeared in two forms, a larger and a smaller. The larger is said to survive in Florence only (Pal. 272 and 313, the latter MS. containing only six Books) ; this was published at Rome in 1 594, and, remarkably enough, some copies of 1 KlamrOth, pp. jort — S. * Steinschneider, p. 98. Heiberg has Quoted the whole of this preface in the Ztitschrift fur Math. ft, Phytik, XXIX., hist.-litt. Abth. p. 16. * This seems to include a rearrangement of the contents of Books xiv., xv. added to the EltmftUs. 78 INTRODUCTION |ch. vu this edition are to be found with 12 and some with 13 Books, some with a Latin title and some without'. But the book was printed in Arabic, so that Kastner remarks that he will say as much about it as can be said about a book which one cannot read*. The shorter form, which however, in most MSS., is in 15 Books, survives at Berlin, Munich, Oxford, British Museum (974, 1334*, 1335), Paris (2465, 2466), India Office, and Constantinople ; it was printed at Constantinople in 1 80 1, and the first six Books at Calcutta in 1824*. At-Tusi's work is however not a trattslatiott of Euclid's text, but a re-written Euclid based on the older Arabic translations. In this respect it seems to be like the Latin version of the Elements by Campanus (Campano), which was first published by Erhard Ratdolt at Venice in 1482 (the first printed edition of Euclid*). Campanus (13th c.) was a mathematician, and it is likely enough that he allowed himself the same liberty as at-TGst in reproducing Euclid. What- ever may be the relation between Campanus' version and that of Athelhard of Bath (about 1 1 20), and whether, as Curtze thinks*, they both used one and the same Latin version of 10th — 1 ith c, or whether Campanus used Athel hard's version in the same way as at- T fist used those of his predecessors 7 , it is certain that both versions came from an Arabian source, as is evident from the occurrence of Arabic words in them*. Campanus' version is not of much service for the purpose of forming a judgment on the relative authenticity of the Greek and Arabian tradition ; but it sometimes preserves traces of the purer source, as when it omits Theon's addition to vi. 33'. A curious circumstance is that, while Campanus' version agrees with at-Tusi's in the number of the propositions in all the genuine Euclidean Books except V. and IX., it agrees with At hd hard's in having 34 propositions in Book V. (as against 25 in other versions), which confirms the view that the two are not independent, and also leads, as Klamroth says, to this dilemma: either the additions to Book V. are Athelhard's own, or he used an Arabian Euclid which is not known to us". Heiberg also notes that Campanus' Books XIV., XV. show a certain agreement with the preface to the Thabit-Ishiq version, in which the author claims to have (1) given a method of inscribing spheres in the five regular solids, (2) carried further the solution of the problem how 1 Suter, Dit Afatkeniatiier und Asironomen dtr Arobtr, p. iji. The Lai in title- a Euilidii tltttteniorum geometrkorum libri trtdtcim. Ex traditions dottissimi Nasiridini Ttami nunc frimum arabite itnprtssi. Kumae in typograpbia Medicea MDXCIV. Cum licentia superiorum. 1 Kastner, Gesehiehie der Matntmatik, I. p. 367. * Suter has a note that this US. is very old, having been copied from the original in the author's lifetime. * Suter, p. iji. ■ Described by Kastner, Gathitkte dtr Mathrmatii, 1. pp. 389 — iog, and by Weiss - enbom, Die Uberutxungtn da Etttlid durch Campano und Zamberti, Halle a. S., 1881, pp. 1 — J. See also infra. Chapter vni, p, 97. ■ Sonderabdruck des Jahrtsberitklts titer die Fortiekrittt dtr klastitcktn AUtrlnumt- wisstnsehaft vem. Okt. 1S79 — 18B3, Berlin, 1884. 7 Klamroth, p. 171. * Curtw, op. tit. p. 10; Heiberg, Exklid-Studitn, p. 178. * Heiberg's Euclid, vol. v. p. ci. l0 Klamroth, pp. 173 — 4. ch. vn] EUCLID IN ARABIA 79 to inscribe any one of the solids in any other and (3) noted the cases where this could not be done 1 . With a view to arriving at what may be called a common measure of the Arabian tradition, it is necessary to compare, in the first place, the numbers of propositions in the various Books. Haji Khalfa says that al-Hajjaj's translation contained 468 propositions, and Thabit's 478 ; this is stated on the authority of at-TQsI, whose own edition contained 468*. The fact that Thabit's version had 478 propositions is confirmed by an index in the Bodleian MS. 279 (called O by Klamroth). A register at the beginning of the Codex Leiden sis 399, 1 which gives Ishaq's numbers (although the translation is that of al-FIajjaj) apparently makes the total 479 propositions (the number in Book XIV. being apparently 11, instead of the 10 of O 1 ). I subjoin a table of relative numbers taken from Klamroth, to which I have added the corresponding numbers in August's and Heiberg's editions of the Greek text The Arabian Euclid The Greek Euclid Books Ishiq at-TOsI Campaniu Gregory August Heiberg 1 48 48 48 48 48 48 11 14 14 >4 14 14 14 1U 36 36 36 37 37 37 IV 16 16 16 16 16 16 V *S n 34 *5 n *s VI 33 32 32 33 33 33 VII 39 39 39 41 41 39 VIII *7 'S IS 37 27 37 IX 3« 36 39 36 36 36 X 109 107 107 117 116 115 XI 41 4' 4< 40 40 39 XII ■5 '5 '5 iS iS 18 XIII 21 iS ■ S ■ S iS 18 462 453 464 470 469 465 [XIV 10 10 18 7 ? XV 6 6 '3 10 478 468 495 "487 a The numbers in the case of Heiberg include all propositions which he has printed in the text ; they include therefore xiii. 6 and in. 12 now to be regarded as spurious, and X. .1 12 — 115 which he brackets as doubtful. He does not number the propositions in Books XIV., XV., but I conclude that the numbers in P reach at least 9 in xiv., and 9 in XV. 1 Heiberg, Zttisthrift fur Math. u. PAjriii, xxix,, hist.-liu. Atrtheilung, p. 11. ' Klamroth, p. 17+; Steinachneider, Zatschrift fur Math. u. Physth, XXXI., hid Mitt. Abth. p. 9*i. ■ BeMhorn- Heiberg read " 11?" aa the number, Klamroth had read it as ji (p. 173). 8o INTRODUCTION [en. vii The Fihrist confirms the number 109 for Book X., from which K lam roth concludes that Ishaq's version was considered as by far the most authoritative. In the text of O, Book IV. consists of 1 7 propositions and Book XIV. of 12, differing in this respect from its own table of contents ; IV. 15, 16 in O are really two proofs of the same proposition. In al-Hajjaj's version Book I. consists of 47 propositions only, I. 45 being omitted. It has also one proposition fewer in Book III., the Heron ic proposition m. 12 being no doubt omitted. In speaking of particular propositions, I shall use Heiberg's numbering, except where otherwise stated. The difference of 10 propositions between Tha bit- Ishaq and at-TusT is accounted for thus : (1) The three propositions vi. 12 and X, 28, 39 which both Ishaq and the Greek text have are omitted in at-Tusi, (2) Ishaq divides each of the propositions xm. 1 — 3 into two, making six instead of three in at-Tusi and in the Greek. {3) Ishaq has four propositions (numbered by him vm. 24, 2$, IX. 30, 3 1) which are neither in the Greek Euclid nor in at-Tusi. Apart from the above differences al-iiajjaj (so far as we know), Ishaq and at-Tusi agree , but their Euclid shows many differences from our Greek text. These differences we will classify as follows 1 . 1. Prepositions. The Arabian Euclid omits VII. 20, 22 of Gregory's and August's editions (Heiberg, App. to VoL 11. pp. 428-32) ; vm. 16, 17; X. 7, 8, 13, 16, 24, 112, 113, 114, besides a lemma vulgo X. 13, the proposition X. 1 1 7 of Gregory's edition, and the scholium at the end of the Book (see for these Heiberg's Appendix to Vol. III. pp. 382, 408 — 416) ; XI. 38 in Gregory and August (Heiberg, App. to Vol. I v. p. 354); XII. 6, 13, 14 ; (also all but the first third of Book xv.). The Arabian Euclid makes III. 11, 12 into one proposition, and divides some propositions (X. 31, 32 ; xi, 31, 34; xin. 1 — 3) into two each. The order is also changed in the Arabic to the following extent. v. 12, 13 are interchanged and the order in Books Vi,, vn, IX. — XIII. is : VL 1—8, 13, II, 12, 9, IO, 14—17, 19, 20, 18, 21, 22, 24, 26, 23, 25. 27—30, 32, 31, 33. VII. I — 20, 22, 21, 23 — 28, 31, 32, 29, 30, 33 — 39. IX i — 13, 20, 14—19, 21 — 25, 27, 26, 28 — 36, with two new pro- positions coming before prop. 30. x. 1—6, 9—12. IS. 14. i7— 2 3. 26—28, 25, 29—30, 31, 32, 33— in, us- XL 1—30,31.32,34.33.35—39 xii. 1 — 5, 7, 9, 8, 10, 12, 11, 15, 16 — 18. xm. 1—3, 5. 4, 6, 7, 12, 9, io, 8, 11, 13, is, 14. 16—18. 1 See KJamroth, pp. 175 — 6, 180, 161 — 4, %\± — 15, 516 ; Heibeig, vol. v. pp. xcvi, xcvii. CH.vn] EUCLID IN ARABIA Si 2. Definitions. The Arabic omits the following definitions: iv. Deff. 3 — 7, VII. Def. 9 (or io), xi. Deff, 5 — 7, 15, 17. 23, 25—28; but it has the spurious definitions VI. Deff. 2, 5, and those of proportion and ordered proportion in Book V. (Deff. 8, 19 August), and wrongly interchanges v. Deff. 1 1, 12 and also vi. Deff. 3, 4. The order of the definitions is also different in Book VII. where, after Def. II, the order is 12, 14, 13, 15, 16, 19, 20, 17, 18, 21, 22, 23, and in Book xi. where the order is 1, 2, 3,4, 8, io, 9, 13, 14, 16, 12, 21, 22, 18, 19, 20, 11, 24. 3. Lemmas and porisms. All are omitted in the Arabic except the porisms to vi. 8, vin, 2, X. 3 ; but there are slight additions here and there, not found in the Greek, e.g. in vm. 14, 15 (in K). 4. Alternative proofs. These are all omitted in the Arabic, except that in X, 105, 106 they are substituted for the genuine proofs ; but one or two alternative proofs are peculiar to the Arabic (vi. 32 and vm. 4, 6). The analyses and syntheses to XIII. I — 5 are also omitted in the Arabic. K lam roth is inclined, on a consideration of all these differences, to give preference to the Arabian tradition over the Greek (1) "on historical grounds," subject to the proviso that no Greek MS. as ancient as the 8th c, is found to contradict his conclusions, which are based generally (2) on the improbability that the Arabs would have omitted so much if they had found it in their Greek MSS., it being clear from the Fihrist that the Arabs had already shown an anxiety for a pure text, and that the old translators were subjected in this matter to the check of public criticism. Against the " historical grounds," Heiberg is able to bring a considerable amount of evidence 1 . First of all there is the British Museum palimpsest (L) of the 7th or the beginning of the 8th c. This has fragments of propositions in Book X. which are omitted in the Arabic ; the numbering of one proposition, which agrees with the numbering in other Greek ms., is not comprehensible on the assumption that eight preceding propositions were omitted in it, as they are in the Arabic; and lastly, the readings in L are tolerably like those of our MSS., and surprisingly tike those of B. It is also to be noted that, although P dates from the 10th c. only, it contains, according to all appearance, an ante-Theonine recension. Moreover there is positive evidence against certain omissions by the Arabians. At-Tusi omits VI. 12, but it is scarcely possible that, if Eutocius had not had it, he would have quoted Vi. 23 by that number*. This quotation of VI. 23 by Eutocius also tells against Ishaq who has the proposition as vi. 25. Again, Simplicius quotes VI. 10 by that number, whereas it is VI. 13 in Ishaq ; and Pappus quotes, by number, XIII. 2 (Ishaq 3, 4), X.III. 4 (Ishaq 8), XIII. 16 (Ishaq 19). 1 Heiberg in Ztitxhrift fur Math. u. Pfytit, XXIX., bilt.-litt. Ablh. p. 3 sqq. 1 Apollonius, ed. Heiberg, vol. II. p. »i8, 3 — 5. 8a INTRODUCTION [en. vii On the other hand the contraction of III. II, 12 into one proposition in the Arabic tells in favour of the Arabic. Further, the omission of certain porisms in the Arabic cannot be supported; for Pappus quotes the porism to XIII. 17 1 , Procius those to II. 4, in. 1, vii. 2\ and Simplicius that to IV. 15. Lastly, some propositions omitted in the Arabic are required in later propositions. Thus X. 13 is used in X. 18, 22, 23, 26 etc. ; X. 17 is wanted in X. 18, 26, 36; xn. 6, 13 are required for XII. 1 1 and XII. 15 respectively. It must also be remembered that some of the things which were properly omitted by the Arabians are omitted or marked as doubtful in Greek MSS. also, especially in P, and others are rightly suspected for other reasons (e.g. a number of alternative proofs, lemmas, and porisms, as well as the analyses and syntheses of XII I. 1—5). On the other hand, the Arabic has certain interpolations peculiar to our inferior MSS. (cf. the definition VI. Def. 2 and those of proportion and ordered proportion), Heiberg comes to the general conclusion that, not only is the Arabic tradition not to be preferred offhand to that of the Greek mss., but it must be regarded as inferior in authority. It is a question how far the differences shown in the Arabic are due to the use of Greek MSS. differing from those which have been most used as the basis of our text, and how far to the arbitrary changes made by the Arabians themselves. Changes of order and arbitrary omissions could not surprise us, in view of the preface above quoted from the Oxford MS. of Thabit-Ishaq, with its allusion to the many important and necessary things left out by Abu '1 Wafa and to the author's, own rearrangement of Books XII., xm. But there is evidence of differences due to the use by the Arabs of other Greek Mss. Heiberg' is able to show considerable resemblances between the Arabic text and the Bologna MS. b in that part of the MS. where it diverges so remarkably from our other MSS. (see the short description of it above, p. 49) ; in illustration he gives a comparison of the proofs of XII. 7 in b and in the Arabic respectively, and points to the omission in both of the proposition given in Gregory's edition as XI 38, and to a remark- able agreement between them as regards the order of the propositions of Book XII. As above stated, the remarkable divergence of b only affects Books xi. (at end) and XII. ; and Book xm. in b shows none of the transpositions and other peculiarities of the Arabic. There are many differences between b and the Arabic, especially in the definitions of Book XX, as well as in Book xm. It is therefore a question whether the Arabians made arbitrary changes, or the Arabic form is the more ancient, and b has been altered through contact with other MSS. Heiberg points out that the Arabians must be alone responsible for their definition of a prism, which only covers a prism with a triangular base. This could not have been Euclid's own, for the word prism already has the wider meaning in Archimedes, and 1 Pippus, V. p. 436, 5. * Procius, pp. 303 — 4. * Ztttschrift fur Math. u. Physik, XXIX., hbt.-titt. Al.ib. p. 6«qq. CH, vii] EUCLID IN ARABIA 83 Euclid himself speaks of prisms with parallelograms and polygons as bases (xi. 39; XII. 10). Moreover, a Greek would not have been likely to leave out the definitions of the " Platonic " regular solids. Heiberg considers that the Arabian translator had before him a MS. which was related to b, but diverged still further from the rest of our MSS. He does not think that there is evidence of the existence of a redaction of Books I. — X. similar to that of Books XI., XII. in b ; for K I am roth observes that it is the Books on solid geometry (XI. — XIIL) which are more remarkable than the others for omissions and shorter proofs, and it is a noteworthy coincidence that it is just in these Books that we have a divergent text in b. An advantage in the Arabic version is the omission of VII. Def. 10, although, as Iambliehus had it, it may have been deliberately omitted by the Arabic translator. Another advantage is the omission of the analyses and syntheses of XI II. I — 5 ; but again these may have been omitted purposely, as were evidently a number of porisms which are really necessary. One or two remarks may be added about the Arabic versions as compared with one another, Al-Hajjaj's object seems to have been less to give a faithful reflection of the original than to write a useful and convenient mathematical text-book. One characteristic of it is the careful references to earlier propositions when their results are used. Such specific quotations of earlier propositions are rare in Euclid ; but in al-Hajjaj we find not only such phrases as "by prop, so and so," " which was proved " or " which we showed how to do in prop, so and so," but also still longer phrases. Sometimes he repeats a construction, as in I. 44 where, instead of constructing " the parallelo- gram BEFG equal to the triangle C in the angle EBG which is equal to the angle })" and placing it in a certain position, he produces AB to G, making BG equal to half DE (the base of the triangle CDE in his figure), and on GB so constructs the parallelogram BHKG by I. 42 that it is equal to the triangle CDE, and its angle GBH is equal to the given angle. Secondly, al-Hajjaj, in the arithmetical books, in the theory of proportion, in the applications of the Pythagorean 1. 47, and generally where possible, illustrates the proofs by numerical examples. It is true, observes Klamroth, that these examples are not apparently separated from the commentary of an-Nairtzi, and might not there- fore have been due to al-Hajjaj himself; but the marginal notes to the Hebrew translation in Municn MS. 36 show that these additions were in the copy of al-Hajjaj used by the translator, for they expressly give these proofs in numbers as variants taken from al-Hajjaj 1 . These characteristics, together with al-Hajjaj 's freer formulation of the propositions and expansion of the proofs, constitute an in- telligible reason why Ishaq should have undertaken a fresh translation from the Greek. Klamroth calls Ishaq's version a model of a good translation of a mathematical text ; the introductory and transitional Klamruth, p. 310 ; Steiiischneidet, pp. 85 — 6. 84 INTRODUCTION fc H - vii phrases are stereotyped and few in number, the technical terms are simply and consistently rendered, and the less formal expressions connect themselves as closely with the Greek as is consistent with intelligibility and the character of the Arabic language. Only in isolated cases does the formulation of definitions and enunciations differ to any considerable extent from the original. In general, his object seems to have been to get rid of difficulties and unevennesses in the Greek text by neat devices, while at the same time giving a faithful reproduction of it 1 . There are curious points of contact between the versions of al-Hajjaj and Thabit-Ishaq. For example, the definitions and enunciations of propositions are often word for word the same. Presumably this is owing to the fact that Ishaq found these de- finitions and enunciations already established in the schools in his time, where they would no doubt be leamt by heart, and refrained from translating them afresh, merely adopting the older version with some changes'. Secondly, there is remarkable agreement between the Arabic versions as regards the figures, which show considerable variations from the figures of the Greek text, especially as regards the letters ; this is also probably to be explained in the same way, all the later translators having most likely borrowed al-Hajjaj's adaptation of the Greek figures'. Lastly, it is remarkable that the version of Books XI. — XIII. in the KjfSbenhavn MS. (K), purporting to be by al-Hajjaj, is almost exactly the same as the Thabit-Ishaq version of the same Books in O. Klamroth conjectures that Ishaq may not have translated the Books on solid geometry at all, and that Thabit took them from al-Hajjaj, only making some changes in order to fit them to the translation of Ishaq'. From the facts (l) that at-TusI's edition had the same number of propositions (468) as al-Hajjaj's version, while Thabit- 1 shaq's had 478, and (2) that at-Tusl has the same careful references to earlier propositions, Klamroth concludes that at-Tusi deliberately preferred al-Hajjaj's version to that of Ishaq', Heiberg, however,, points out (1) that at-Tusi left out VI. 12 which, if we may judge by Klamroth's silence, al-Hajjaj had, and (2) al-Hajjaj's version had one proposition less in Books 1. and in. than at-Tusl has. Besides, in a passage quoted by Hajl Khalfa' from at-Tusi, the latter says that "he separated the things which, in the approved editions, were taken from the archetype from the things which had been added thereto," indicating that he had compiled his edition from both the earlier translations'. There were a large number of Arabian commentaries on, or reproductions of, the Elements or portions thereof, which will be 1 Klamroth, p. 390, illustrates Ish&q's method by his way of distinguishing l^op^dftur (to be congruent with) and lipapubfadai (to be applied to), the confusion of which by trans- lators was animadverted on by Savile. Ishiq avoided the confusion by using two entirely different words. ' Klamroth, pp. JIO— I. ' ibid. p. 187, • ibid. pp. 304—5. * ibid- p. "a 74. • Hail Khalfa, I. p. 383. 7 Heiberg, Zritschrift ftir Math. u.Phjrsik, XXIX., hist.-litt. Abth. pp.3, 3. en. vn] EUCLID IN ARABIA 85 found fully noticed by Steinschn eider 1 . I shall mention here the commentators etc. referred to in the Fikrist, with a few others. 1. Abu '1 'Abbas al-Fadl b. Hatim an-Nairlzi (born at Nairiz, died about 922) has already been mentioned'. His commentary survives, as regards Books 1. — VI., in the Codex Leidensis 399, 1, now edited, as to four Books, by Besthorn and Heiberg, and as regards Books I. — x. in the Latin translation made by Gherard of Cremona in -the 12th c. and now published by Curtze from a Cracow MS.* Its importance lies mainly in the quotations from Heron and Simplicius. 2. Ahmad b. 'Umar al-KarablsI (date uncertain, probably 9th — 10th c), " who was among the most distinguished geometers and arithmeticians'." 3. A 1-' Abbas b. Sa'ld al-Jauhan (fl. 830) was one of the astro- nomical observers under al-Ma'mun, but devoted himself mostly to geometry. He wrote a commentary to the whole of the Elements, from the beginning to the end ; also the " Book of the propositions which he added to the first book of Euclid*." 4. Muh. >. 'Isa Abu 'Abdallah al-Mahan! (d. between 874 and 884) wrotej according to the Fikrist, (1) a commentary on Eucl. Book v., (2) "On proportion," (3) "On the 26 propositions of the first Book of Euclid which are proved without reductio ad absurdum*." The work " On proportion " survives and is probably identical with, or part of, the commentary on Book V. 7 He also wrote, what is not mentioned by the Fihrist, a commentary on Eucl. Book X., a fragment of which survives in a Paris MS." 5. Abu Ja'far al- Khazin (i.e. " the treasurer " or " librarian "), one of the first mathematicians and astronomers of his time, was born in Khurasan and died between the years 961 and 971. The Fikrist speaks of him as having written a commentary on the whole of the Elements*, but only the commentary on the beginning of Book X. survives (in Leiden, Berlin and Paris) ; therefore either the notes on the rest of the Books have perished, or the Fihrist is in error 10 . The latter would seem more probable, for, at the end of his commentary, al-Khazin remarks that the rest had already been commented on by Sulaiman b. 'Usma (Leiden MS.) 11 or "Oqba (Surer), to be mentioned below. At-Khazin's method is criticised unfavourably in the preface to the Oxford MS. quoted by Nicoll (see p. 77 above). 6. Abu '1 WaiS al-Buzjanl (940-997), one of the greatest Arabian mathematicians, wrote a commentary on the Elements, but 1 Steinschneider, Zeiischriftfiir Mali. u. Physii, XXXI., hist.-litt. Abth, pp. 86 sqq. 1 Steinschneider, p. 86, Fihrist (Ir. Suter), pp. [6, 67 ; Suter, Dit Maiktmaliktr ttnd Astrmniien Jrr Arabcr (1900), p. +J. ' SuppitHUHtum ad Eudiftis Optra omnia, ed. Heiberg and Mtnge, Leipzig, 1899. 1 Fikrist , pp. 16, 38 [ Sleinschneider, p. 87 ; Suter, p. 6j. * Fihrist, pp. 16, 15; Steinschneider, p. gg . Suter, p. 12. * FUrist, pp. 16, ii, 58. 7 Suter, p. 16, note, quotes the Para MS. 1467, 16 s containing the work "on proportion" u the authority for this conjecture. 1 MS. i+S7, 39° (ef. Woepcke in Mhn. pr4s. a raead. dit sriemes, XI v., i8j6\ p. 669). * Fikrist, p. 17. ™ Suter, p. j8, note b. " Steintchneider, p. 89. 86 INTRODUCTION [ch. vii did not complete it 1 . His method is also unfavourably regarded in the same preface to the Oxford MS. 28a According to Haji Khalfa, he also wrote a book on geometrical constructions, in thirteen chapters. Apparently a book answering to this description was compiled by a gifted pupil from lectures by Abu '1 Wafa, and a Paris MS. (Anc. fonds 169) contains a Persian translation of this work, not that of Abu '1 Wafa himself. An analysis of the work was given by Woepcke*, and some particulars will be found in Cantor'. Abu '1 Wafa also wrote a commentary on Diophantus, as well as a separate "book of proofs to the propositions which Diophantus used in his book and to what he (Abu '1 Wafa) employed in his commentary*." 7. Ibn Rahawaihi al-Arjanl also commented on Eucl. Book X.* 8. 'All b, Ahmad Abu '1-Qasim al- AntakI (d. 987) wrote a commentary on the whole book 1 ; part of it seems to survive (from the 5th Book onwards) at Oxford (Catal. MSS. orient. II. 28 1) 7 . 9. Sind b. 'AH Abu 't-Taiyib was a Jew who went over to Islam in the time of al-Ma'mun, and was received among his astro- nomical observers, whose head ht became* (about 830); he died after 864. He wrote a commentary on the whole of the Elements ; " Abu 'All saw nine books of it, and a part of the tenth 8 ." His book " On the Apotomae and the Medials," mentioned by the Fihrist, may be the same as, or part of, his commentary on Book x. 10. Abu Yusuf Ya'qQb b, Muh. ar-Razi "wrote a commentary on Book X., and that an excellent one, at the instance of Ibn al- •Amid"." 11. The Fihrist next mentions al-Kindi (Abu Yusuf Ya'qub b. Ishaq b. as-Sabbah al-Kindi, d. about 873), as the author (1) of a work * on the objects of Euclid's book," in which occurs the statement that the Elements were originally written by Apollonius, the carpenter (see above, p. 5 and note), (2) of a book "on the improvement of Euclid's work," and (3) of another "on the improvement of the 14th and 15th Books of Euclid." "He was the most distinguished man of his time, and stood alone in the knowledge of the old sciences collectively ; he was called ' the philosopher of the Arabians ' ; his writings treat of the most different branches of knowledge, as logic, philosophy, geometry, calculation, arithmetic, music, astronomy and others"." Among the other geometrical works of al-Kindi mentioned by the Fihrist 1 * are treatises on the closer investigation of the results of Archimedes concerning the measure of the diameter of a circle in terms of its circumference, on the construction of the figure of the two mean proportionals, on the approximate determination of the chords ^r. v. T. v. pp. 118 — 156 and 309— $$$. * Fihrist, p. 17 ; Suter, p. ij. ■ Fihrist, p. 17. 7 Suter, p, 64. • Fikrisl, p. 17, 19 j Suter, pp. 13, i+. » Fihrist, p. r;. " Fihrist, p. 17; Suter, p. 66. " Fihrist, p. 17, 10 — [J. u The mere catalogue of al-KindT's works on the various branches of science takes up four octavo pages {1 1— ij) of Suter's translation of the Fihrist. I ch. vii] EUCLID IN ARABIA 87 of the circle, on the approximate determination of the chord (side) of the nonagon, on the division of triangles and quadrilaterals and con- structions for that purpose, on the manner of construction of a circle which is equal to the surface of a given cylinder, on the division of the circle, in three chapters etc. 12. The physician Nazif b. Yumn (or Yaman) al-Qass ("the priest ") is mentioned by the Fikrist as having seen a Greek copy of Eucl. Book X. which had 40 more propositions than that which was in general circulation (containing 109), and having determined to translate it into Arabic 1 . Fragments of such a translation exist at Paris, Nos. 18 and 34. of the MS. 24s 7 (952, 2 Suppl. Arab, in Woepcke's tract); No. 18 contains "additions to some propositions of the 10th Book, existing in the Greek language'." Nazif must have died about 990*, 13. YGhanna b. Yusuf b. al-Harith b. al-Bitriq al-Qass (d. about 980) lectures' on the Elements and other geometrical books, made translations from the Greek, and wrote a tract on the " proof" of the case of two straight lines both meeting a third and making with it, on one side, two angles together less than two right angles*. Nothing of his appears to survive, except that a tract " on rational and irrational magnitudes," No. 48 in the Paris MS. just mentioned, is attributed to him. 14. Abu Muh. al- Hasan b. 'Ubaidallah b. Sulatman b. Wahb (d. 901) was a geometer of distinction, who wrote works under the two distinct titles " A commentary on the difficult parts of the work of Euclid " and " The Book on Proportion'." Suter thinks that an- other reading is possible in the case of the second title, and that it may refer to the Euclidean work " on the divisions (of figures)*." 15. Qusta b. Luqi al-Ba'labakkl (d. about 912), a physician, philosopher, astronomer, mathematician and translator, wrote " on the difficult passages of Euclid's book" and "on the solution of arith- metical problems from the third book of Euclid 7 "; also an "intro- duction to geometry," in the form of question and answer". 16. Thabit b. Qurra (826-90 1), besides translating some parts of Archimedes and Books V. — VII. of the Conks of Apollonius, and revising Ishaq's translation of Euclid's Elements, also revised the trans- lation of the Data by the same Ishaq and the book On divisions of figures translated by an anonymous writer. We are told also that he wrote the following works : (1) On the Premisses (Axioms, Postulates etc.) of Euclid, (2) On the Propositions of Euclid, (3) On the propositions and questions which arise when two straight lines are cut by a third (or on the "proof" of Euclid's famous postulate). The last tract is extant in the MS. discovered by Woepcke (Paris 2 457> 3*°)- He is also credited with "an excellent work" in the shape of an " Introduction to the Book of Euclid," a treatise on 1 Fihriit, pp. 16, 1 j. * Woepcke, Mtm. pris. i I'acad. da scittua, XIV. pp. 666, 668. * Suter, p. 68. * FiAmt, p. 38 ; Suter. p. 60. 1 Fikrist, p. 16, rad Suter' 5 Dote, p. 60. * Suter, p. tit, note 13, ' Fikritt, p. 43, * Fihriit, p. +3 ; Soter, p. 41. 88 INTRODUCTION [ch. vii Geometry dedicated to Ismail b, Bulbul, a Compendium of Geometry, and a large number of other works for the titles of which reference may be made to Suter, who also gives particulars as to which are extant 1 . 17. Abu Sa'ld Sinan b. T habit b, Qurra, the son of the translator of Euclid, followed in his father's footsteps as geometer, astronomer and physician. He wrote an "improvement of the book of on the Elements of Geometry, in which he made various additions to the original." It is natural to conjecture that Euclid is the name missing in this description (by Ibn abl Usaibi'a); Casiri has the name Aqaton 1 . The latest editor of the Ta'rikk al-Hukamd, however, makes the name to be Iflaton (= Plato), and he refers to the statement by the Fikrist and Ibn al-Qiftl attributing to Plato a work on the Elements of Geometry translated by Qusta. It is just possible, therefore, that at the time of Qusta the Arabs were acquainted with a book on the Elements of Geometry translated from the Greek, which they attri- buted to Plato*. Sinan died in 94.3. 18. Abu Sahl Wijan (or Waijan) b. Rustam al-Kuhi (ft 988), born at Kuh in Tabaristan, a distinguished geometer and astronomer, wrote, according to the Fikrist, a " Book of the Elements" after that of Euclid*; the 1st and 2nd Books survive at Cairo, and a part of the 3rd Book at Berlin (5922)'. He wrote also a number of other geometrical works : Additions to the 2nd Book of Archimedes on the Sphere and Cylinder (extant at Paris, at Leiden, and in the India Office), On the finding of the side of a heptagon in a circle (India Office and Cairo), On two mean proportionals (India Office), which last may be only a part of the Additions to Archimedes' On the Sphere and Cylinder, etc. 19. Abu Nasr Muh. b. Muh. b. Tarkhan b. Uzlag al-Farabl (870-950) wrote a commentary on the difficulties of the introductory matter to Books I. and V. s This appears £0 survive in the Hebrew translation which is, with probability, attributed to Moses b. Tibbon'. 20. Abu 'All al-Hasan b. al- Hasan b, al-Haitham (about 965- 1039), known by the name Ibn al-Haitham or Abu 'AHal-Basri, was a man of great powers and knowledge, and no one of his time approached him in the field of mathematical science. He wrote several works on Euclid the titles of which, as translated by Woepcke from Usaibi'a, are as follows 5 : 1. Commentary and abridgment of the Etements. 2. Collection of the Elements of Geometry and Arithmetic, drawn from the treatises of Euclid and Apollonius. 3. Collection of the Elements of the Calculus deduced from the principles laid down by Euclid in his Elements. 1 Suter, pp. 34—8. 1 Fikritt (ed. Suter), p. 59, note 131 ; Suter, p. ji, note b. * See Suter in Bibliathtca Mathematical Jv a> 1903-4, pp. 396 — 7, review of JuIluk Lippert's Ihn al-Qifti. Ta^rith al-hiikamd, Leipzig, 1903. ■ Fikritt, p. 40. • Suter, p. - IE> EK TON 6EON02 ZYNOYSICN. Et? tov avrov to wp&TOV, ifyyriftdT&v UpoicXov /9*/9\. 5. Adiecta praefatiuncula in qua de disciplinis Mathematicis nonnihil. BASILEAE APVD 10 AN. HERVAGIVM ANNO M.D.XXXIII. MENSE SEPTEMBRI. The editor was Simon Grynaeus the elder (d. 1541), who, after working at Vienna and Ofen, Heidelberg and Tubingen, taught last of all at Basel, where theology was his main subject. His "prae- fatiuncula " is addressed to an Englishman, Cuthbert Tonstall { 1474- ■559)i who, having studied first at Oxford, then at Cambridge, where he became Doctor of Laws, and afterwards at Padua, where in addi- tion he leamt mathematics — mostly from the works of Regiomontanus and Paciuolo— wrote a book on arithmetic 1 as "a farewell to the sciences," and then, entering politics, became Bishop of London and member of the Privy Council, and afterwards (1530) Bishop of Durham. Grynaeus tells us that he used two MSS. of the text of the Elements, entrusted to friends of his, one at Venice by "Lazarus Bayfius" (Lazare de BaJf, then the ambassador of the King of France at Venice), the other at Paris by " loann. Rvellius " (J ean Ruel, a French doctor and a Greek scholar), while the commentaries of Proclus were put at 1 Dt arte supputandi libri quatuar. ch. viu] TRANSLATIONS AND EDITIONS 101 the disposal of Grynaeus himself by " loann. Claymundus" at Oxford, Heiberg has been able to identify the two mss. used for the text ; they are (i) cod. Venetus Marcianus 301 and (2) cod. Paris, gr. 2343 of the 16th c, containing Books 1. — xv., with some scholia which are embodied in the text. When Grynaeus notes in the margin the readings from " the other copy," this " other copy " is as a rule the Paris MS., though sometimes the reading of the Paris MS. is taken into the text and the " other copy " of the margin is the Venice MS. Besides these two mss. Grynaeus consulted Zamberti, as is shown by a number of marginal notes referring to " Zampertus " or to " latin um exemplar" in certain propositions of Books IX. — XL When it is con- sidered that the two MSS. used by Grynaeus are among the worst, it is obvious how entirely unauthoritative is the text of the editio princeps. Yet it remained the source and foundation of later editions of the Greek text for a long period, the editions which followed being designed, not for the purpose of giving, from other MSS., a text more nearly representing what Euclid himself wrote, but of supplying a handy compendium to students at a moderate price. 1536. Orontius Finaeus (Oronce Fine) published at Paris {apurf Simonem Colinaeum) "demonstrations on the first six books of Euclid's elements of geometry," " in which the Greek text of Euclid himself is inserted in its proper places, with the Latin translation of Barth. Zamberti of Venice," which seems to imply that only the enunciations were given in Greek. The preface, from which Kastner quotes', says that the University of Paris at that time required, from all who aspired to the laurels of philosophy, a most solemn oath that they had attended lectures on the said first six Books. Other editions of Fine's work followed in 1544 and 1551. 1545. The enunciations of the fifteen Books were published in Greek, with an Italian translation by Angelo Caiani, at Rome (apud Antonium Bladum Asulanum). The translator claims to have cor- rected the books and " purged them of six hundred things which did not seem to savour of the almost divine genius and the perspicuity of Euclid'" 1549. Joachim Camerarius published the enunciations of the first six Books in Greek and Latin (Leipzig). The book, with preface, purports to be brought out by Rhaeticus (1514-1576), a pupil of Copernicus. Another edition with proofs of the propositions of the first three Books was published by Moritz Steinmetz in 1 577 (Leipzig) ; a note by the printer attributes the preface to Camerarius himself. 155a loan. Scheubel published at Basel (also per loan. Her- vagium) the first six Books in Greek and Latin "together with true and appropriate proofs of the propositions, without the use of letters " (i.e. letters denoting points in the figures), the various straight lines and angles being described in words 1 . 1557 (also 1558). Stephanus Gracilis published another edition (repeated 1 573, 1 578, 1 598) of the enunciations (alone) of Books I, — XV. 1 Kastner, I. p. j6o. * Heiberg, vol. v. p. crii. ' Kastner, 1. p. 359. 10a INTRODUCTION [ch. vm in Greek and Latin at Paris {apud Gulielmurn Cavellai). He remarks in the preface that for want of time he had changed scarcely anything in Books I. — VI., but ; n the remaining Books he had emended what seemed obscure or inelegant in the Latin translation, while he had adopted in its entirety the translation of Book x. by Pierre Mondore (PetrusMontaureus), published separately at Paris in 1551. Gracilis also added a few " scholia." 1564. In this year Conrad Dasypodius (Rauchfuss), the inventor and maker of the clock in Strassburg cathedral, similar to the present one, which did duty from 1571 to 1789, edited (Strassburg, Chr. Mylius) (1) Book 1. of the Elements in Greek and Latin with scholia, (2) Book 11. in Greek and Latin with Barlaam's arithmetical version of Book II., and (3) the enunciations of the remaining Books III. — XIII. Book 1. was reissued with " vocabula quaedam geometrica " of Heron, the enunciations of all the Books of the Elements, and the other works of Euclid, all in Greek and Latin. In the preface to (1) he says that it had been for twenty-six years the rule of his school that all who were promoted from the classes to public lectures should learn the first Book, and that he brought it out, because there were then no longer any copies to be had, and in order to prevent a good and fruitful regulation of his school from falling through. In the preface to the edition of 157 1 he says that the first Book was generally taught in all gymnasia and that it was prescribed in his school for the first class. In the preface to (3) he tells us that he published the enunciations of Books in.— xill. in order not to leave his work unfinished, but that, as it would be irksome to carry about the whole work of Euclid in extenso, he thought it would be more convenient to students of geometry to learn the Elements if they were compressed into a smaller book. 1620. Henry Briggs {of Briggs' logarithms) published the first six Books in Greek with a Latin translation after Commandinus, "corrected in many places" (London, G. Jones). 1703 is the date of the Oxford edition by David Gregory which, until the issue of Heiberg and Menge's edition, was still the only edition of the complete works of Euclid 1 . In the Latin translation attached to the Greek text Gregory says that he followed Comman- dinus in the main, but corrected numberless passages in it by means of the books in the Bodleian Library which belonged to Edward Bernard (1638- 1 696), formerly Savilian Professor of Astronomy, who had conceived the plan of publishing the complete works of the ancient mathematicians in fourteen volumes, of which the first was to contain Euclid's Elements I. — xv. As regards the Greek text, Gregory tells us that he consulted, as far as was necessary, not a few MSS, of the better sort, bequeathed by the great Savile to the University, as well as the corrections made by Savile in his own hand in the margin of the Basel edition. He had the help of John Hudson, Bodley's Librarian, who 1 ETKAEIAOT TA SOZOMBNA. Euclidis qnae supersunl omnia. Ex recensions Davidis Gregorii M.D. Asttonomiae Professoris Saviliani el R.S.S. Oxoniae, o Theatio Sheldoniano, An. Dora, mdccici. cm. vin] TRANSLATIONS AND EDITIONS 103 punctuated the Basel text before it went to the printer, compared the Latin version with the Greek throughout, especially in the Elements and Data, and, where they differed or inhere he suspected the Greek text, consulted the Greek msS. and put their readings in the margin if they agreed with the Latin and, if they did not agree, affixed an asterisk in order that Gregory might judge which reading was geo- metrically preferable. Hence it is clear that no Greek MS., but the Basel edition, was the foundation of Gregory's text, and that Greek MSS. were only referred to in the special passages to which Hudson called attention. 1 is 14-1 818. A most important step towards a good Greek text was taken by F. Peyrard, who published at Paris, between these years, in three volumes, the Elements and Data in Greek, Latin and French 1 . At the time (1808) when Napoleon was having valuable MSS. selected from Italian libraries and sent to Paris, Peyrard managed to get two ancient Vatican MSS. (190 and 1038) sent to Paris for his use (Vat. 304 was a ' so at Paris at the time, but all three were restored to their owners in 1 8 14). Peyrard noticed the excellence of Cod. Vat. 190, adopted many of its readings, and gave in an appendix a conspectus of these readings and those of Gregory's edition ; he also noted here and there readings from Vat. 1038 and various Paris MSS. He there- fore pointed the way towards a better text, but committed the error of correcting the Basel text instead of rejecting it altogether and starting afresh. 1824-1825. A most valuable edition of Books I. — VI. is that of J. G. Camerer (and C. F, Hauber) in two volumes published at Berlin*. The Greek text is based on Peyrard, although the Basel and Oxford editions were also used. There is a Latin translation and a collection of notes far more complete than any other I have seen and well nigh inexhaustible. There is no editor or commentator of any mark who is not quoted from ; to show the variety of important authorities drawn upon by Camerer, I need only mention the following names : Proclus, Pappus, Tartaglia, Command inus, Clavius, Peletier, Barrow, Borelli, Watlis, Tacquet, Austin, Simson, Playfair. No words of praise would be too warm for this veritable encyclopaedia of information. 1825. J. G. C. Neide edited, from Peyrard, the text of Books I. — VI., XI. and XII. (Halts Saxoniae). 1826-9. The last edition of the Greek text before Heiberg's is that of E. F. August, who followed the Vatican MS. more closely than Peyrard did, and consulted at all events the Viennese MS. Gr, 103 (Heiberg's V). August's edition (Berlin, 1826-9) contains Books I. — XIII. 1 Eudidis quae supersuni. Las tEuvres a*Eudide t m Gree, en Latin ei en Fronfais d'aprls an msnuscrit tris-atvitn, qui Itait rerte* iwonnu jusqn'a nos jeurt. Par F, Peyrard. Ouvrijje approuv* par I'lnttitut de Franc* (Paris, chei M. Patris). 1 Euclidis elemettforum hbri lex prwrrs graice et Inline tommentarie e scriptis veierum ae rteentforum mathtmatxarum ei Pjltiaertri maxime illustraii [Berolnii, suEnptibus G. Rcimeri). Tom. I. 1814 ; torn. II. t8i{. io4 INTRODUCTION [ch. viii III. Latin versions or commentaries after 1533. 1545. Petrus Ramus (Pierre de la Ramee, 151 5-1572) is credited with a translation of Euclid which appeared in 1545 and again in 1549 at Paris 1 . Ramus, who was more rhetorician and logician than geometer, also published in his Scholae mathematical ( iSS9i Frankfurt; 1569, Base!) what amounts to a series of lectures on Euclid's Elements, in which he criticises Euclid's arrangement of his propositions, the definitions, postulates and axioms, all from the point of view of logic. 1557. Demonstrations to the geometrical Elements of Euclid, six Books, by Peletarius (Jacques Peletier). The second edition (1610) contained the same with the addition of the "Greek text of Euclid"; but only the enunciations of the propositions, as well as the defini- tions etc., are given in Greek (with a Latin translation), the rest is in Latin only. He has some acute observations, for instance about the "angle" of contact •559- Johannes Buteo, or Borrel (1492-1573), published in an appendix to his book De quadratura circuit some notes " on the errors of Campanus, Zambertus, Orontius, Peletarius, Pena, interpreters of Euclid." Buteo in these notes proved, by reasoned argument based on original authorities, that Euclid himself and not Theon was the author of the proofs of the propositions. 1 566. Franciscus Flussates Candalla (Francois de Foix, Comte de Candale, 1 502-1 594) "restored" the fifteen Books, following, as he says, the terminology of Zamberti's translation from the Greek, but drawing, for his proofs, on both Campanus and Theon (i.e. Zamberti) except where mistakes in them made emendation necessary. Other editions followed in 1578, 1602, 1695 (in Dutch). 1572. The most important Latin translation is that of Com- mand irtus (1509-1575) of Urbino, since it was the foundation of most translations which followed it up to the time of Peyrard, including that of Simson and therefore of those editions, numerous in England, which give Euclid "chiefly after the text of Simson." Simson 's first (Latin) edition (1756) has "ex versione Latina Federici Commandini" on the title-page. Commandinus not only followed the original Greek more closely than his predecessors but added to his translation some ancient scholia as well as good notes of his own. The title of his work is Euclidis elementorum libri xv, una cum sckeliis antiquis. A Federico Contmandino Urbinate nuper in latinum conversi, commentariisque quibusdam illustrati (Pisauri, apud Camillum Francischinum). He remarks in his preface that Orontius Finaeus had only edited six Books without reference to any Greek ms., that Peletarius had followed Campanus' version from the Arabic rather than the Greek text, and that Candalla had diverged too far from Euclid, having rejected as inelegant the proofs given in the Greek text and substituted faulty proofs of his own. Commandinus appears to have 1 Described by Boncompagni, BuiUitino, it. p. 3S9. ch. vm] TRANSLATIONS AND EDITIONS 105 used, in addition to the Basel editio princeps, some Greek ms., so far not identified ; he also extracted his " scholia antiqua " from a US. of the class of Vat. 192 containing the scholia distinguished by Heiberg as " Schol. Vat." New editions of Commandinus' translation followed in 1575 (in Italian), 1619, 1749 (in English, by Keill and Stone), 1756 (Books 1. — VI., XI., XII. in Latin and English, by Simson), 1763 (Keill). Besides these there were many editions of parts of the whole work, e.g. the first six Books. [574. The first edition of the Latin version by Clavius' (Christoph Klau [?], born at Bamberg 1537, died 1612) appeared in 1574, and new editions of it in 1589, 1591, 1603, 1607, 1612. It is not a translation, as Clavius himself states in the preface, but it contains a vast amount of notes collected from previous commentators and editors, as well as some good criticisms and elucidations of his own. Among other things, Clavius finally disposed of the error by which Euclid had been identified with Euclid of Megara. He speaks of the differences between Campanus who followed the Arabic tradition and the " commentaries of Theon," by which he appears to mean the Euclidean proofs as handed down by Theon ; he complains of predecessors who have either only given the first six Books, or have rejected the ancient proofs and substituted worse proofs of their own, but makes an exception as regards Commandinus, " a geometer not of the common sort, who has lately restored Euclid, in a Latin translation, to his original brilliancy." Clavius, as already stated, did not give a translation of the Elements but rewrote the proofs, com- pressing them or adding to them, where he thought that he could make them clearer. Altogether his book is a most useful work. 1 62 1. Henry Savile's lectures {Praelectiones tresdecttn in prin- cipium EUtnentorum Etulidis Oxoniae habitae MDC.XX., Oxonii 1621), though they do not extend beyond I. 8, are valuable because they grapple with the difficulties connected with the preliminary matter, the definitions etc., and the tacit assumptions contained in the first propositions, 1654 Andre 1 Tacquet's Elementa geometriae planae et solidae containing apparently the eight geometrical Books arranged for general use in schools. It came out in a large number of editions up to the end of the eighteenth century. 1655, Barrow's Euclidis Eiementorum Libri XV breviter demon- strati is a book of the same kind. In the preface (to the edition of [659) he says that he would not have written it but for the fact that Tacquet gave only eight Books of Euclid. He compressed the work into a very small compass (less than 400 small pages, in the edition of 1659, for the whole of the fifteen Books and the Data) by abbre- viating the proofs and using a large quantity of symbols (which, he says, are generally Oughtred's). There were several editions up to 1732 (those of 1660 and 1732 and one or two others are in English). 1 EudiMs clem/ntor um libri XV. Aictuit xvi. de selidorum regaiarium comparatism. Omnes pefipituis dtmonstratiimibm, accurattsqut nhotiii iltustrati. Auclorc ChristophoTO Ctaaio (Romae, apud Vinccotium Accoltum), 1 vols. to6 INTRODUCTION [ch. vm 1658. Giovanni Alfonso Borelli (1608- 1 679) published Euclides restitutus, on apparently similar lines, which went through three more editions (one in Italian, 1663). 166a Claude Francois Milliet Dechales' eight geometrical Books of Euclid's Elements made easy. Dechales' versions of the Elements had great vogue, appearing in French, Italian and English as well as Latin. Riccardi enumerates over twenty editions. 1733. Saccheri's Euclides ab omni naevo vindicatus sive const us geometrkus quo stabiliuntur prima ipsa geomelriae principia is important for his elaborate attempt to grove the parallel-postulate, forming an important stage in the history of the development of non- Euclidean geometry. 1756. Simson's first edition, in Latin and in English. The Latin title is Euclidis elementorum libri priores sex, item undecimus et duo- decimus, ex versione latina Federici Commandtni; sublatis iis quibus olim libri hi a Tkeone, aliisve, vitiali sunt, et quibusdam Euclidis demonstrationibus restitutis. A Roberto Simson M.D. Glasguae, in aedibus Academicis excudebant Robertus et Andreas Foulis, Academiae typographi. 1802. Euclidis elementorum libri priores XII ex Commandini et Gregorii versianibus latinis. In usum juventutis Academicae..,hy Samuel Horsley, Bishop of Rochester. (Oxford, Clarendon Press.) IV. Italian versions or commentaries. 1543. Tartaglia's version, a second edition of which was pub- lished in 1565', and a third in 1585. It does not appear that he used any Greek text, for in the edition of 1565 he mentions as available only "the first translation by Campano," "the second made by Bartolomeo Zamberto Veneto who is still alive," "the editions of Paris or Germany in which they have included both the aforesaid translations," and "our own translation into the vulgar (tongue)." IS7S- Commandinus' translation turned into Italian and revised by him. 1613. The first six Books "reduced to practice" by Pietro Antonio Cataldi, re-issued in 1620, and followed by Books VH, — ix. (1621) and Book X. (1625). 1663. Borelli's Latin translation turned into Italian by Domenico Magni. 1680. Euclide restitute by Vitale Giordano. 169a Vincenzo Vivian i's Eletnenti piani e solid t di Euclide (Book v. in 1674). 1 The title-page of the edition of 1 56; is ss follows : Euclide Megarcnse phihsophv. sail introduttorc deile identic msthematicc, diligtnttmente rasiettatc, et alia inicgrita ridotls, per it degno pr&fessore di tat identic Nicola Tartatca Briscians, scconde U due tratioltioni. con una ampla espositione delle istcsso tradottore di nuouo aggiutita. ialmente ckiara, cat ogni mediocre ingigno, stitta la notiiia, aver tvffragio di alcutT a/tra icientia con facUita jrrd capacc a peter la inttnderi. In Venetia, Appresso Curtio Troiano, 1565. ch. vm] TRANSLATIONS AND EDITIONS 107 173 1. Elementi geomttrici ptant e solid 'i di Eu elide by Guido Grandi. No translation, but an abbreviated version, of which new editions followed one another up to 1806. 1749. Italian translation of" Dechales with Ozanam's corrections and additions, re-issued 1785, 1797. 1752. Leonardo Ximenes (the first six Books). Fifth edition, 1 819. 181 8. Vincenzo Flauti's Corso di geometries eUtnentare e sublime (4 vols.) contains (Vol. I.) the first six Books, with additions and a dissertation on Postulate 5, and (Vol. II.) Books xi„ XH. Flauti also published the first six Books in 1827 and the Elements of geometry of Euclid m 1843 and 1854 V. German. 1558. The arithmetical Books vir. — ix. by Scheubel" (cf. the edition of the first six Books, with enunciations in Greek and Latin, mentioned above, under date 1 5 50). 1562. The version of the first six Books by Wilhelm Holtzmann (Xylander)*. This work has its interest as the first edition in German, but otherwise it is not of importance. Xylander tells us that it was written for practical people such as artists, goldsmiths, builders etc., and that, as the simple amateur is of course content to know facts, without knowing how to prove them, he has often left out the proofs altogether. He has indeed taken the greatest possible liberties with Euclid, and has not grappled with any of the theoretical difficulties, such as that of the theory of parallels. 1651. Heinrich Hoffmann's Teutscher Euclides (2nd edition 1653), not a translation. 1694. Ant Ernst Burkh. v. Pirckenstein's Teutsch Redender Euclides (eight geometrical Books), "for generals, engineers etc." "proved in a new and quite easy manner." Other editions 1699, 1744. 1697. Samuel Reyher's In teutscher Sfiraclie vorgestellter Euclides (six Books), "made easy, with symbols algebraical or derived from the newest art of solution." 1714. Euclidu xv Bilcker teutsch, "treated in a special and brief manner, yet completely," by Chr. Schessler (another edition in 1729). 1773. The first six Books translated from the Greek for the use of schools by J. F, Lorenz. The first attempt to reproduce Euclid in German word for word. 1 78 1. Books XI., xji. by Lorenz (supplementary to the pre- ceding). Also Euklid's Etemente fUnfzekn Biicker translated from 1 Das libcndochl und taunt buck da kochherumbun Mathematici Euclidis Mtgartnsis... durck Magistrum Jahaun Sckeybf, der loblichen univcrsittt zu Tubingen, des EucHdis und Arithmetic Ordinarien t uuss dem lutein ins teutsch gebra£kt * Di* seeks erste Hucher Euclidis vom an fang oder grund der Geometry. ..Asess Grserhtscker sfiraek in die Teutsch gebraekt aigentlick erkldrt. . . Demassen vcr/uals in Teutscher spraek nie imhen warden... Durch Wilktlm Holttman gtnattt Xylander von Augsfurg. Getnickht m io8 INTRODUCTION [ch. viii the Greek by Lorenz (second edition 1798; editions of 1809, 1818, 1824 by Mollweide, of 1840 by Dippe). The edition of 1824, and I presume those before it, are shortened by the use of symbols and the compression of the enunciation and "setting-out" into one. 1807. Books 1.— vi., xi. ( xil. "newly translated from the Greek," by J. K. F. Hauff. 1828. The same Books by Joh. Jos. Ign. Hoffmann "as guide to instruction in elementary geometry," followed in 1832 by observa- tions on the text by the same editor. 1833. Die Geometric des Euklid und das Wesen derselben by E. S. Unger; also 1838, 1851. 1901. Max Simon, Euclid und die seeks planimetrischen Bucket. VI, French. 1564-1566. Nine Books translated by Pierre Forcadel, a pupil and friend of P. de la Ramee. 1604. The first nine Books translated and annotated by Jean Errard de Bar-Ie-Duc ; second edition, 1605. 161 5. Denis Hen r ion's translation of the 15 Books (seven editions up to 1676), 1639. The first six Books "demonstrated by symbols, by a method very brief and intelligible," by Pierre Hengone, mentioned by Barrow as the only editor who, before him, had used symbols for the exposition of Euclid. 1672. Eight Books "rend us plus faciles" by Claude Francois Mi Diet Dec hales, who also brought out Les ilemens d'Euclide ex- pliquis d'une maniere nouvelie et trh facile, which appeared in many editions, 1672, 1677, 1683 etc. (from 1709 onwards revised by Ozanam), and was translated into Italian (1749 etc.) and English (by William Halifax, 1685). 1 804. In this year, and therefore before his edition of the Greek text, F. Peyrard published the Elements literally translated into French. A second edition appeared in 1 809 with the addition of the fifth Book, As this second edition contains Books 1. — vi. XL, xn. and x. 1, it would appear that the first edition contained Books 1. — iv., VI., XI., XII. Peyrard used for this translation the Oxford Greek text and Simson. VII. Dutch. 1606. Jan Pieterszoon Dou (six Books). There were many later editions. Kastner, in mentioning one of 1702, says that Dou explains in his preface that he used Xylander's translation, but, having after- wards obtained the French translation of the six Books by Errard de Bar-le-Duc (see above),' the proofs in which sometimes pleased him more than those of the German edition, he made his Dutch version by the help of both. 1 61 7. Frans van Schooten, "The Propositions of the Books of Euclid's Elements"; the fifteen Books in this version " enlarged " by Jakob van Leest in 1662. 1695. C. J. Vooght, fifteen Books complete, with Candalla's " 16th." ch. vin] TRANSLATIONS AND EDITIONS 109 1702. Kendfik Coets, six Books (also in Latin, 1692); several editions up to 1752. Apparently not a translation: but an edition for school use. 1763. Pybo Steenstra, Books I. — vr„ XI., XII., likewise an abbre- viated version, several times reissued until 1825. VIII. English. 1570 saw the first and the most important translation, that of Sir Henry Billingsley. The title-page is as follows : THE ELEMENTS OF GEOMETRIE of the most auncient Philosopher E VCLWE of Megara Faithfully (now first) translated into the English toung, by H. Billingsley, Citizen of London. Whereunto are annexed certaine Scholies, Annotations, and Inuentions, of the best Mathematiciens, both of time past, and in this our age. With a very fruitfull Preface by M. I. Dee, specifying the chiefe Mathetnaticall Sciiees, what they are, and whereunto commodious: where, also, are disclosed certaine new Secrets Mathetnaticall and Afechanicall, vntill these our dales, greatly missed. Imprinted at London by John Daye. The Preface by the translator, after a sentence observing that with- out the diligent study of Euclides Efementes it is impossible to attain unto the perfect knowledge of Geometry, proceeds thus. " Wherefore considering the want and lacke of such good authors hitherto in our Englishe tounge, lamenting also the negligence, and lacke of zeale to their countrey in those of our nation, to whom God hath geuen both knowledge and also abilitie to translate into our tounge, and to publishe abroad such good authors and bookes (the chiefe instrumentes of all learninges): seing moreouer that many good wittes both of gentlemen and of others of all degrees, much desirous and studious of these artes, and seeking for them as much as they can, sparing no paines, and yet frustrate of their intent, by no meanes attaining to that which they seeke : I haue for their sakes, with some charge and great trauaile, faithfully translated into our vulgar e touge, and set abroad in Print, this booke of Euclide. Whereunto I haue added easie and plaine declarations and examples by figures, of the defini- tions. In which booke also ye shall in due place finde manifolde additions, Scholies, Annotations, and Inuentions; which I haue gathered out of many of the most famous and chiefe Mathematicies, both of old time, and in our age : as by diligent reading it in course, ye shall well perceaue...," It is truly a monumental work, consisting of 464 leaves, and there- fore 928 pages, of folio size, excluding the lengthy preface by Dee. The notes certainly include all the most important that had ever been tto INTRODUCTION [cm. vin written, from those of the Greek commentators, Proclus and the others whom he quotes, down to those of Dee himself on the la3t books. Besides the fifteen Books, Billingsley included the "sixteenth" added by Candalla, The print and appearance of the book are worthy of its contents ; and, in order that it may be understood how no pains were spared to represent everything in the clearest and most perfect form, I need only mention that the figures of the propositions in Book XI. are nearly all duplicated, one being the figure of Euclid, the other an arrangement of pieces of paper (triangular, rectangular etc.) pasted at the edges on to the page of the book so that the pieces can be turned up and made to show the real form of the solid figures represented. Billingsley was admitted Lady Margaret Scholar of St John's College, Cambridge, in 1551, and he is also said to have studied at Oxford, but he did not take a degree at either University. He was afterwards apprenticed to a London haberdasher and rapidly became a wealthy merchant. Sheriff of London in 1584, he was elected Lord Mayor on 31st December, 1596, on the death, during his year of office, of Sir Thomas Skinner. From 1589 he was one of the Queen's four " customers," or farmers of customs, of the port of London. In 1 591 he founded three scholarships at St John's College for poor students, and gave to the College for their maintenance two messuages and tenements in Tower Street and in Mark Lane, Allhallows, Barking. He died in 1606. 1651. Elements of Geometry. The first VI Bootks: In a compen- dious form contracted and demonstrated by Captain Thomas Rudd, with the mathematicall preface of John Dee (London). 1660. The first English edition of Barrow's Euclid (published in Latin in 1655)1 appeared in London. It contained "the whole fifteen books compendiously demonstrated"; several editions followed, in 1705, 1722, 1732, 1751. 1 66 1. Euclid 's Elements of Geometry, with a supplement of divers Propositions and Corollaries. To which is added a Treatise of regular Soli lis by Campane and Ftussat ; likewise Euclid's Data and Marinus his Preface. Also a Treatise of the Divisions of Superficies, ascribed to Machomet Bagdedine, but published by Commandine at the request of J, Dee of London, Published by care and industry of John Leeke and Geo. Serle, students in the Math. (London). According to Potts this was a second edition of Billingsley's translation. 1685. William Halifax's version of Dechales' " Elements of Euclid explained in a new but most easy method " (London and Oxford). 1705. The English Euclide; being tlie first six Elements of Geometry, translated out of the Greek, with annotations and useful/ supplements by Edmund Scarburgh (Ox ford ). A noteworthy and useful edition. 1708. Books I. — VL, XL, xii., translated from Command in us' Latin version by Dr John Keill, Savilian Professor of Astronomy at Oxford. Keill complains in his preface of the omissions by such editors as Tacquet and Dechales of many necessary propositions (e.g. VI. 27 — 29), and of their substitution of proofs of their own for Euclid's. He praises Barrow's version on the whole, though objecting to the " algebraical " ca. viii] TRANSLATIONS AND EDITIONS in form of proof adopted in Book n., and to the excessive use of notes and symbols, which (he considers) make the proofs too short and thereby obscure: his edition was therefore intended to hit a proper mean between Barrow's excessive brevity and Clavius' prolixity. Keill's translation was revised by Samuel Cunn and several times reissued. 1749 saw the eighth edition, 1772 the eleventh, and 1782 the twelfth. 1714. W. Whiston's English version (abridged) of The Elements of Euclid with select theorems out of Archimedes by the learned Andr. Tacquet. 1756. Simson's first English edition appeared in the same year as his Latin version under the title : The Elements of Euclid, vis. the first six Books together with the eleventh and twelfth. In this Edition the Errors by which Theon or others have long ago vitiated these Books are corrected and some of Euclid's Demonstrations are restored. By Robert Simson (Glasgow). As above stated, the Latin edition, by its title, purports to be " ex version© latina Federici Commandini," but to the Latin edition, as well as to the English editions, are appended Notes Critical and Geometrical ; containing an Account of those things in which this Edition differs from the Greek text; and the Reasons of the Alterations which have been made. As also Obser- vations on some of the Propositions. Simson says in the Preface to some editions (e.g. the tenth, of 1799) that "the translation is much amended by the friendly assistance of a learned gentleman." Simson's version and his notes are so well known as not to need any further description. The book went through some thirty suc- cessive editions. The first five appear to have been dated 1756, 1762, 1767, 1772 and 1775 respectively; the tenth 1799, the thirteenth 1806, the twenty-third 1830, the twenty-fourth 1834, the twenty-sixth 1844. The Data "in like manner corrected '' was added for the first time in the edition of 1762 (the first octavo edition). 1781, 1788. In these years respectively appeared the two volumes containing the complete translation of the whole thirteen Books by James Williamson, the last English translation which reproduced Euclid word for word. The title is The Elements of Euclid, with Dissertations intended to assist and encourage a critical examination of these Elements, as the most effectual means of establishing a j'uster taste upon mathematical subjects than that which at present prevails. By James Williamson. In the first volume (Oxford, 1781) he is described as " M.A. Fellow of Hertford College," and in the second (London, printed by T. Spilsbury, 1788) as "B.D." simply. Books v., VI. with the Con- clusion in the first volume are paged separately from the rest 1 78 1 . 4 n examination of the first six Books of Euclid's Elements, by William Austin (London). '79S- John Playfair's first edition, containing "the first six Books of Euclid with two Books on the Geometry of Solids." The book ii3 INTRODUCTION [ch. vm reached a fifth edition in 1819, an eighth in 1831, a ninth in 1836, and a tenth in 1846. 1826. Riccardi notes under this date Euclid's Elements of Geo- metry containing the whole twelve Books translated into English, from the edition of Peyrard, by George Phillips. The editor, who was President of Queens' College, Cambridge, 1857-1892, was born in 1804 and matriculated at Queens' in 1 826, so that he must have published the book as an undergraduate. 1828. A very valuable edition of the first six Books is that of Dionysius Lardner, with commentary and geometrical exercises, to which he added, in place of Books XI., XII., a Treatise on Solid Geometry mostly based on Legend re, Lardner compresses the pro- positions by combining the enunciation and the setting-out, and he gives a vast number of riders and additional propositions in smaller print The book had reached a ninth edition by 1846, and an eleventh by 1855. Among other things, Lardner gives an Appendix "on the theory of parallel lines," in which he gives a short history of the attempts to get over the difficulty of the parallel' postulate, down to that of Legendre. 1833. T. Perronet Thompson's Geometry without axioms, or the first Book of Euclid's Elements with alterations and notes ; and an intercalary book in which the straight line and plane are derived from properties of the sphere, with an appendix containing notices of methods proposed for getting over the difficulty in the twelfth axiom of Euclid. Thompson (1783-1869) was 7th wrangler 1802, midshipman 1803, Fellow of Queens* College, Cambridge, 1804, and afterwards general and politician. The book went through several editions, but, having been well translated into French by Van Tenac, is said to have received more recognition in France than at home. 1 845. Robert Potts' first edition (and one of the best) entitled : Euclid's Elements of Geometry chiefly from the text of Dr Simson with explanatory notes... to which is prefixed an introduction containing a brief outline of the History of Geometry. Designed for the use of the higher forms in Public Schools and students in the Universities (Cambridge University Press, and London, John W. Parker), to which was added (1847) An Appendix to the larger edition of Euclid's Elements of Geometry, containing additional notes on the Elements, a short tract on trans- versals, and hints for the solution of the problems etc. 1862. Todhunter's edition. The later English editions 1 will not attempt to enumerate; their name is legion and their object mostly that of adapting Euclid for school use, with all possible gradations of departure from his text and order. IX, Spanish. 1576. The first six Books translated into Spanish by Rodrigo Camorano. 1637. The first six Books translated, with notes, by L. Carduchi. 1689. Books 1.— vi, XI, XII, translated and explained by Jacob Knesa. ch. vni] TRANSLATIONS AND EDITIONS 113 X. Russian. 1739. Ivan Astaroff* (translation from Latin). 1 789. Pr. Suvoroff and Yos. Nikitin {translation from Greek). 1880. Vachtchenko-Zakhartchenko. {181 7, A translation into Polish by Jo. Czecha.) XI. Swedish. 1744. Mitten Stromer, the first six Books ; second edition 1748. The third edition (1753) contained Books XI.— XU. as well; new editions continued to appear till 1884, 1836. H. Falk, the first six Books. 1844, 1845, 1859. P. R. Brakenhjelm, Books I. — VI., XL, xn. 1850. F. A. A. Lundgren. 1850. H. A. Witt and M. E. Areskong, Books I.— VI., XI., XII. XII. Danish. 1745. Ernest Gottlieb Ziegenbalg. 1803. H. C. Linderup, Books I. — VI. XIII. Modern Greek. 1820. Benjamin of Lesbos, I should add a reference to certain editions which have appeared in recent years. A Danish translation (Euklid's Eletnenter oversat af Thyra Eibe) was completed in 1912 ; Books I. — II. were published (with an Intro- duction by Zeuthen) in 1897, Books in. — IV. in 1900, Books v. — VI. in 1904, Books VII. — XIII. in 19 1 2, The Italians, whose great services to elementary geometry are more than once emphasised in this work, have lately shown a note- worthy disposition to make the ipsissima verba of Euclid once more the object of study. Giovanni Vacca has edited the text of Book I. (// prima libro degli Elementi. Testo greco, versione italiana, intro- duzione e note, Firenze 1 916,) Federigo Enriques has begun the publication of a complete Italian translation (Gli Elementi d' Enclide e la critica antica e moderna); Books I. — IV. appeared in 1925 (Alberto Stock, Roma). An edition of Book I. by the present writer was published in 1918 {Euclid in Greek \ Book [., with Introduction and Notes, Camb. Univ. Press). CHAPTER IX. j I, ON THE NATURE OF ELEMENTS. It would not be easy to find a more lucid explanation of the terms element and elementary, and of the distinction between them, than is found in Prod us \ who is doubtless, here as so often, quoting from Geminus. There are, says Proclus, in the whole of geometry certain leading theorems, bearing to those which follow the relation of a principle, all- pervading, and furnishing proofs of many properties. Such theorems are called by the name of elements ; and their function may be compared to that of the letters of the alphabet in relation to language, letters being indeed called by the same name in Greek (orotyeta). The term elementary, on the other hand, has a wider application : it is applicable to things " which extend to greater multiplicity, and, though possessing simplicity and elegance, have no longer the same dignity as the elements, because their investigation is not of general use in the whole of the science, e.g. the proposition that in triangles the perpendiculars from the angles to the transverse sides meet in a point." " Again, the term element is used in two senses, as Menaechmus says. For that which is the means of obtaining is an element of that which is obtained, as the first proposition in Euclid is of the second, and the fourth of the fifth. In this sense many things may even be said to be elements of each other, for they are obtained from one another. Thus from the fact that the exterior angles of rectilineal figures are (together) equal to four right angles we deduce the number of right angles equal to the internal angles (taken together)*, and vice versa. Such an element is like a lemma. But the term element is otherwise used of that into which, being more simple, the composite is divided ; and in this sense we can no longer say that everything is an element of everything, but only that things which are more of the nature of principles are elements of those which stand to them in the relation of results, as postulates are elements of theorems. It is 1 Proclus, Coium, on Eucl. I., ed. Friedlein, pp. Jssqq. * t4 TrkTjSin rw irrit dfi0a(j tnur. If the te*t is right, we must apparently take it as "the number of the angles equal to right angles that there are inside/' i.e. that are made up by the internal angles. ch. ix. J i] ON THE NATURE OF ELEMENTS 115 according to this signification of the term element that the elements found in Euclid were compiled, being partly those of plane geometry, and partly those of stereometry. In like manner many writers have drawn up elementary treatises in arithmetic and astronomy. " Now it is difficult, in each science, both to select and arrange in due order the elements from which all the rest proceeds, and into which all the rest is resolved. And of those who have made the attempt some were able to put together more and some less ; some used shorter proofs, some extended their investigation to an indefinite length ; some avoided the method of reductio ad absurdum, some avoided proportion-, some contrived preliminary steps directed against those who reject the principles ; and, in a word, many different methods have been invented by various writers of elements. "It is essential that such a treatise should be rid of everything superfluous (for this is an obstacle to the acquisition of knowledge) ; it should select everything that embraces the subject and brings it to a point (for this is of supreme service to science) ; it must have great regard at once to clearness and conciseness (for their opposites trouble our understanding); it must aim at the embracing of theorems in general terms (for the piecemeal division of instruction into the more partial makes knowledge difficult to grasp). In all these ways Euclid's system of elements will be found to be superior to the rest ; for its utility avails towards the investigation of the primordial figures 1 , its clearness and organic perfection are secured by the progression from the more simple to the more complex and by the foundation of the investigation upon common notions, while generality of demonstration is secured by the progression through the theorems which are primary and of the nature of principles to the things sought. As for the things which seem to be wanting, they are partly to be discovered by the same methods, like the construction of the scalene and isosceles (triangle), partly alien to the character of a selection of elements as introducing hopeless and boundless complexity, like the subject of unordered irrationals which Apoilonius worked out at length", and partly developed from things handed down (in the elements) as causes, like the many species of angles and of lines. These things then have been omitted in Euclid, though they have received full discussion in other works ; but the knowledge of them is derived from the simple (elements)." Proclus, speaking apparently on his own behalf, in another place distinguishes two objects aimed at in Euclid's Elements. The first has reference to the matter of the investigation, and here, like a good Platonist, he takes the whole subject of geometry to be concerned with the "cosmic figures," the five regular solids, which in Book XHi. 1 rwt Apxix&r as in the following passages. " That, when equals are taken from equals, the remainders are equal is (a) common (principle) in the case of alt quantities, but mathematics takes a separate department (atroXa^ovo'a) and directs its investigation to some portion of its proper subject-matter, as e.g. lines or angles, numbers, or any of the other quantities*.' 1 "The common (principles), e.g. that one of two contradictories must be true, that equals taken from equals etc., and the like" " " With regard to the principles of demonstration, it is questionable whether they belong to one science or to several. By principles of demonstration I mean the common opinions from which all demonstration proceeds, e.g. that one of two contradictories must be true, and that it is impossible for the same thing to be and not be 4 ." Similarly "every demonstrative (science) investigates, with regard to some subject-matter, the essential attributes, starting from the common opinions'." We have then here, as Heiberg says, a sufficient explanation of Euclid's term for axioms, 1 Anal. pest. I. a, 71a 14—14. ' Mttaph. 1061 b 19—14. * Anal. post. \. n, 77 a 30. * Mttaph. 996 b 36 — 30. 5 Mttaph. 907 a so — 11. ch. ix. §3] FIRST PRINCIPLES lai viz. common notions (koivoI evvoiat), and there is no reason to suppose it to be a substitution for the original term due to the Stoics : cf. Proclus' remark that, according to Aristotle and the geometers, axiom and common notion are the same thing 1 . Aristotle discusses the indemonstrable character of the axioms in the Metaphysics. Since "all the demonstrative sciences use the axioms'," the question arises, to what science does their discussion belong*? The answer is that, like that of Being (oiJct-mi), it is the province of the (first) philosopher 1 . It is impossible that there should be demonstration of everything, as there would be an infinite series of demonstrations: if the axioms were the subject of a demonstrative science, there would have to be here too, as in other demonstrative sciences, a subject-genus, its attributes and corresponding axioms 1 \ thus there would be axioms behind axioms, and so on continually. The axiom is the most firmly established of all principles*. It is ignorance alone that could lead anyonetotryto prove the axioms' ; the supposed proof would be a.petilio principii'. If it is admitted that not every- thing can be proved, no one can point to any principle more truly indemonstrable'. If any one thought he could prove them, he could at once be refuted ; if he did not attempt to say anything, it would be ridiculous to argue with him i he would be no better than a vegetable 10 . The first condition of the possibility of any argument whatever is that words should signify something both to the speaker and to the hearer: without this there can be no reasoning with any one. And, if any one admits that words can mean anything to both hearer and speaker, he admits that something can be true without demon- stration. And so on". It was necessary to give some sketch of Aristotle's view of the first principles, if only in connexion with Proclus' account, which is as follows. As in the case of other sciences, so " the compiler of elements in geometry must give separately the principles of the science, and after that the conclusions from those principles, not giving any account of the principles but only of their consequences. No science proves its own principles, or even discourses about them : they are treated as self-evident. . .Thus the first essential was to dis- tinguish the principles from their consequences. Euclid carries out this plan practically in every book and, as a preliminary to the whole enquiry, sets out the common principles of this science. Then he divides the common principles themselves into hypotheses, postulates, and axioms. For all these are different from one another : an axiom, a postulate and a hypothesis are not the same thing, as the inspired Aristotle somewhere says. But, whenever that which is assumed and ranked as a principle is both known to the learner and convincing in itself, such a thing is an axiom, e.g. the statement that things which are equal to the same thing are also equal to one another. When, on 1 Proclus, p. 194, 8. ' Mtlaph. 997 a 10. ' ibid. 996 b 16, * ibid. 100s a 11 — b 11. ' Hid. 997 a 5 — 8. ■ ibid. 1005 b II — 17. ' ibid, loofia 5. s ibid. 1006 a 17. ibid. 1006a 10. I0 ibid. 1006a 11 — 15. 1} ibid. 1006a iSsqq. u» INTRODUCTION [cH.1x.j3 the other hand, the pupil has not the notion of what is told him which carries conviction in itself, but nevertheless lays it down and assents to its being assumed, such an assumption is a hypothesis. Thus we do not preconceive by virtue of a common notion, and without being taught, that the circle is such and such a figure, but, when we are told so, we assent without demonstration. When again what is asserted is both unknown and assumed even without the assent of the learner, then, he says, we call this a postulate, e.g. that all right angles are equal. This view of a postulate is clearly implied by those who have made a special and systematic attempt to show, with regard to one of the postulates, that it cannot be assented to by any one straight off. According then to the teaching of Aristotle, an axiom, a postulate and a hypothesis are thus distinguished 1 ." We observe, first, that Proclus in this passage confuses hypotheses and definitions, although Aristotle had made the distinction quite plain. The confusion may be due to his having in his mind a passage of Plato from which he evidently got the phrase about " not giving an account of" the principles. The passage is a : " I think you know that those who treat of geometries and calculations (arithmetic) and such things take for granted (inroQefievoi) odd and even, figures, angles of three kinds, and other things akin to these in each subject, implying that they know these things, and, though using them as hypotheses, do not even condescend to give any account of them either to themselves or to others, but begin from these things and then go through everything else in order, arriving ultimately, by recognised methods, at the conclusion which they started in search of." But the hypothesis is here the assumption, e.g, ' that there may he sttch a thing as length without breadth, henceforward called a line',' and so on, without any attempt to show that there is such a thing ; it is mentioned in connexion with the distinction between Plato's 'superior' and 'inferior' intellectual method, the former of which uses successive hypotheses as stepping-stones by which it mounts upwards to the idea of Good. We pass now to Proclus' account of the difference between postu- lates and axioms. He begins with the view of Geminus, according to which " they differ from one another in the same way as theorems are also distinguished from problems. For, as in theorems we propose to see and determine what follows on the premisses, while in problems we are told to find and do something, in like manner in the axioms such things are assumed as are manifest of themselves and easily apprehended by our untaught notions, while in the postulates we assume such things as are easy to find and effect (our understanding suffering no strain in their assumption), and we require no complication of machinery*."..." Both must have the characteristic of being simple 1 Proclos, pp. 7S, 10—77, *• ' Republic, VI, 510 c. Cf. Aristotle, jVk. Eth. 1151a 17. * H. Jackson, Journal of Philology, vol. x. p. 144- * Proclus, pp. 178, 1? — (79,8* In illustration Proclus contrasts the drawing of a straight Line or a circle with the drawing of a " single- turn spiral " or of an equilateral triangle, the ch. ix. §3] FIRST PRINCIPLES 123 and readily grasped, I mean both the postulate and the axiom ; but the postulate bids us contrive and find some subject-matter (uXj,) to exhibit a property simple and easily grasped, while the axiom bids us assert some essential attribute which is self-evident to the learner, just as is the fact that fire is hot, or any of the most obvious things 1 ," Again, says Proclus, " some claim that alt these things are alike postulates, in the same way as some maintain that all things that are sought are problems. For Archimedes begins his first book on /«- equilibrium' 1 with the remark ' I postulate that equal weights at equal distances are in equilibrium,' though one would rather call this an axiom. Others call them all axioms in the same way as some regard as theorems everything that requires demonstration'." " Others again will say that postulates are peculiar to geometrical subject-matter, while axioms are common to all investigation which is concerned with quantity and magnitude. Thus it is the geometer who knows that all right angles are equal and how to produce in a straight line any limited straight line, whereas it is a Common notion that things which are equal to the same thing are also equal to one another, and it is employed by the arithmetician and any scientific person who adapts the general statement to his own subject 1 ." The third view of the distinction between a postulate and an axiom is that of Aristotle above described'. The difficulties in the way of reconciling Euclid's classification of postulates and axioms with any one of the three alternative views are next dwelt upon. If we accept the first view according to which an axiom has reference to something known, and a postulate to something done, then the 4th postulate (that all right angles are equal) \% not a postulate ; neither is the 5th which states that, if a straight line falling on two straight lines makes the interior angles on the same side less than two right angles, the straight lines, if produced indefinitely, will meet on that side on which are the angles less than two right angles. On the second view, the assumption that two straight lines cannot enclose a space, " which even now," says Proclus, " some add as an axiom," and which is peculiar to the subject-matter of geometry, like the fact that all right angles are equal, is not an axiom. According to the third (Aristotelian) view, "everything which is confirmed (irtTat) by a sort of demonstration spiral requiring more complex machinery and even the equilateral triangle needing a certain method. " For the geometrical intelligence will say that hy conceiving a straight tine fined at one end but, as regards the other end, moving round the fixed end, and a point moving along the straight line from the fined end, 1 have described the single- turn spiral ; for the end of the straight line descrihing a circle, and the point moving on the straight line simul- taneously, when they arrive and meet at the same point, complete such a spiral. And again, if I draw equal circles, join their common point to the centres of the circles and draw a straight line from one of the centres to the other, I shall have the equilateral triangle. These things then are far from being completed by means of a single act or of a moment's thought" (p. 180. S— ji). 1 Proclus, p. i Si, 4 — 11. 1 It is necessary to coin a word to render tontapfxirtur, which is moreover in the plural. The title of the treatise as we have it is Equilibria of plana or centres of gravity of planes in Book r and Equilibria of planes in Book It. • Proclus, p. 181, 16—13. * **S p. 181, 6—14. ■ Pp. 118, 119. i»4 INTRODUCTION [ch. ix. §3 will be a postulate, and what is incapable of proof will be an axiom 1 ." This last statement of Proclus is loose, as regards the axiom, because it omits Aristotle's requirement that the axiom should be a self- evident truth, and one that must be admitted by any one who is to learn anything at all, and, as regards the postulate, because Aristotle calls a postulate something assumed without proof though it is "matter of demonstration" (aTroSetierhv Sv), but says nothing of a quasi -demon strati on of the postulates. On the whole I think it is from Aristotle that we get the best idea of what Euclid understood by a postulate and an axiom or common notion. Thus Aristotle's account of an axiom as a principle common to all sciences, which is self-evident, though incapable of proof, agrees sufficiently with the contents of Euclid's common notions as reduced to five in the most recent text (not omitting the fourth, that " things which coincide are equal to one another"). As regards the postulates, it must be borne in mind that Aristotle says elsewhere' that, "other things being equal, that proof is the better which proceeds from the fewer postulates or hypotheses or propositions." I f then we say that a geometer must lay down as principles, first certain axioms or common notions, and then an irreducible minimum of postulates in the Aristotelian sense concerned only with the subject-matter of geometry, we are not far from describing what Euclid in fact does. As regards the postulates we may imagine him saying : ■ Besides the common notions there are a few other things which I must assume without proof, but which differ from the common notions in that they are not self-evident. The learner may or may not be disposed to agree to them ; but he must accept them at the outset on the superior authority of his teacher, and must be left to convince himself of their truth in the course of the investigation which follows. In the first place certain simple constructions, the drawing and producing of a straight line, and the drawing of a circle, must be assumed to be possible, and with the constructions the existence of such things as straight lines and circles ; and besides this we must lay down some postulate to form the basis of the theory of parallels." It is true that the admission of the 4th postulate that all right angles are equal still presents a difficulty to which we shall have to recur. There is of course no foundation for the idea, which has found its way into many text-books, that " the object of the postulates is to declare that the only instruments the use of which is permitted in geometry are the rule and compass'." § 4. THEOREMS AND PROBLEMS. " Again the deductions from the first principles," says Proclus, "are divided into problems and theorems, the former embracing the 1 Proclus, pp. i8j, 11—183, '3- * Anal. past. 1. aj, 86* 33—35. 8 Cf. Lardner's Euclid : aha Todhunter. ch. ix. 5 4 ] THEOREMS AND PROBLEMS i*5 generation, division, subtraction or addition of figures, and generally the changes which are brought about in them, the latter exhibiting the essential attributes of each 1 ," " Now, of the ancients, some, like Speusippus and Amphinomus, thought proper to call them all theorems, regarding the name of theorems as more appropriate than that of problems to theoretic sciences, especially as these deal with eternal objects. For there is no becoming in things eternal, so that neither could the problem have any place with them, since it promises the generation and making of what has not before existed, e.g. the construction of an equilateral triangle, or the describing of a square on a given straight line, or the placing of a straight line at a given point. Hence they say it is better to assert that all (propositions) are of the same kind, and that we regard the generation that takes place in them as referring not to actual making but to knowledge, when we treat things existing eternally as if they were subject to becoming: in other words, we may say that everything is treated by way^ of theorem and not by way of problem* (irdvra 6eapr)fjiaTiKw oX\* ov TrpQJ3\iifiaTtic ri^Trzav. This should apparently be the fourth because in the next words it is implied that none of the first three propositions ate required in proving it. * Proclus, pp. 141, it) — 543, 11. * ibid. pp. 133, 11— 1134, 6. 1*8 INTRODUCTION [ch. ix. §4 Prop. 2, and must therefore precede it But Prop. 1 showing how to construct an equilateral triangle on a given base is not important, in relation to Prop. 4, as dealing with the " production of triangles " in general : for it is of no use to say, as Proclus does, that the construc- tion of the equilateral triangle is " common to the three species (of triangles) 1 ," as we are not in a position to know this at such an early stage. The existence of triangles in general was doubtless assumed as following from the existence of straight lines and points in one plane and from the possibility of drawing a straight line from one point to another. Proclus does not however seem to reject definitely the view of Carpus, for he goes on* : "And perhaps problems are in order before theorems, and especially for those who need to ascend from the arts which are concerned with things of sense to theoretical investigation. But in dignity theorems are prior to problems.... It is then foolish to blame Geminus for saying that the theorem is more perfect than the problem. For Carpus himself gave the priority to problems in respect of order, and Geminus to theorems in point of more perfect dignity," so that there was no real inconsistency between the two. Problems were classified according to the number of their possible solutions. Amphinomus said that those which had a unique solution (itowa^tSt) were called " ordered " (the word has dropped out in Proclus, but it must be rerayftiva, in contrast to the third kind, aTattra) ; those which had a definite number of solutions " inter- mediate" (fiea-a); and those with an infinite variety of solutions "un- ordered " (ajaMTa)'. Proclus gives as an example of the last the problem To divide a given straight line into three parts in continued proportion*. This is the same thing as solving the equations %+y+s=a, xs =_y i . Proclus' remarks upon the problem show that it was solved, like all quadratic equations, by the method of " application of areas." The straight line a was first divided into any two parts, (jr-M)and_y, subject to the sole limitation that (x + s) must not be less than 2y, which limitation is the Stopto-fws, or condition of possibility. Then an area was applied to (x + z), or («— y), "falling short by a square figure" {eWetirov tXhei rerpaywy) and equal to the square on y. This determines x and z separately in terms of a and y. For, if b be the side of the square by which the area (i.e. rectangle) "falls short," we have {{a —y) — z\z ™j**, whence 2z «= (a —y) ± n/[(a —yf — 4y 3 }. And y may be chosen arbitrarily, provided that it is not greater than 0/3. Hence there are an infinite number of solutions. If y = a(i, then, as Proclus remarks, the three parts are equal. Other distinctions between different kinds of problems are added by Proclus. The word " problem," he says, is used in several senses. In its widest sense it may mean anything " propounded " (irpojewo- pi-i>ov), whether for the purpose of instruction {paBrio-em) or construc- tion (■n-oojo'fci)?). (In this sense, therefore, it would include a theorem.) 1 Proclus, p. 13^, Ji. ■ ibid, p. 543, 11—15. ■ ibid, p. »jo> 7 — 13. * ibid. pp. no, 16 — 111, 6. ch. ix. §4] THEOREMS AND PROBLEMS 139 But its special sense in mathematics is that of something "propounded with a view to a theoretic construction 1 ." Again you may apply the term (in this restricted sense) even to something which is impossible, although it is more appropriately used of what is possible and neither asks too much nor contains too little in the shape of data. According as a problem has one or other of these defects respectively, it is called (1) a problem in excess (irXeovdgov) or (2) a deficient problem (iK\nre<; irp60\T}p.a), The problem in excess (1) is of two kinds, (a) a problem in which the properties of the figure to be found are either inconsistent (atrii p/Sara) or non-existent {avvirapicTa), in which case the problem is called impossible, or (b) a problem in which the enunciation is merely redundant : an example of this would be a problem requiring us to construct an equilateral triangle with its vertical angle equal to two-thirds of a right angle ; such a problem is possible and is called "more than a problem '' (p.el&v 7j Trp6ft\T)fta). The deficient problem (2) is similarly called " less than a problem " (tKaatrov rj Trpdft\i)f*a), its characteristic being that something has to be added to the enunciation in order to convert it from indeterminateness (aopttr-ria) to order (raf «) and scientific deter- minateness (Spas evta-TtjtioviKos) : such would be a problem bidding you " to construct an isosceles triangle," for the varieties of isosceles triangles are unlimited. Such "problems" are not problems in the proper sense {/cupim through its consequences to something admitted (to be) true. " Synthesis is an assumption of that which is admitted < and the passage > through its consequences to the finishing or attainment of what is sought." The language is by no means clear and has, at the best, to be filled out. Pappus has a fuller account 1 : " The so-called ivakv&fMPW (' Treasury of Analysis ") is, to put it shortly, a special body of doctrine provided for the use of those who, after finishing the ordinary Elements, are desirous of acquiring the power of solving problems which may be set them involving (the construction of) lines, and it is useful for this alone. It is the work of three men, Euclid the author of the Elements, Apollonius of Perga, and Aristaeus the elder, and proceeds by way of analysis and synthesis. " Analysis then takes that which is sought as if it were admitted and passes from it through its successive consequences to something which is admitted as the result of synthesis: for in analysis we assume that which is sought as if it were (already) done (yeyovos), and we inquire what it is from which this results, and again what is the ante- cedent cause of the latter, and so on, until by so retracing our steps we come upon something already known or belonging to the class of first principles, and such a method we call analysis as being solution backwards (deaira\tv \vtrtr). " But in synthesis, reversing the process, we take as already done that which was last arrived at in the analysis and, by arranging in their natural order as consequences what were before antecedents, and successively connecting them one with another, we arrive finally at the construction of what was sought ; and this we call synthesis. " Now analysis is of two kinds, the one directed to searching for the truth and called theoretical, the other directed to finding what we are told to find and called problematical, (1) In the theoretical kind we assume what is sought as if it were existent and true, after which we pass through its successive consequences, as if they too were true and established by virtue of our hypothesis, to something admitted : then (a), if that something admitted is true, that which is sought will also be true and the proof will correspond in the reverse order to the analysis, but (d), if we come upon something admittedly false, that which is sought will also be false. (2) In the problematical kind we assume that which is propounded as if it were known, after which we pass through its successive consequences, taking them as true, up to something admitted : if then (a) what is admitted is possible and obtainable, that is, what mathematicians call given, what was originally proposed will also be possible, and the proof will again correspond in 1 Pappus, VII. pp. 63+ — 6. ch. ix. $6] OTHER TECHNICAL TERMS 139 reverse order to the analysis, but if {&) we come upon something admittedly impossible, the problem will also be impossible." The ancient Analysis has been made the subject of careful studies by several writers during the last half-century, the most complete being those of Hankel, Duhamel and Zeuthen.; others by Ofterdinger and Cantor should also be mentioned'. The method is as follows. It is required, let us say, to prove that a certain proposition A is true. We assume as a hypothesis that A is true and, starting from this we find that, if A is true, a certain other proposition B is true ; if B is true, then C ; and so on until we arrive at a proposition K which is admittedly true. The object of the method is to enable us to infer, in the reverse order, that, since K is true, the proposition A originally assumed is true. Now Aristotle had already made it clear that false hypotheses might lead to a conclusion which is true. There is therefore a possibility of error unless a certain precaution is taken. While, for example, B may be a necessary consequence of A, it may happen that A is not a necessary consequence of B, Thus, in order that the reverse inference from the truth of K that A is true may be logically justified, it is necessary that each step in the chain of inferences should be unconditionally convertible. As a matter of fact, a very large number of theorems in elementary geometry are unconditionally convertible, so that in practice the difficulty in securing that the successive steps shall be convertible is not so great as might be supposed. But care is always necessary. For example, as Hankel says 1 , a proposition may not be uncon- ditionally convertible in the form in which it is generally quoted. Thus the proposition "The vertices of all triangles having a common base and constant vertical angle lie on a circle " cannot be converted into the proposition that "All triangles with common base and vertices lying on a circle have a constant vertical angle"; for this is only true if the further conditions are satisfied (]) that the circle passes through the extremities of the common base and {2) that only that part of the circle is taken as the locus of the vertices which lies on one side of the base. If these conditions are added, the proposition is unconditionally convertible. Or again, as Zeuthen remarks 3 , K may be obtained by a series of inferences in which A or some other proposition in the series is only apparently used ; this would be the case e.g. when the method of modem algebra is being employed and the expressions on each side of the sign of equality have been inadvertently multiplied by some composite magnitude which is in reality equal to zero. Although the above extract from Pappus does not make it clear that each step in the chain of argument must be convertible in the case taken, he almost implies this in the second part of the definition jf Analysis where, instead of speaking of the consequences B, C... 1 lAx^.^ZurGtstkichlederMathetnatikinAlterihumMttdMUitiatter^ 1874, pp. 137 — 150; Duhamd, Damithodis dans la scititra de raittmnement. Part I., 3 ed., Paris, 1885, pp. 39 — 68 ; Zeutben, Gmkuktt der MtUhematik im Attcrtnm und Miltclalttr, 1896, pp. 92 — 104; Oflerdinger, Beitrdp zur Gisehkhte der gritchischm MathmnUii, Ulm, i860; Cantor, Geschkkte der Mathtmatiki l t , pp. 110 — i. 1 Hankel, p. 139. * Zeiuhen, p. 103. Mo INTRODUCTION [ch. ix. 5 6 successively following from A, he suddenly changes the expression and says that we inquire what it is (B)frem which A follows (A being thus the consequence of B, instead of the reverse), and then what (viz. C) is the antecedent cause of B; and in practice the Greeks secured what was wanted by always insisting on the analysis being confirmed by subsequent synthesis, that is, they laboriously worked backwards the whole way from K. to A, reversing the order of the analysis, which process would undoubtedly bring to light any flaw which had crept into the argument through the accidental neglect of the necessary precautions. Reductio ad absurdum a variety of analysis. In the process of analysis starting from the hypothesis that a proposition A is true and passing through B, C... as successive con- sequences we may arrive at a proposition K which, instead of being admittedly true, is either admittedly false or the contradictory of the original hypothesis A or of some one or more of the propositions B, C... intermediate between A and K. Now correct inference from a true proposition cannot lead to a false proposition ; and in this case there- fore we may at once conclude, without any inquiry whether the various steps in the argument are convertible or not, that the hypo- thesis A is false, for, if it were true, all the consequences correctly inferred from it would be true and no incompatibility could arise. This method of proving that a given hypothesis is false furnishes an indirect method of proving that a given hypothesis A is true, since we have only to take the contradictory of A and to prove that it is false. This is the method of reductio ad absurdum, which is therefore a variety of analysis. The contradictory of A, or not-A, will generally include more t han one case and, in order to prove its falsity, each of the cases must be separately disposed of: e.g., if it is desired to prove that a certain part of a figure is equal to some other part, we take separately the hypotheses (i) that it is greater, (2) that it is less, and prove that each of these hypotheses leads to a conclusion either admittedly false or contradictory to the hypothesis itself or to some one of its consequences. Analysis as applied to problems. It is in relation to problems that the ancient analysis has the greatest significance, because it was the one general method which the Greeks used for solving all "the more abstruse problems" (ra deraG(TTepa twv TrpopKufiaTwvy '. We have, let us suppose, to construct a figure satisfying a certain set of conditions If we are to proceed at all methodically and not by mere guesswork, it is first necessary to "analyse" those conditions. To enable this to be done we must get them clearly in our minds, which is only possible by assuming all the conditions to be actually fulfilled, in other words, by supposing the problem solved. Then v;e have to transform those conditions, by all the means which practice in such cases has taught us to employ, into other conditions which are necessarily fulfilled if the original conditions are, and to continue this 1 Proclus, p. 141, 16, 17. CH. ix. $6] OTHER TECHNICAL TERMS 141 transformation until we at length arrive at conditions which we are in a position to satisfy 1 . In other words, we must arrive at some relation which enables us to construct a particular part of the figure which, it is true, has been hypothetically assumed and even drawn, but which nevertheless really requires to be found in order that the problem may be solved. From that moment the particular part of the figure becomes one of the data, and a fresh relation has to be found which enables a fresh part of the figure to be determined by means of the original data and the new one together. When this is done, the second new part of the figure also belongs to the data ; and we proceed in this way until all the parts of the required figure are found 1 . The first part of the analysis down to the point of discovery of a relation which enables us to say that a certain new part of the figure not belonging to the original data is given, Hankel calls the transformation ; the second part, in which it is proved that all the remaining parts of the figure are "given," he calls the resolution. Then follows the synthesis, which also consists of two parts, (1) the construction, in the order in which it has to be actually carried out, and in general following the course of the second part of the analysis, the resolution ; (2) the demonstration that the figure obtained does satisfy all the given conditions, which follows the steps of the first part of the analysis, the transformation, but in the reverse order. The second part of the analysis, the resolution, would be much facilitated and shortened by the existence of a systematic collection of Data such as Euclid's book bearing that title, consisting of propositions proving that, if in a figure certain parts or relations me. given, other parts or relations are also given. As regards the first part of the analysis, the trans- formation, the usual rule applies that every step in the chain must be unconditionally convertible ; and any failure to observe this condition will be brought to light by the subsequent synthesis. The second part, the resolution, can be directly turned into the construction since that only is given which can be constructed by the means provided in the Elements. It would be difficult to find a better illustration of the above than the example chosen by Hankel from Pappus*. Given a circle ABC and two points D, E external to it, to draw straight lines DB, KB from D, E to a point B on the circle such that, if DB, KB produced meet the circle again in C, A, AC shall be parallel A>DE. Analysis. Suppose the problem solved and the tangent at A drawn, meeting ED produced in F. (Part I. Transformation.) Then, since AC is parallel to DE, the angle at C is equal to the angle CDE. But, since FA is a tangent, the angle at C is equal to the angle FAE. Therefore the angle FAE is equal to the angle CDE, whence A, B, D, Fact concyclic. 1 Zcutben, p. 93. * Hankel, p. 141. ' Pappus, vn. pp. 630 — 1. 14* INTRODUCTION [ch. IX. 56 Therefore the rectangle AE, EB is equal to the rectangle FE, ED. (Part II, Resolution.) But the rectangle AE, EB is given, because it is equal to the square on the tangent from E. Therefore the rectangle FE, ED is given ; and, since ED is given, FE is given (in length). [Data, 57.J But FE is given in position also, so that F is also given. [Data, 27.] Now FA is the tangent from a given point F to a circle A BC given in position ; therefore FA is given in position and magnitude. [Data, 90.] And F is given ; therefore A is given. But £ is also given ; therefore the straight line AE is given in position, [Data, 26.] And the circle ABC is given in position ; therefore the point B is also given. [Data, 25.] But the points D, E are also given ; therefore the straight lines DB, BE are also given in position. Synthesis. (Part I. Construction) Suppose the circle ABC and the points D, E given. Take a rectangle contained by ED and by a certain straight line EF equal to the square on the tangent to the circle from E. From F draw FA touching the circle in A ; join ABE and then DB, producing DB to meet the circle at C. Join AC. 1 say then that A C is parallel to DE. (Part II. Demonstration.) Since, by hypothesis, the rectangle FE, ED is equal to the square on the tangent from E, which again is equal to the rectangle AE, EB, the rectangle AE, EB is equal to the rectangle FE, ED, Therefore A, B, D, F are concyclic, whence the angle FAE is equal to the angle BDE. But the angle FAE is equal to the angle ACB in the alternate segment ; therefore the angle A CB is equal to the angle BDE. Therefore AC Is parallel to DEk In cases where a &iopt' lavTrjii Anal. post. II. 1, 90 a 31. ' Anal. post. II. », 90 & 1} — 11. * ibid. u. 8, 93 * 17. * ibid, 93 a 10. ' Trendelenburg, Erlduttrungtn, p. no. * Di amma II, 1, 413 a 13— JO. * Anal. posU II. a, 00 a 6, •■Mttvu, 94 a 18. ISO INTRODUCTION [ch. ix. §7 it should be with the definition, A definition which is to show the genesis of the thing defined should contain the middle term or cause ; otherwise it is a mere statement of a conclusion. Consider, for instance, the definition of " quadrature " as " making a square equal in area to a rectangle with unequal sides," This gives no hint as to whether a solution of the problem is possible or how it is solved : but, if you add that to find the mean proportional between two given straight lines gives another straight line such that the square on it is equal to the rectangle contained by the first two straight lines, you supply the necessary middle term or cause 1 . Technical term a not defined by Euclid. It will be observed that what is here defined, " quadrature " or " squaring " (TeTpaiyawio-^oc), is not a geometrical figure, or an attribute of such a figure or a part of a figure, but a technical term used to describe a certain problem. Euclid does not define such things ; but the fact that Aristotle alludes to this particular definition as well as to definitions of deflection (tce/ckd-a-Oat) and of verging {vevew) seems to show that earlier text-books included among definitions explanations of a number of technical terms, and that Euclid deliberately omitted these explanations from his Elements as surplusage. Later the tendency was again in the opposite direction, as we see from the much expanded Definitions of Heron, which, for example, actually include a definition of a deflected line (KettXaajUw) ypa/i,^)'. Euclid uses the passive of icTJiv occasionally*, but evidently considered it unnecessary to explain such terms, which had come to bear a recognised meaning. The mention too by Aristotle of a definition of verging (vcveiv) suggests that the problems indicated by this term were not excluded from elementary text-books before Euclid. The type of problem (vev Things which coincide with one another are equal to one another. [8] 5. The whole is greater than the part Definition i. Si)/i ttov ia-Ttv, uv [icpiK oi$iy. A point is that which has no pari. An exactly parallel use of fiepos (itrrt) in the singular is found in Aristotle, Metaph. 1035 b 32 fitpos /11F ow tort (tat tqv mSovi, literally "There is a part even of the form "; Borsitz translates as if the plural were used, "Theile giebt es," and the meaning is simply "even the form is divisible (into parts)." Accordingly it would be quite justifiable to translate in this case "A point is that which is indivisible into parts." Martianus Capella (5th c. a.d.) alone or almost alone translated differently, "Punctum est cuius pars nihil est," "a point is that a part of which is netting." Notwithstanding that Max Simon (Euclid vnd die sechs planimetrischen Sucker, 1 901) has adopted this translation (on grounds which I shall presently mention), I cannot think that it gives any sense. If a part of a point is nothing, Euclid might as well have said that a point is itself "nothing," which of course he does not do. Pre -Euclidean definitions. It would appear that this was not the definition given in earlier text- books; for Aristotle (Topics vi. 4, 141 b 20), in speaking of "the definitions" of point, line, and surface, says that they alt define the prior by means of the posterior, a point as an extremity of a line, a line of a surface, and a surface of a solid The first definition of a point of which we hear is that given by the Pythagoreans (cf. Proclus, p. 95, 21), who defined it as a "monad having position" or "with position added" (m°vo; irpoo-kafjowra 9i% a'wXaT«. A line is breadthless length. This definition may safely be attributed to the Platonic School, if not to Plato himself. Aristotle (Topics vi. 6, 143 b 11) speaks of it as open to objection because it "divides the genus by negation," length being necessarily either breadthless or possessed of breadth ; it would seem however that the objection was only taken in order to score a point against the Platonists, since he says (ibid. 143 b 29) that the argument is "of service only against those who assert that the genus [sc. length] is one numerically, that ts, those who assume ideas," e.g. the idea of length (amo u^*o$) which they regard as a genus : for if the genus, being one and self-existent, could be divided into two species, one of which asserts what the other denies, it would be self- contradictory (WaiU), Proclus (pp. 96, 21—97, 3) observes that, whereas the definition of a point is merely negative, the line introduces the first "dimension," and so its definition is to this extent positive, while it has also a negative element which denies to it the other " dimensions " (Suur-rdtrtii). The negation of both breadth and depth is involved in the single expression "breadthless" (atrXarti), since everything that is without breadth is also destitute of depth, though the converse is of course not tnie. Alternative definitions. The alternative definition alluded to by Proclus, fiiyiOtn iip' ty Suumjov " magnitude in one dimension " or, better perhaps, " magnitude extended one way " (since Suuttoitk as used with reference to line, surface and solid scarcely corresponds to our use of " dimension " when we speak of "one," " two," or " three dimensions "), is attributed by an-Nairlzs to " Heromides," who must presumably be the same as " Herundes," to whom he attributes a certain definition of a point. It appears however in substance in Aristotle, though Aristotle does not use the adjective Sunmirw, nor does he apparently use 8«wrrao-« except of body as having three " dimensions " or " having dimension (or extension J a// ways (vavrg)," the "dimensions" being in his view (1) up and down, (2) before and behind, and (3) right and left, and " up " being the principle or beginning of length, " right " of breadth, and " before " of depth (De cat So 11. 2, 284 b 24). A line is, according to Aristotle, a magnitude " divisible in one way only " (jtattajrg iuLifttrov), in contrast to a magnitude divisible in two ways (&XB ZuxtpiTOv), or a surface, and a magnitude divisible "in all or in three ways" (iramg koX rpixfi jtnipcToV), or a body (Metaph, 1016 b 25 — 27); or it is a magnitude "continuous one way (or in one direction)," as compared with magnitudes continuous tovo ways or three ways, i. def. 2] NOTES ON DEFINITIONS i, 2 159 which curiously enough he describes as " breadth " and " depth " respectively (jtiytdos Zi to p.iv lift tv uvvi^h firjitm, to 8' hrl Suo tAiitos, to 8' twi rpia (3a8o<;, Metaph. 1020 a 11), though he immediately adds that " length " means a line, " breadth " a surface, and " depth " a body. Proclus gives another alternative definition as "flux of a point " (/5ucr« tnjfMuiv), i.e. the path of a point when moved. This idea is also alluded to in Aristode (De anima 1. 4, 409 a 4 above quoted) : " they say that a line by its motion produces a surface, and a point by its motion a line." "This definition," says Proclus (p. 97, 8 — r3), "is a perfect one as showing the essence of the line : he who called it the flux of a point seems to define it from its genetic cause, and it is not every line that he sets before us, but only the immaterial line , for it is this that is produced by the point, which, though itself indivisible, is the cause of the existence of things divisible." Proclus (p. r 00, 5 — 19) adds the useful remark, which, he says, was current in the school of Apollonius, that we have the notion of a line when we ask for the length of a road or a wall measured merely as length ; for in that case we mean something irrespective of breadth, viz. distance in one *' dimension." Further we can obtain sensible perception of a line if we look at the division between the light and the dark when a shadow is thrown on the earth or the moon ; for clearly the division is without breadth, but has length. Species of "lines." After defining the " line " Euclid only mentions one species of line, the straight line, although of course another species appears in the definition of a circle later. He doubtless omitted all classification of lines as unnecessary for his purpose, whereas, for example, Heron follows up his definition of a line by a division of lines into (1) those which are " straight " and {2} those which are not, and a further division of the latter into (a) " circular circumferences," {&) "spiral-shaped" (iiutottScw) lines and (r) "curved" (xajim-vAai) lines generally, and then explains the four terms. Aristotle tells us {Metaph. 986 a 25) that the Pythagoreans distinguished straight (iJW) and curved (ku/u t',W), and this distinction appears in Plato (cf. Republic x. 602 c) and in Aristotle (cf. " to a line belong the attributes straight or curved," Anal. post. 1. 4, 73 b 19; "as in mathematics it is useful to know what is meant by the terms straight and curved," De anima I. 1, 402 b 19). But from the class of " curved " lines Plato and Aristotle separate off the urtpi^ipijs or " circular " as a distinct species often similarly contrasted with straight. Aristotle seems to recognise broken lines forming an angle as one line : thus "a line, if it be bent (xosafi- pin}), but yet continuous, is called one" (Me tap A. 1 01 6 a 2); "the straight line is more one than the bent line" (Hid. 1016 a 12). Cf. Heron, Def. 12, "A broken line (jeckW/u'vij y pappy) so-called is a line which, when produced, does not meet itself." When Proclus says that both Plato and Aristotle divided lines into those which are "straight," "circular" (wtpujxpife) or "a mixture of the two," adding, as regards Plato, that he included in the last of these classes " those which are called helicoidal among plane (curves) and (curves) formed about solids, and such species of curved lines as arise from sections of solids " (p. 104, 1 — 5), he appears to be not quite exact. The reference as regards Plato seems to be to Parmenides 145 B: "At that rate it would seem that the one must have shape, either straight or round (arpoyyvKov) or some combination of the two"; but this scarcely amounts to a formal classification of lines. As regards i6o BOOK I [l. DEF. 2 Aristotle, Proclus seems to have in mind the passage (De each i. 2, 268 b 17) where it is stated that " all motion in space, which we call translation ($apd), is (in) a straight line, a circle, or a combination of the two ; for the first two ate the only simple (motions)." For completeness it is desirable to add the substance of Proclus' account of the classification of lines, for which he quotes Geminus as his authority. Geminus* first classification of lines. This begins {p. in, 1 — 9) with a division of lines into composite [avtSrrtK) and incomposite (io-u'ec'rrov). The only illustration given of the composite class is the "broken line which forms an angle" (9 KftXavjiivy) ™1 ywiW ttoumaa) ; the subdivision of the incomposite class then follows (in the text as it stands the word " composite " is clearly an error for " incomposite "). The subdivisions of the incomposite class are repeated in a later passage (pp. 176, 27 — 177, 23) with some additional details. The following diagram reproduces the effect of both versions as far as possible (all the illustrations mentioned by Proclus being shown in brackets). lines composite (broken line forming an angle) incomposite forming a figure GXytHiTtrroiQvaa.i or determinate (circle, ellipse, eissoid) not forming a figure or indeterminate d6pt&Toi and extending without limit i-r Arttpop iKfia.W&fin'a.t (straight line, parabola, hyperbola, conchoid) The additional details in the second version, which cannot easily be shown in the diagram, are as follows : (1) Of the lines which extend without limit, some do not form a figure at all (viz. the straight line, the parabola and the hyperbola); but some first "come together and form a figure" (i.e. have a loop), "and, for the rest, extend without limit " (p. 177, 8). As the only other curve, besides the parabola and the hyperbola, which has been mentioned as proceeding to infinity is the conchoid (of Nicomedes), we can hardly avoid the conclusion of Tannery 1 that the curve which has a loop and then proceeds to infinity is a variety of the conchoid itself. As is 1 Notes. pour thistoire des ligncs et surf aits tourbes dans Fatttiquiii in Bulletin des sriemts mathim, ct astronam. 1 ser, vm. (1884), pp. 108—0 (Minmires seientifiaues, ri. p. 13). I. def. 2] NOTE ON DEFINITION 2 161 well known, the ordinary conchoid (which was used both for doubling the cube and for trisecting the angle) is obtained in this way. Suppose any number of rays passing through a fixed point (the pole) and intersecting a fixed straight line ; and suppose that points are taken on the rays, beyond the fixed straight line, such that the portions of the rays intercepted between the fixed straight line and the point are equal to a constant distance (hiacrnj/ia), the locus of the points is a conchoid which has the fixed straight line for asymptote. If the "distance" a is measured from the intersection of the ray with the given straight line, not in the direction away from the pole, but towards the pole, we obtain three other curves according as a is less than, equal to, or greater than b, the distance of the pole from the fixed straight line, which is an asymptote in each case. The case in which a ■■■■ l> gives a curve which forms a loop and then proceeds to infinity in the way Proclus describes. Now we know both from Eutocius {Csmm. on Archimedes, ed. Heiberg, in. p. 98) and Proclus (p. 272, 3 — 7) that Nicomedes wrote on conchoidr (in the plural), and Pappus 3 ~ l 9)- 2. Mixed lines. It might be supposed, says Proclus (p. 105, n), that the cylindrical helix, being homoeomerie, like the straight line and the circle, must like them be simple. He replies that it is not simple, but mixed, because it is generated by two unlike motions. Two like motions, said Geminus, e.g. two motions at the same speed in the directions of two adjoining sides of a square, produce a simple line, namely a straight line (the diagonal) ; and again, if a straight line moves with its extremities upon the two sides of a right angle respectively, this same motion gives a simple curve (a circle) for the locus of the middle point of the straight line, and a mixed curve (an ellipse) for the locus of any Other point on it (p. 106, 3—15). Geminus also explained that the term " mixed," as applied to curves, and as applied to surfaces, respectively, is used in different senses. As applied to curves, "mixing" neither means simple "putting together" (js). And the spiric sections are three according to these three differences" (p. 119, 8-r 7 ). " When the hippopede, which is one of the spiric curves, forms an angle with itself, this angle also is contained by mixed lines" (p. 1.27, 1 — 3). " Perseus showed for spirics what was their property {a-vinrr^iia) " (P- 356. ")■ Thus the spiric surface was what we call a tore, or (when open) an anchor- ring. Heron (Def. 97) says it was called alternatively spire {uirflpa} or ring (xpuur;); he calls the variety in which "the circle cuts itself," not "interlaced," but " cross ing-ltse If" (iTrakkaTTUwra). Tannery 1 has discussed these passages, as also did Schiaparelli*. It is clear that Prochis' remark that the difference in the three curves which he mentions corresponds to the difference between the three surfaces is a slip, due perhaps to too hurried transcribing from Geminus ; all three arise from plane sections of the open anchor-ring. If r is the radius of the revolving circle, a the distance of its centre from the axis of rotation, d the distance of the plane section (supposed to be parallel to the axis) from the axis, the three curves described in the first extract correspond to the following cases : (1) d=a~r. In this case the curve is the hippopede, of which the lemniscate of Bernoulli is a particular case, namely that in which a = zr. The name hippopede was doubtless adopted for this one of Perseus' curves on the ground of its resemblance to the hippopede of Eudoxus, which seems to have been the curve of intersection of a sphere with a cylinder touching it internally. (2) a + r>d>a. Here the curve is an oval, (3) a > rf> a - r. The curve is now narrowest in the middle. Tannery explains the " three lines upon (in addition to) five sections " thus. He points out that with the open tore there are two other sections corresponding to (4) d= a : transition from (2) to (3). (5) a- r>d> o, in which case the section consists of two symmetrical ovals. He then shows that the sections of the closed or continuous tore, corre- sponding to a = r, give curves corresponding to (2), (3) and (4) only. Instead of (t) and (5) we have only a section consisting of two equal circles touching one another. On the other hand, the third spire (the interlaced variety) gives three new forms, which make a group of three in addition to the first group oifive sections. 1 Poiw fkisteire des lignts et surfaces conrbts dans tantiquiti in Bulletin tits stientes mat him. et astranom. vill. (i88+|, pp. 35— % 7 [Mlnairts stientifiqiw, II. pp. 14— 18). s Du kemocentrischen SpAarcn des EudaxuSi ties Kallippus und dts Aristottles {Adhartd* lungm tur Gesch. der Math. 1. Heft, 187;, pp. 14Q — i6 4 BOOK J [l. DF-F. a The difficulty which I see in this interpretation is the fact that, just after " three lines on five sections " are mentioned, Proclus describes three curves which were evidently the most important ; but these three belong to three of the five sections of the open tore, and are not separate from them, 4. The cissoid. This curve is assumed to be the same as that by means of which, according to Eutocius (Comm. on Archimedes, in. p. 66 sqq.), Diodes in his book ir«pi ■Kvpiwv (On burning-glasses) solved the problem of doubling the cube. It is the locus of points which he found by the following construction. Let AC, BD be diameters at right angles in a circle with centre O. Let E, Fbe points on the quadrants BC, BA respectively such that the arcs BE, BE ait equal. Draw EG, FH perpendicular to CA. D Join AE, and let P be its intersection with FH. The cissoid is the locus of all the points P corresponding to different posi- tions of E on the quadrant BC and of F at an equal distance from B along the arc BA. A ts the point on the curve correspond- ing to the position C for the point E, and B the point on the curve corresponding to the position of E in which it coincides with B. It is easy to see that the curve extends in the direction AB beyond B, and that CK drawn perpendicular to CA is an asymptote. It may be regarded also as having a branch AD symmetrical with AB, and) beyond D, approaching KC produced as asymptote. If OA, 0£> are coordinate axes, the equation of the curve is obviously /(« + *) = ("-*)*> where a is the radius of the circle. There is a cusp at A, and it agrees with this that Proclus should say (p. 1 a 6, 34^ that "cissoidal lines converging to one point like the leaves of ivy — for this is the origin of their name — form an angle." He makes the slight correction (p. 1 28, 5) that it is not two farts of a curve, but one curve, which in this case makes an angle. But what is surprising is that Proclus seems to have no idea of the curve passing outside the circle and having an asymptote, for he several times speaks of it as a closed curve (forming a figure and including an area) : cf. p. 152, 7, "the plane (area) cut off by the cissoidal line has one bounding (line), but it has not in it a centre such that all (straight lines drawn to the curve) from it are equal." It would appear as if Proclus regarded the cissoid as formed by the/our symmetrical cissoidal arcs shown in the figure. Even more peculiar is Proclus' view of the 5, "Single-turn Spiral." This is really the spiral of Archimedes traced by a point starting from the fixed extremity of a straight line and moving uniformly along it, while /' .--'' ° H '•'-.J 5 E 7 s~"^ / j !. Dvrr. 2—4] NOTES ON DEFINITIONS 2—4 165 simultaneously the straight line itself moves uniformly in a plane about its fixed extremity. In Archimedes the spiral has of course any number of turns, the straight line making the same number of complete revolutions. Yet I'roclus, while giving the same account of the generation of the spiral (p. 180, 8 — 12), regards the single-turn spiral us actually stopping short at the point reached after one complete revolution of the straight line : " it is necessary to know that extending without limit is not a property of all lines ; for it neither belongs to the circle nor to the cissoid, nor in general to lines which form figures ; nor even to those which do not form figures. For even the single- turn spiral does not extend without limit— -for it is constructed between two points^nor does any of the other lines so generated do so" (p. 187, 19 — 25). It is curious that Pappus (vm. p. n 10 sqq. } uses the same term finvoVrpoi^ns \\ii to denote one turn, not of the spiral, but of the cylindrical helix. Definition 3. Tp*Hprj$ SI Tripara {TTj^Mld. The extremities of a line art points. It being unscientific, as Aristotle said, to define a point as the " extremity of a line " (iripas ypa/iftjjs), thereby explaining the prior by the posterior, Euclid defined a point differently ; then, as it was necessary to connect a point with a line,, he introduced this explanation after the definitions of both had been given. This compromise is no doubt his own idea; the same thing occurs with reference to a surface and a line as its extremity in Def. 6, and with reference to a solid and a surface as its extremity in xt. Def. 2. We miss a statement of the facts, equally requiring to be known, that a " division " (Suupio-ij) of a line, no less than its " beginning " or " end," is a point (this is brought out by Aristotle: cf, Metapk. 1060 b 15), and that the intersection of two lines is also a point. If these additional explanations had been given, Proclus would have been spared the difficulty which he finds in the fact that some of the tines used in Euclid (namelv infinite straight lines on the one hand, and circles on the other) have no " extremities." So also the ellipse, which Proclus calls by the, old name (foptfc (" shield "). In the case of the circle and ellipse we can, he observes (p. 105, 7), take a portion bounded by points, and the definition applies to that portion. His rather far-fetched distinction between two aspects of a circle or ellipse as a line and as a closed figure (thus, while you are describing a circle, you have two extremi- ties at any moment, but they disappear when it is finished) is an unnecessarily elaborate attempt to establish the literal universality of the "definition," which is really no more than an explanation that, if a line has extremities, those extremities are points. Definition 4. Eu#eiu yfnififurj tarty, $rtf ci laov toIs ioir ratv iaxdrotv btiTrpottOfv p). Aristotle quotes it in equivalent terms {Topics vi. n, 148 b 27), o5 to /h'o-oc &r«rnotr#«I row inpao-ii' ; and, as he does not mention the name of its author, but states it in combina- tion with the definition of a line as the extremity of a surface, we may assume that he used it as being well known. Proclus also quotes the definition as Plato's in almost identical terms, fc to fU tauTtji r« i( Urov, " there all are on a footing of equality." Slightly different are the uses in Aristotle, Eth, Nie. x. 8, 1178 a 25 rur /ikr yap iwyimri'w xP tul *<" i£ Urov Itrrta, "both need the necessaries of life to the same extent, Set us say"; Topics ix, 15, 174 a 32 i$ urov iroioWa rije tpiinjatv, "asking the question indifferently" {i.e. without showing any expectation of one answer being given rather than another). The natural meaning would therefore appear to be "evenly placed" (or balanced), "in equal measure," " indifferently" or "without bias" one way or the other. Next, is the dative rots i' io.vrr}t o"i)/u£ok constructed with i( urov or with MttBt? In the first case the phrase must mean "that which he&evenfy with (or in respect to) the points on it," in the second apparently "that which, in (or by) the points on it, lies (or is placed) evenly (or uniformly)." Max Simon takes the first construction to give the sense "die Gerade liegt in gleicher Weise wie ihre Punkte." If the last words mean " in the same way as (or in like manner as) its points," I cannot see that they tell us anything, although Simon attaches to the words the notion of distance (Abstand) like Proclus. The second construction he takes as giving " die Gerade liegt fur (durch) ihre Punkte gleichmassig," " the straight line lies symmetrically for (or through) its points"; or, if k«t h), "continuous one way" (i tr trunx^), or "divisible in one way" (^iowiy^ Statpfroi-), so a surface is a magnitude extended or continuous two ways (hi Sw>), or divisible in two ways (Sixi)). As in Euclid a surface has " length and breadth " only, so in Aristotle " breadth " is characteristic of the surface and is once used as synonymous with it (Metaph. tojo a ix), and again "lengths are made up of long and short, surfaces of broad and narrow, and solids (oyi™) of deep and shallow" (Metaph. 1085 a 10). Aristotle mentions the common remark that a line by its motion produces a surface (De anima ]. 4, 409 a 4). He also gives the a posteriori description of a surface as the "extremity of a solid" (Topics vi. 4, [41 b 21}, and as "the section (rofnf) or division {S«up«rii) of a body" (Metaph. 1060 b 14). Proclus remarks (p. 114, jo) that we get a notion of a surface when we measure areas and mark their boundaries in the sense of length and breadth ; and we further get a sort of perception of it by looking at shadows, since these have no depth {for they do not penetrate the earth) but only have length and breadth. Classification of surfaces. Heron gives (Def. 74, p. 50, ed. Heiberg) two alternative divisions or surfaces into two classes, corresponding to Gemirtus' alternative divisions of lines, viz. into (1) incomposite and composite and (2) simple and mixed. (1) Incomposite surfaces are "those which, when produced, fall into (or coalesce with) themselves" (wtim iK0akXo/ityai, airat tiaff iavrwv Wwrowie), i.e. are of continuous curvature, e.g. the sphere. Composite surfaces are "those which, when produced, cut one another." Of composite surfaces, again, some are (a) made up of non-homogeneous (elements) (i$ dvofiotoyow) such as cones, cylinders and hemispheres, others (&) made up of homogeneous (elements), namely the rectilineal (or polyhedral) surfaces. (2) Under the alternative division, simple surfaces are the plane and the spherical surfaces, but no others ; the mixed class includes all other surfaces whatever and is therefore infinite in variety. Heron specially mentions as belonging to the mixed class (a) the surface of cones, cylinders and the like, which are a mixture of plane and circular (fitKTal i£ tTrnrt'Sou Kal irepic^epttas) and (h) spirie surfaces, which are "a mixture of two circumferences " (by which he must mean a mixture of two circular elements, namely the generating circle and its circular motion about an axis in the same plane). Proclus adds the remark that, curiously enough, mixed surfaces may arise by the revolution either of simple curves, e.g. in the case of the spire, or of mixed curves, e.g. the "right-angled conoid" from a parabola, "another conoid" from the hyperbola, the "oblong" (krifinKts, in Archimedes iropa- pa-xts) and " flat " (VwtjrAa-nJ) spheroids from an ellipse according as it revolves about the major or minor axis respectively (pp. 119, 6 — 120, 2). The homoeo- meric surfaces, namely those any part of which will coincide with any other part, are two only (the plane and the spherical surface), not three as in the case of lines (p. no, 7). i. deff. 6, 7] NOTES ON DEFINITIONS 5— 7 i?» Definition 6. 'Eirt^avtias Si -ripara. ypappmi. The extremities of a surface are lines. It being unscientific, as Aristotle says, to define a line as the extremity of a surface, Euclid avoids the error of defining the prior by means of the posterior in this way, and gives a different definition not open to this objection. Then, by way of compromise, and in order to show the connexion between a line and a surface, he adds the equivalent of the definition of a line previously current as an explanation. As in the corresponding Def. 3 above, he omits to add what is made clear by Aristotle (Metapk. 1060 b 15) that a "division" (Siaipco-is) or " section " (to/ii;) of a solid or body is also a surface, or that the common boundary at which two parts of a solid fit together (Categories 6, 5 a a) may be a surface. Proclus discusses how the fact stated in Def. 6 can be said to be true of surfaces like that of the sphere "which is bounded (wtiripairrat), it is true, but not by lines." His explanation (p. 116, 8 — 14) is that, "if we take the surface (of a sphere), so far as it is extended two ways (BtyjJ &aobachewsky evolved the plane is of course equivalent to the definition of a plane as the locus of all points equidistant from two fixed points in space. iy6 BOOK I [i. deff, 7—9 Leibniz in a letter to Giordano defined a plane as thai surface which divides space into turn congruent farts. Adverting to Giordano's criticism that you could conceive of surfaces and lines which divided space or a plane into two congruent parts without being plane or straight respectively, Beez ( liber Euklidischc und Nicht-Euklidische Geometric, 1888) pointed out that what was wanted to complete the definition was the further condition that the two congruent spaces could be slid along each other without the surfaces ceasing to coincide, and claimed priority for his completion of the definition in this way. But the idea of all the parts of a plane fitting exactly on all other parts is ancient, appearing, as we have seen, in Heron, Def. 9. Definitions 8, 9. 8. '£mV«Sos &i yiMivla iariv y iv cwnri&w Svo ypafijjL&v OriTTOfUvtav a\kijktitv jcat i*t] tV iiOua.% Ktifiivtiiv TTfm aAAijXuf ruv ypu^/Hav xAurtt. 9. "Orav &* al Trfptixowai ttjv yutvcW ypaftfXal tv&4itu umtiv, tv&uypaftftos jraA.«Tarflf), where again an angle is supposed to be formed by one broken line or surface. Still more interesting, perhaps, is the definition by " those who say that the first distance under the point (ri •wp&rar i. Dbff. 8, 9] NOTES ON DEFINITIONS 7—9 177 &uiim)iia thro to mffuiw) is the angle. Among these is Plutarch, who insists that Apollonius meant the same thing ; for, he says, there must be some first distance under the breaking (or deflection) of the including lines or surfaces, though, the distance under the point being continuous, it is impossible to obtain the actual first, since every distance is divisible without limit" (**•' ajrttpov). There is some vagueness in the use of the word " distance" (Suwrnj/ia) ; thus it was objected that " if we anyhow separate off the first " (distance being apparently the word understood) " and draw a straight line through it, we get a triangle and not one angle." In spite of the objection, I cannot but see in the idea of Plutarch and the others the germ of a valuable conception in infinitesimals, an attempt (though partial and imperfect) to get at the rate of divergence between the lines at their point of meeting as a measure of the angle between them. A third view of an angle was that of Carpus of Antioch, who said " that the angle was a quantity (too-ov), namely a distance (SuMmf/m) between the lines or surfaces containing it This means that it would be a distance (or divergence) in one sense (ie£y<> yoivta? urac cIAA^'Xacs TOtp, Ap&i) itcaripa i£r law ywetuie fart, Hal ij l$tt^fp6ypajifu>v) plane figure Dounded by one line" {De taeh n. 4, 286 b 13 — 16); "the plane equal (i.e. extending equally all ways) from the middle " (hrUtfov to i*. tov p.iao\s urov), meaning a circle (Rhetoric 111. 6, 1407 b 27); he also contrasts with the circle "any other figure which has not the lines from the middle equal, as for example an egg-shaped figure" {De eaelo n. 4, 287 a 19). The word "centre" {nivrpov) was also regularly used : cf. Produs' quotation from the " oracles " (Wyto), " the centre from which all (lines extending) as far as the rim are equal." The definition as it stands has no genetic character. It says nothing as to the existence or non-existence of the thing defined or as to the method of constructing it. It simply explains what is meant by the word " circle," and is a provisional definition which cannot be used until the existence of circles is proved or assumed. Generally, in such a case, existence is proved by actual construction ; but here the possibility of constructing the circle as defined, and consequently its existence, are postulated (Postulate 3). A genetic definition might state that a circle is the figure described when a straight line, always remaining in one plane, moves about one extremity as a fixed point until it returns to its first position (so Heron, Def. ay). Simplicius indeed, who points out that the distance between the feet of a pair of compasses is a straight line from the centre to the circumference, will have it that Euclid intended by this definition to show how to construct a circle by the revolution of a straight line about one end as centre ; and an- Nairlzi points to this as the explanation (r) of Euclid's definition of a circle as a plane figure, meaning the whole surface bounded by the circumference, and not the circumference itself, and (2) of his omission to mention the " circumference," since with this construction the circumference is not drawn separately as a line. But it is not necessary to suppose that Euclid himself did more than follow the traditional view ; for the same conception of the circle as a plane figure appears, as we have seen, in Aristotle. While, however, i. Deff. 15-17] NOTES ON DEFINITIONS 15—17 185 Euclid is generally careful to say the "circumference of a circle " when he means the circumference, or an arc, only, there are cases where "circle" means "circumference of a circle," e.g. in ill. 10 1 "A circle does not cut a circle in more points than two." Heron, Proclus and Simplicius are all careful to point out that the centre is not the only point which is equidistant from all points of the circumference. The centre is the only point in the plane of the circle ("lying within the figure," as Euclid says) of which this is true; any point not in the same plane which is equidistant from all points of the circumference is a pole. If you set up a "gnomon " (an upright stick) at the centre of a circle (i.e. a line through the centre perpendicular to the plane of the circle), its upper extremity is a pole (Proclus, p. 153, 3); the perpendicular is the locus of all such poles. Definition 17. ktAfitrpof Si rov kvkKuv ttrrlv tvOttOr tls &ia tw Ktvrpov Tjyfj&vrj k. Stya Ti/ivfi tov kvx\qv. A diameter of the circle is any straight line drawn through the centre and terminated in both directions by the circumference of the circle, and such a straight line also bisects the circle. The last words, literally " which (straight line) also bisects the circle," are omitted by Simson and the editors who followed him. But they are necessary even though they do not " belong to the definition " but only express a property of the diameter as defined. For, without this explanation, Euclid would not have been justified in describing as a .fcrjw-circle a portion of a circle bounded by a diameter and the circumference cut off by it. Simplicius observes that the diameter is so called because it passes through the whole surface of a circle as if measuring it, and also because it divides the circle into two equal parts. He might however have added that, in general, it is a line passing through a figure where it is widest, as well as dividing it equally : thus in Aristotle ri nara Sia/ttrpw KtCptva, " things diametrically situated " in space, are at their maximum distance apart. Diameter was the regular word in Euclid and elsewhere for the diameter of a square, and also of a parallelogram; diagonal (Stayuptot) was a later term, defined by Heron (Def. by] as the straight line drawn from an angle to an angle. Proclus (p. 157, 10) says that Thales was the first to prove that a circle is bisected by its diameter; but we are not told how he proved it. Proclus gives as the reason of the property "the undeviating course of the straight line through the centre " (a simple appeal to symmetry), but adds that, if it is desired to prove it mathematically, it is only necessary to imagine the diameter drawn and one part of the circle applied to the other ; it is then clear that they must coincide, for, if they did not, and one fell inside or outside the other, the straight lines from the centre to the circumference would not all be equal : which is absurd. Saccheri's proof is worth quoting. It depends on three " Lemmas " immediately preceding, (1) that two straight lines cannot enclose a space, (z) that two straight lines cannot have one and the same segment common, (3) that, if two straight lines meet at a point, they do not touch, but cut one another, at it. " Let MDHNKM be a circle, A its centre, MN a diameter. Suppose i86 BOOK I [i. Deff. 17, tS the portion MNKM of the circle turned about the fixed points M, N, so that it ultimately comes near to or coincides with the remaining portion MNHDM. "Then (i) the whole diameter MAN', with all its points, clearly remains in the 3ame position, since otherwise two straight lines would enclose a space (contrary to the first Lemma). " (ii) Clearly no point K of the circumference NKM falls within or outside the surface enclosed by the diameter MANand the other part, NHDM, of the circumference, since otherwise, contrary to the nature of the circle, a radius as AK would be less or greater than another radius as All. " (iii) Any radius MA can clearly be rectilineally produced only along a single other radius AN, since otherwise (contrary to the second Lemma) two lines assumed straight, e.g. MAN, MAH, would have one and the same common segment. " (iv) All diameters of the circle obviously cut one another in the centre (Lemma 3 preceding), and they bisect one another there, by the general properties of the circle. " From all this it is manifest that the diameter MAN divides its circle and the circumference of it just exactly into two equal parts, and the same may be generally asserted for every diameter whatsoever of the same circle ; which was to be proved." Simson observes that the property is easily deduced from tit. 31 and 24 ; for it follows from lit. 31 that the two parts of the circle are "similar segments" of a circle (segments containing equal angles, in. Def. 11), and from tti. 24 that they are equal to one another. Definition 18. H/iuruKXiov hi itTTt to Trtptt^ofitvov tr)(fffia. inrd re t?/s BiafUTpov Kat Trj% Ivokafi^awo/Ltv^^ iiir avr^e vtpi^tpiia^. nivTfujv Si tov ijfiiKvtt\{oii to avrd, & Ktll TOV KVtikrtv CffTtV. A semicircle if the figure contained by the diameter and the circumference cut off by it. And the centre of the semicircle is the same as that of the circle. The last words, "And the centre of the semicircle is the same as that of the circle," are added from Proclus to the definition as it appears in the MSS. Scarburgh remarks that a semicircle has no centre, properly speaking, and thinks that the words are not Euclid's, but only a note by Proclus. I am however inclined to think that they are genuine, if only because of the very futility of an observation added by Proclus. He explains, namely, that the semicircle is the only plane figure that has its centre on its perimeter (!), "so that you may conclude that the centre has three positions, since it may be within the figure, as in the case of a circle, or on the perimeter, as with the semicircle, or outside, as with some conic lines (the single-branch hyperbola presumably)" ! Proclus and Simplicius point out that, in the order adopted by Euclid for these definitions of figures, the first figure taken is that bounded by on* line (the circle), then follows that bounded by two lines (the semicircle), then the triangle, bounded by three lines, and so on. Proclus, as usual, distinguishes I. Deff. 18-ai] NOTES ON DEFINITIONS 17—31 187 different kinds of figures bounded by two lines (pp. 159, 14 — 160, 9). Thus they may be formed (1) by circumference and circumference, e.g. (a) those forming angles, as a tunc (to pijvottSn) and the figure included by two arcs with convexities outward, and (b) the angle-less (iytinov), as the figure included between two concentric circles (the coronal) ; (2) by circumference and straight line, e.g. the semicircle or segments of circles (tty£S * our* i 31 it seems to be a quadrilateral, and in Metapk. 1054 b 2, " equal and equiangular Ttrpdyaiva," it cannot be anything else but quadri- lateral if "equiangular" is to have any sense. Though, by introducing Tcrpdir\tvpav for any quadrilateral, Euclid enabled ambiguity to be avoided, there seem to be traces of the older vague use of Ttrpdytowr in much later writers. Thus Heron (Def. 100) speaks of a cube as "contained by six equi- lateral and equiangular Ttrpdywra" and Proclus (p. 166, 10) adds to his remark about the " four-sided triangle " that " you might have rtrpaybiva with more than the four sides," where rtrpayvsva can hardly mean squares. tTtpofiyKn, oblong (with sides of different length), is also a Pythagorean term. The word right-angled {ipBoyiiytor) as here applied to quadrilaterals must mean rectangular (i.e., practically, having all its angles right angles) ; for, although it is tempting to take the word in the same sense for a I. Def. u] NOTES ON DEFINITIONS so— a j 189 square as for a triangle (i.e. " having one right angle "), this will not do in the case of the oblong, which, unless it were stated that three of its angles are right angles, would not be sufficiently defined. If it be objected, as it was by Todhunter for example, that the definition of a square assumes more than is necessary, since it is sufficient that, being equilateral, it should have one right angle, the answer is that, as in other cases, the superfluity does not matter from Euclid's point of view ; on the contrary, the more of the essential attributes of a thing that could be included in its definition the better, provided that the existence of the thing defined and its possession of all those attributes is proved before the definition is. actually used ; and Euclid does this in the case of the square by construction in 1. 46, making no use of the definition before that proposition. The word rhombus (po/t/Jot) is apparently derived from fiipfioi, to turn round and round, and meant among other things a spinning-top. Archimedes uses the term solid rhombus to denote a solid figure made up of two right cones with a common circular base and vertices turned in opposite directions. We can of course easily imagine this solid generated by spinning; and, if the cones were equal, the section through the common axis would be a plane rhombus, which would also be the apparent form of the spinning solid to the eye. The difficulty in the way of supposing the plane figure to have been named after the solid figure is that in Archimedes the cones forming the solid are not necessarily equal. It is however possible that the solid to which the name was originally given was made up of two equal cones, that the plane rhombus then received its name from that solid, and that Archimedes, in taking up the old name again, extended its signification (cf. J. H. T. Miiller, Beitrdge zur Terminologie der griechisehen Mathematiker, i860, p. 20). Proclus, while he speaks of a rhombus as being like a shaken, i.e. deformed, square, and of a rhomboid as an oblong that has been moved, tries to explain the rhombus by reference to the appearance of a spinning square {rsrpaymvov poftfioiittvor). It is true that the definition of a rhomboid says more than is necessary in describing it as having its opposite sides and angles equal to one another. The answer to the objection is the same as the answer to the similar objection to the definition of a square. Euclid makes no use in the Elements of the oblong, the rhombus and the rhomboid. The explanation of his inclusion of definitions of these figures is no doubt that they were taken from earlier text- books. From the words "let quadrilaterals other than these be called trapezia" we may perhaps infer that trapezium was a new name or a new application of an old name. As Euclid has not yet defined parallel lines and does not anywhere define a parallelogram, he is not in a position to make the more elaborate classification of quadrilaterals attributed by Proclus to Posidonius and appearing also in Heron's Definitions. It may be shown by the following diagram, distinguishing seven species of quadrilaterals. Quadrilaterals parallelograms non- parallelograms rectangular non-rectangular two sides parallel no sides parallel [traptxtum] {traprt&t) r-U squart oblong rhombus rhomboid i iGscdts trapezium scaitnt trapezium 190 BOOK I [1. Bepf. 2s, a3 It will be observed that, while Euclid in the above definition classes as trapezia all quadrilaterals other than squares, oblongs, r ho in hi, and rhomboids, the word is in this classification restricted to quadrilaterals having two sides (only) parallel, and trapezoid is used to denote the rest Euclid appears to have used trapezium in the restricted sense of a quadrilateral with two sides parallel in his book, vtpi Suuptatav (on divisions of figures). Archimedes uses it in the same sense, but in one place describes it more precisely as a trapezium with its two sides parallel. Definition 23. IlapaX^7]\ai turiv ciStitu, a*riMf iv to) uurw hrtwi&o ov&at mil «*/3aXAo/j.*i'ai tit irnpov itji' imiTtpa. Ta /Mpf in ftrfiirtpa tni/iviirTOtxTiv &XXykiuf. Parallel straight tines are straight lines which, being in the same plane and being produced indefinitely in both directions, do not meet one another in either direction. Wap6XKvj\\K (alongside one another) written in one word does not appear in Plato ; but with Aristotle it was already a familiar term. «(? aire tpok cannot be translated " to infinity " because these words might seem to suggest a region or place infinitely distant, whereas •« airttpoy, which seems to be used indifferently with «jt' Zirnpuv, is adverbial, meaning "without limit," i.e. "indefinitely." Thus the expression is used of a magnitude being "infinitely divisible," or of a series of terms extending without limit in both directions, i tWtiM i»' aiMipw /ijrtnTMif, c£ a ,r< i,ij^ (tcrtv ai toJv Buo op&uv eX<£crf/in'€!. TXrt/, if a straight line falling on two straight lints make the interior angles on the same side less than two right angles, the two straight lines, if produeed indefinitely, meet on that side on whieh are the angles less than the tit, at K. Then, since the angles AFG, CGFsze. less than two right angles, and the angles AFG, GFB are equal to two right angles, take away the common angle AFG, and the angle CGF is less than the angle BFG; that is, the exterior angle of the triangle KFG is less than the interior arid opposite angle BFG : which is impossible. Therefore AB, CD do not meet towards B, D. (v) But they do meet, and therefore they must meet in one direction or the other: therefore they meet towards A, B, that is, on the side where are the angles less than two right angles. The flaw in Ptolemy's argument is of course in the part of his proof of i. 2 9 which I have italicised. As Proclus says, he is not entitled to assume that, if AB, CD are parallel, whatever is true of the interior angles on one side of FG (i.e. that they are together equal to, greater than, or less than, two right angles} is necessarily true at the same time of the interior angles on the other side. Ptolemy justifies this by saying that FA, GC are no more parallel in one direction than FB, GD are in the other : which is equivalent to the assumption that through any point only one parallel can be drawn to a given straight line. That is, he assumes an equivalent of the very Postulate he is endeavouring to prove. Proclus. Before passing to his own attempt at a proof, Proclus (p. 368, 26 sqq.) examines an ingenious argument (recalling somewhat the famous one about Achilles and the tortoise) which appeared to show that it was impossible for the lines described in the Postulate to meet. Let AB, CD make with AC the angles BAC, ACD together less than two right angles. Bisect AC a.t E and along AB, CD respectively measure AF, CG so that each is equal to AE. gl \ |h Bisect FG at K and mark off FK, GL each equal to FH; and so on. Then AF, CG will not meet at any point on FG ; for, if that were the case, two sides of a triangle would be together equal to the third : which is impossible. 1. Post. 5] NOTE ON POSTULATE 5 *>7 Similarly, AB, CD will not meet at any point on KL; and "proceeding like this indefinitely, joining the non-coincident points, bisecting the lines so drawn, and cutting off from the straight lines portions equal to the half of these, they say they thereby prove that the straight lines AB, CD will not meet anywhere." It is not surprising that Proclus does not succeed in exposing the fallacy here (the fact being thai the process will indeed be endless, and yet the straight lines will intersect within a finite distance). But Proclus' criticism contains nevertheless something of value. He says that the argument will prove too much, since we have only to join A G in order to see that straight lines making svme angles which are together less than two right angles do in fact meet, namely AG, CG. "Therefore it is not possible to assert, without some definite limitation, that the straight lines produced from angles less than two right angles do not meet. On the contrary, it is manifest that MM straight lines, when produced from angles less than two right angles, do meet, although the argument seems to require it to be proved that this property belongs to all such straight lines. For one might say that, the lessening of the two right angles being subject to no limitation, with such and such an amount of lessening the straight tines remain non-secant, but with an amount of lessening in excess of this they meet (p. 371, 2 — 10)." [Here then we have the germ of such an idea as that worked out by Lobachewsky, namely that the straight lines issuing from a point in a plane can be divided with reference to a straight line lying in that plane into two classes, "secant" and "non-secant," and that we may define as parallel the two straight lines which divide the secant from the non-secant class.] Proclus goes on (p. 371, io) to base his own argument upon "an axiom such as Aristotle too used in arguing that the universe is finite. For, if from one point two straight lines forming an angle be produced indefinitely, the distance (StaiTTairit, Arist. htaimjjta) between the said straight tines produced indefinitely will exceed any finite magnitude. Aristotle at all events showed that, if the Straight lines drawn from the centre to the circumference are infinite, the interval between them is infinite. For, if it is finite, it is impossible to increase the distance, so that the straight lines {the radii) are not infinite. Hence the straight lines, when produced indefinitely, will be at a distance from one another greater than any assumed finite magnitude." This is a fair representation of Aristotle's argument in De caelo ]. 5, 271 b 28, although of course it is not a proof of what Proclus assumes as an axiom. This being premised, Proclus proceeds (p. 371, 24): I. " I say that, if any straight line cuts one of two parallels, it will cut the other also. "For let AB, CD be parallel, and let EFG cut AB; I say that it will cut CD also. " For, since BF, FG are two straight lines from .E one point F, they have, when produced indefinitely, ft X^ B a distance greater than any magnitude, so that it will ^\ also be greater than the interval between the parallels. " Whenever therefore they are at a distance from one another greater than the distance between the parallels, FG will cut CD. " Therefore etc." 2o8 BOOK I [i. Post. 5 II. " Having proved this, we shall prove, as a deduction from it, the theorem in question. "For let AB, CD be two straight lines, and let EF falling on them make the angles BEF, DFE less than two right angles. "I say that the straight lines will meet on that side on which are the angles less than two right angles. " For, since the angles BEF, DFE are less than two right angles, let the angle HEB be equal to the excess of two right angles (over them), and let HE be produced to K. " Since then EF falls on KH, CD and makes the two interior angles HEF, DFE equal to two right angles, the straight lines MX, CD are parallel. "And AB cuts KH\ therefore it will also cut CD, by what was before shown. " Therefore AB, CD will meet on that side on which are the angles less than two right angles. " Hence the theorem is proved." Clavius criticised this proof on the ground that the axiom from which it starts, taken from Aristotle, itself requires proof. He points out that, just as you cannot assume that two lines which continually approach one another will meet (witness the hyperbola and its asymptote), so you cannot assume that two lines which continually diverge will ultimately be so far apart that a perpendicular from a point on one let fall on the other will be greater than any assigned distance ; and he refers to the conchoid of Nkomedes, which continually approaches its asymptote, and therefore continually gets farther away from the tangent at the vertex • yet the perpendicular from any point on the curve to that tangent will always be less than the distance between the tangent and the asymptote. Saccheri supports the objection. Proclus' first proposition is open to the objection that it assumes that two "parallels" (in the Euclidean sense) or, as we may say, two straight liius which have a common perpendicular, are (not necessarily equidistant, but) so related that, when they are produced indefinitely, the perpendicular from a point of one upon the other remains finite. This last assumption is incorrect on the hyperbolic hypothesis ; the "axiom" taken from Aristotle does not hold on the elliptic hypothesis, Nasiraddln at.-TusI. The Persian-bom editor of Euclid, whose date is 1201— 1274, has three lemmas leading up to the final proposition. Their content is substantially as follows, the first lemma being apparently assumed as evident I. (o) If AB, CD be two straight lines such that successive perpen- diculars, as EF, GH, KL, from points on AB to CD always make with A3 unequal angles, which are always acute on the side towards B and always obtuse on the side towards A, then the lines AB, CD, so long as they do not cut, approach continually nearer in the direction of the acute angles and diverge continually in the direction of the obtuse angles, and the perpendiculars diminish towards B, D, and in- crease towards A, C. (6) Conversely, if the perpendiculars so drawn L H F C continually become shorter in the direction of B, D, and longer in the i. Post. 3] NOTE ON POSTULATE 5 zog direction of A, C, the straight lines AB, CD approach continually nearer in the direction of B, D and diverge continually in the other direction ; also each perpendicular will make with AB two angles one of which is acute and the other is obtuse, and all the acute angles will lie in the direction towards B, D, and the obtuse angles in the opposite direction. [Saccheri points out that even the first part {a) requires proof. As regards the converse (b) he asks, why should not the successive acute angles made by the perpendiculars with AB, while remaining acute, become greater and greater as the perpendiculars become smaller until we arrive at last at a perpendicular which is a common perpendicular to both lines? If that happens, alt the author's efforts are in vain. And, if you are to assume the truth of the statement in the lemma without proof, would it not, as Wall is said, be as easy to assume as axiomatic the statement in Post. 5 without more ado?] II. ^AC, BI) be drawn from the extremities of AB at right angles to it and on the same side, and if AC, EDfc made equal to one another and CD be joined, each of the angles ACD, BDC will be right, and CD will be equal to AB, The first part of this lemma is proved by redwtio ad absurdum from the preceding lemma. If, e.g., the angle A CD is not right, it must either be acute or obtuse. Suppose it is acute ; then, by lemma 1, A C is greater than BD, which '\s contrary to the hypothesis. And so on. The angles ACD, BDC being proved to be right angles, it is easy to prove that AB, CD are equal. [It is of course assumed in this " proof" that, if the angle ACD is acute, the angle BDC is obtuse, and vice versa.] III. /// any triangle the three angles are together equal to two right angles. This is proved for a right-angled triangle by means of the foregoing lemma, the four angles of the quadrilateral ABCD of that lemma being all right angles. The proposition is then true for any triangle, since any triangle can be divided into two right-angled triangles IV. Here we have the final " proof " of Post. 5. Three cases are distinguished, 'but it is enough to show the case where one of the interior angles is right and the other acute. Suppose AB, CD to be two straight lines met by FCE making the angle ECD a right angle and the angle CEB an acute angle. 'lake any point G on EB, and draw GH perpendicular to EC. Since the angle CEG is acute, the perpendicular GH will fall on the side of E towards D, and will either coincide with CD or not coincide with it. In the former case the proposition is proved. If GH does not coincide with CD but falls on the side of it towards F, CD, being within the triangle formed by the perpendicular and by CE, EG, must cut EG. [An axiom is here used, namely that, if CD be produced far enough, it must pass outside the triangle and therefore cut some side, which must be EB, since it cannot be the perpendicular (1. 27), or CE.) Lasdy, let C.fffall on the side of CD towards E. aro BOOK I [i. Post. 5 Along HC set off HK, KL etc., each equal to EH, until we get the first point of division, as M, beyond C. Along GB set off GN, NO etc, each equal to EG, until EP is the same multiple of EG that EM is of EH. Then we can prove that the perpendiculars from N, 0, P on EC fall on the points K, L, M respectively. For take the first perpendicular, that from N, and call it NS. Draw EQaa right angles to EH and equal to GH, and set off SR along SiValso equal to GH, Join QG, GR, Then (second lemma) the angles EQG, QGHaie right, and QG = EH. Similarly the angles SRG, RGHwe right, and RG-SH Thus RGQ is one straight line, and the vertically opposite angles NGR, EGQ are equal. The angles NRG, EQG are both right, and NG = GE, by construction. Therefore (1. 26") EG =■ GQ ; whence SH= HE = KH, and S coincides with K. We may proceed similarly with the other perpendiculars. Thus PM is perpendicular to EE. Hence CD, being parallel to MP and within the triangle PME, must cut EP, if produced far enough. John Wallis. As is well known, the argument of Wallis (1616—1703) assumed as a postulate that, given a figure, another figure is possible which is similar to (he given one and of any sine whatever. In fact Wallis assumed this for triangles only. He first proved (1) that, if a finite straight line is placed on an infinite straight line, and is then moved in its own direction as far as we please, it will always lie on the same infinite straight line, (2) that, if an angle be moved so that one leg always slides along an infinite straight line, the angle will remain the same, or equal, (3) that, if two straight lines, cut by a third, make the interior angles on the same side less than two right angles, each of the exterior angles is greater than the opposite interior angle (proved by means of 1. 13). p x (4) UAB, CD make, with AC, the interior angles less than two right angles, suppose AC (with AB rigidly attached to it) to move along _ ^ \- g AF to the position ay, such that a coincides with C. If AB then takes the position aft, o£ lies entirely outside CD (proved by means of {3) above). (5) With the same hypotheses, the straight line aft, sr AB, during Us motion, and before a reaches C, must cut the straight tine CD. S6) Here is enunciated the postulate stated above. 7-) Postulate S is now proved thus. Let AB, CD be the straight lines which make, with the infinite straight line ACF meeting them, the interior angles BA C, DC A together less than two right angles. Suppose AC (with AB rigidly attached to it) to move along ACF until AB takes the position of aft cutting CD in it. Then, «CVr being a triangle, we can, by the above postulate, suppose a triangle drawn on the base CA similar to the triangle aCVr. Let it be ACF. [Wallis here interposes a defence of the hypothetical construction.] 1. Post. 5] NOTE ON POSTULATE 5 sit Thus CP and AP meet at P; and, as by the definition of similar figures the angles of the triangles PCA, rCa are respectively equal, the angle PCA being equal to the angle rCa and the angle PAC to the angle miCor BAC, it follows that CP, APMe on CD, A3 produced respectively. Hence AB, CD meet on the side on which are the angles less than two right angles. [The whole gist of this proof lies in the assumed postulate as to the existence of similar figures ; and, as Saccheri points out, this is equivalent to unconditionally assuming the "hypothesis of the right angle," and consequently Euclid's Postulate 5.] Gerolamo Saccheri. The book Euclides ab omni naevo vindicatus (1733) by GeTolamo Saccheri {1667 — 1733), a Jesuit, and professor at the University of Pavia, is now accessible (1) edited in German by Engel and Stackel, Die Theorie dtr Parallellinien von Euhtid bis auf Gauss, 1895, pp. 41 — 136, and (2) in an Italian version, abridged but annotated, L'Euclide emendato del P. Gerolamo Saccheri, by G. Boccardini (Hoepli, Milan, 1904}. It is of much greater importance than all the earlier attempts to prove Post. 5 because Saccheri was the first to contemplate the possibility of hypotheses other than that of Euclid, and to work out a number of consequences of those hypotheses. He was therefore a true precursor of Legendie and of Lobachewsky, as Beltrami called him (1889), and, it might be added, of Riemann also. For, as Veronese observes {Fondamenti di geometria, p, 570), Saccheri obtained a glimpse of the theory of parallels in all its generality, while Legendre, Lobachewsky and G, Bolyai excluded a priori, without knowing it, the "hypo- thesis of the obtuse angle," or the Riemann hypothesis. Saccheri, however, was the victim of the preconceived notion of his time that the sole possible geometry was the Euclidean, and he presents the curious spectacle of a man laboriously erecting a structure upon new foundations for the very purpose of demolishing it afterwards ; he sought for contradictions in the heart of the systems which he constructed, in order to prove thereby the falsity of his hypotheses. For the purpose of formulating his hypotheses he takes a plane quadri- lateral ABDC, two opposite sides of which, A C, BD, are equal and perpendicular to a third AB. Then the angles at C and D are easily proved to be equal. On the Euclidean hypothesis they are both right angles; but apart from this hypothesis they might be both obtuse or both acute. To the three possibilities, whicfc Saccheri distinguishes by the names (1) the hypothesis of the right angle, (i) the hypothesis of the obtuse angle and (3) the hypothesis of the acute angle respectively, there corresponds a certain group of theorems ; and Saccheri's point of view is that the Postulate will be completely proved if the consequences which follow from the last two hypotheses comprise results inconsistent with one another. Among the most important of his propositions are the following : (1) If the hypothesis of the right angle, or of the obtuse angle, or of the acute angle is proved true in a single case, it is true in every other case. (Props, v., VI., VII.) (2) According as the hypothesis of the right angle, the obtuse angle, or the acute angle is true, the sum of the thru angles of a triangle is equal to, greater than, or less than two right angles. (Prop. i\. ) lit BOOK I [i. Post. 5 (3) From the existence of a single triangle in which the sum of the angles is equal to, greater than, or less than two right angles (lie truth of the hypothesis of the right angle, obtuse angle, or acute angle respectively follows. (Prop, xv.) These propositions involve the following : If in a single triangle the sum of the angles is equal to, greater than, or less than two right angles, then any triangle has the sum of its angles equal to, greater than, or less than tlt'O right angles respectively, which was proved about a century later by Legendre for the two cases only where the sum is equal to or less than two right angles. The proofs are not free from imperfections, as when, in the proofs of Prop. xii. and the part of Prop. xm. relating to the hypothesis of the obtuse angle, Saccheri uses Eucl. 1. 18 depending on 1. 16, a proposition which is only valid on the assumption that straight lines are infinite in length ; for this assumption itself does not hold under the hypothesis of the obtuse angle (the Riemann hypothesis). The hypothesis of the acute angle takes Saccheri much longer to dispose of, and this part of the book is less satisfactory ; but it contains the following propositions afterwards established anew by Lobachewsky and Bolyai, viz. : (4) Two straight lines in a plane {even on the hypothesis of the acute angle) either have a common perpendicular, or must, if produced in one and the same direction, either intersect once at a finite distance or at least continually approach one another. {Prop, xxin.) (5) In a cluster of rays issuing from a point there exist always (on the hypothesis of the acute angle) two determinate straight lines which separate the straight lines which intersect a fixed straight line from those which do not intersect it, ending with and including the straight line which has a common perpendicular with the fixed straight line. (Props. XXX., xxxc, xxxii.) Lambert. A dissertation by G.S. Kliigel, Conatuum praecipuorum tlteoriamparallelarnm demonstrandi recensio (1 763), contained an examination of some thirty " demon- strations" of Post. 5 and is remarkable for its conclusion expressing, apparently for the first time, doubt as to its demenstrability and observing that the certainty which we have in us of the truth of the Euclidean hypothesis is not the result of a series of rigorous deductions but rather of experimental observations. It also had the greater merit that it called the attention of Johann Heinrich Lambert (1728—1777) to the theory of parallels. His Theory of Parallels was written in 1766 and published after his death by G. Bernoulli and C. F. Hindenburg ; it is reproduced by Engel and Stackel (op. sit. pp. 151 — 208). The third part of Lambert's tract is devoted to the discussion of the same three hypotheses as Saccheri's, the hypothesis of the right angle being for Lambert the first, that of the obtuse angle the second, and that of the acute angle the third, hypothesis; and, with reference to a quadrilateral with three right angles from which Lambert starts (that is, one of the halves into which the median divides Saccheri's quadrilateral), the three hypotheses are the assumptions that the fourth angle is a right angle, an obtuse angle, or an acute angle respectively. Lambert goes much further than Saccheri in the deduction of new propositions from the second and third hypotheses. The most remarkable is the following. The area of a plane triangle, under the second and third hypotheses, is proportional to the difference between the sum vf the three angles and two right angles. h Post. 5] NOTE ON POSTULATE 5 «3 Thus the numerical expression for the area of a triangle is, under the third hypothesis &.=;A( r -A-£-C) (1), and under the second hypothesis &. = *(A + £+C-*) (a), where A is a positive constant A remarkable observation is appended (5 82) : "In connexion with this it seems to be remarkable that the second hypothesis holds if spherical instead of plane triangles are taken, because in the former also the sum of the angles is greater than two right angles, and the excess is proportional to the area of the triangle. " It appears still more remarkable that what I here assert of spherical triangles can be proved independently of the difficulty of parallels. This discovery that the second hypothesis is realised on the surface of a sphere is important in view of the development, later, of the Riemann hypothesis (1854). Still more remarkable is the following prophetic sentence : " I am almost inclined to draw the conclusion that the third hypothesis arises with an imaginary spherical surface" (cf. Lobachewsky's Gcome'trie imaginaire, 1837). No doubt Lambert was confirmed in this by the fact that, in the formula (1) above, which, for h = r 1 , represents the area of a spherical triangle, if r V- 1 is substituted for r, and r 1 = k, we obtain the formula (1). Legend re. No account of our present subject would be complete without a full reference to what is of permanent value in the investigations of Adrien Marie Legendre {1752 — 1833) relating to the theory of parallels, which extended over the space of a generation. His different attempts to prove the Euclidean hypothesis appeared in the successive editions of his aliments de Giomklrie from the first {1794) to the twelfth {1823), which last may be said to contain his last word on the subject. Later, in 1833, he published, in the Afhnoires de I'Acadimie Royals des Sciences, xn. p. 367 sqq., a collection of his different proofs under the title Reflexions sur dffirentes maniires de dhnontrer la thiorie des paralteles. His exposition brought out clearly, as Saccheri had done, and kept steadily in view, the essential connexion between the theory of parallels and the sum of the angles of a triangle. In the first edition of the Elements the proposition that the sum of the angles of a triangle is equal to two right angles was proved analytically on the basis of the assumption that the choice of a unit of length does not affect the correctness of the proposition to be proved, which is of course equivalent to Wallis' assumption of the existence of similar figures. A similar analytical proof is given in thj notes to the twelfth edition. In his second edition Legendre proved Postulate 5 by means of the assumption that, given three points not in a straight line, there exists a circle passing through all three. In the third edition (1800) he gave the proposition that the sum of the angles of a triangle is not greater than two right angles ; this proof, which was geometrical, was replaced later by another, the best known, depending on a construction like that of Euclid 1. 16, the continued application of which enables any number of successive triangles to be evolved in which, while the sum of the angles in each remains always equal to the sum of the angles of the original triangle, one of the angles increases and the sum of the other two diminishes continually. But Legendre found the proof of the equally necessary proposition that the sum of the angles of a triangle is ai 4 BOOK I [i. Post, S not less than two right angles to present great difficulties. He first observed that, as in the case of spherical triangles (in which the sum of the angles is greater than two right angles) the excess of the sura of the angles over two right angles is proportional to the area of the triangle, so in the case of rectilineal triangles, if the sum of the angles is less than two right angles by a Mrtain deficit, the deficit will be proportional to the area of the triangle. Hence if, starting from a given triangle, we could construct another triangle in which the original triangle is contained at least m times, the deficit of this new triangle will be equal to at least m times that of the original triangle, so that the sum of the angles of the greater triangle will diminish progressively as m increases, until it becomes zero or negative : which is absurd. The whole difficulty was thus reduced to that of the construction of a triangle containing the given triangle at least twice ; but the solution of even this simple problem requires it to be assumed (or proved) that through a given point within a given angle less than two-thirds of a right angle we can always draw a straight tine which shall meet both sides of the angle. This is however really equivalent to Euclid's Postulate. The proof in the course of which the necessity for the assumption appeared is as follows. It is required to prove that the sum of the angles of a triangle cannot be less than two right angles. Suppose A is the least of the three angles of a triangle ABC. Apply to the opposite side 2?C a triangle DBC, equal to the triangle ACB, and such that the angle DBC is equal to the angle ACB, and the angle DCB to the angle ABC ; and draw any straight line through D cutting AB, AC produced in E, F. If now the sum of the angles of the triangle ABC is less than two right angles, being equal to aB-i say, the sum of the angles of the triangle DBC, equal to the triangte ABC, is also 2^-8. Since the sum of the three angles of the remaining triangles DEB, FDC respectively cannot at all events be greater than two right angles [for I>egendre's proofs of this see below], the sum of the twelve angles of the four triangles in the figure cannot be greater than 4B + {2B - &) + (2B - ty, i.e. %R-al. Now the sum of the three angles at each of the points B, C, D is iR. Subtracting these nine angles, we have the result that the three angles of the triangle AEF cannot be greater than 2R - 28. Hence, if the sum of the angles of the triangle ABC is less than two right angles by £, the sum of the angles of the- larger triangle AEF is less than two right angles by at least 28. We can continue the construction, making a still larger triangle from AEF, and so on. But, however small 8 is, we can arrive at a multiple 2*$ which shall exceed any given angle and therefore tR itself; so that the sum of the three angles of a triangle sufficiently large would be zero or even less than zero : which is absurd. Therefore etc. The difficulty caused by the necessity of making the above-mentioned assumption made Legendre abandon, in his ninth edition, the method of the i. Post. 5] NOTE ON POSTULATE 5 115 editions from the third to the eighth and return to Euclid's method pure and simple. But again, in the twelfth, he returned to the plan of constructing any number of successive triangles such that the sum of the three angles in all of them remains equal to the sum of the three angles of the original triangle, but two of the angles of the new triangles become smaller and smaller, while the third becomes larger and larger ; and this time he claims to prove in one proposition that the sum of the three angles of the original triangle is equal to two right angles by continuing the construction of new triangles indefinitely and compressing the two smaller angles of the ultimate triangle into nothing, while the third angle is made to become a flat angle at the same time. The construction and attempted proof are as follows. Let ABC be the given triangle ; let A B be the greatest side and BC the least ; therefore C is the greatest angle and A the least. From A draw AD to the middle point of BC, and produce AD to C, making AC equal to AB, Produce AB to B 1 , making AB equal to twice AD. The triangle ABC is then such that the sum of its three angles is equal to the sum of the three angles of the triangle ABC. For take AK along AB equal to AD, and join C'K. Then the triangles ABD, ACK havt two sides and the included angles respectively equal, and are therefore equal in all respects ; and C'K is equal to BD or DC. Next, in the triangles BCK, A CD, the angles BKC, ADC are equal, being respectively supplementary to the equal angles AKC, ADB; and the two sides about the equal angles are respectively equal ; therefore the triangles BC'K, A CD are equal in all respects. Thus the angle AC'B is the sum of two angles respectively equal to the angles B, C of the original triangle ; and the angle A in the original triangle is the sum of two angles respectively equal to the angles at A and B' in the triangle ABC. It follows that the sum of the three angles of the new triangle ABC is equal to the sum of the angles of the triangle ABC. Moreover, the side AC, being equal to AB, and therefore greater than AC, is greater than BC which is equal to AC. Hence the angle C'AB'w less than the angle ABC ; so that the angle CAB is less than \A, where A denotes the angle CAB of the original triangle. [It will be observed that the triangle ABC is really the same triangle as the triangle ABB obtained by the construction of Eucl. 1. 16, but differently placed so that the longest side lies along AB.] By taking the middle point D of the side BC and repeating the same construction, we obtain a triangle AB'C" such that (1) the sum of its three angles is equal to the sum of the three angles of ABC, (a) the sum of the ai6 BOOK I [1. Post. 5 two angles CAB", AB"C" is equal to the angle CAB in the preceding triangle, and is therefore less than \A, and (3) the angle CAB' is less than half the angle CAB, and therefore less than \A. Continuing in this way, we shall obtain a triangle Abe such that the sum of two angles, those at A and i, is less than — A, and the angle at c is greater than the corresponding angle in the preceding triangle. If, Legendre argues, the construction be continued indefinitely so that - n A becomes smaller than any assigned angle, the point c ultimately ties on Alt, and the sum of the three angles of the triangle (which is equal to the sum of the three angles of the original triangle) becomes identical with the angle at c, which is then a.Jlat angle, and therefore equal to two right angles. This proof was however shown to be unsound (in respect of the final inference) by J. P. W. Stein in Gergonne's Annaln de Mathimatiques XV., 1824, pp. 77—79. We will now reproduce shortly the substance of the theorems of Legendre which are of the most permanent value as not depending on a particular hypothesis as regards parallels. I. The sum of the three angles of a triangle cannot it greater than two right angles. This Legendre proved in two ways. (r) Mr si proof (in the third edition of the Aliments). Let ABC be the given triangle, and ACf a straight line. Make CE equal to AC, the angle DCE equal to the angle BAC, and DC equal to AB. Join DE. Then the triangle DCE is equal to the triangle BAC in all respects. If then the sum of the three angles of the triangle ABC is greater than %R, the said sum must be greater than the sum of the angles BCA, BCD, DCE, which sum is equal to 2H. Subtracting the equal angles on both sides, we have the result that the angle ABC is greater than the angle BCD. But the two sides AB, BC of the triangle ABC are respectively equal to the two sides DC, CB of the triangle BCD. Therefore the base AC is greater than the base BD (Eucl. 1. 14). Next, make the triangle BEG {by the same construction) equal in all respects to the triangle BAC or DCE ; and we prove in the same way that CE (or AC) is greater than DE. And, at the same time, BD is equal to DE, because the angles BCD, DEE are equal. Continuing the construction of further triangles, however small the difference between AC and BD is, we shall ultimately reach some multiple i. Post. 5] NOTE ON POSTULATE 5 ai7 of this difference, represented in the figure by (say) the difference between the straight line AJ and the composite line BDFHK, which will be greater than any assigned length, and greater therefore than the sum of AB and JJC. Hence, on the assumption that the sum of the angles of the triangle ABC is greater than 2R, the broken line ABDFHKJ may be less than the straight Hne AJ: which is impossible. Therefore etc. (2) Proof substituted later. If possible, let 2^ + a be the sum of the three angles of the triangle ABC, of which A is not greater than either of the others. Bisect BC at H, and produce AH to D, making HD equal to AH ; join BD. Then the triangles AHC, DHB are equal in all respects (l. 4); and the angles CAH,ACHaie respectively equal to the angles BDH, DBH. It follows that the sum of the angles of the triangle ABD is equal to the sum of the angles of the original triangle, i.e. to tR + a. And one of the angles DAB, ADB is either equal to or less than half the angle. CAB. Continuing the same construction with the triangle ADB, we find a third triangle in which the sum of the angles is still zR + a, while one of them is equal to or less than J L CAB)jj\. Proceeding in this way, we arrive at a triangle in which the sum of the angles is 2R + a, and one of them is not greater than ( L CAB)J2 H . And, if n is sufficiently large, this will be less than a. ; in which case we should have a triangle in which two angles are together greater than two right angles : which is absurd. Therefore a is equal to or less than zero. (It will be noted that in both these proofs, as in Eucl. 1. 16," it is taken for granted that a straight line is infinite in length and does not return into itself, which is not true under the Riemann hypothesis.) II. On the assumption that the sum of the angles of a triangle is less than two right angles, if a triangle is made up of two others, the " deficit" of ' tht, former is equal to the sum of the " deficits " of the others. In fact, if the sums of the angles of the component triangles are 2R -a, 2R - fi respectively, the sum of the angles of the whole triangle is (zj¥-a) + (2JP-£)-3.ff = 7Je-fa + P). III. If the sum of the three angles of a triangle is equal to two right angles, the same is true of all triangles obtained by subdividing it by straight lines drawn from a vertex to meet the opposite side. Since the sum of the angles of the triangle ABC is equal to 2R, if the sum of the angles of the triangle ABD were 2R - a, it would follow that the sum of the angles of the triangle A ADC must be »R + q, which is absurd (by I. above). IV. If in a triangle the sum of the three angles is equal to two right angles, a quadrilateral can always be constructed with four right angles and four equal sides B^ £2 io exceeding in length any assigned rectilineal segment. Let ABC be a triangle in which the sum of the angles is equal to two «i8 BOOK I [i. Post. 5 right angles. We can assume ABC to be an isosceles right-angled triangle because we can reduce the case to this by making subdivisions of ABC by straight lines through vertices (as in Prop. III. above). Taking two equal triangles of this kind and placing their hypotenuses together, we obtain a quadrilateral with four right angles and four equal sides. Putting four of these quadrilaterals together, we obtain a new quadrilateral 0/ the same kind but with its sides double of those of the first quadrilateral. After n such operations we have a quadrilateral with four right angles and four equal sides, each being equal to 3" times the side AH. The diagonal of this quadrilateral divides it into two equal isosceles right- angled triangles in each of which the sum of the angles is equal to two right angles. Consequently, from the existence ot one triangle in which the sum of the three angles is equal to two right angles it follows that there exists an isosceles right-angled triangle with sides greater than any assigned rectilineal segment and such that the sum of its three angles is also equal to two right angles. V. If the sum of the three angles of one triangle is equal to two right angles, the sum of the three angles of any other triangle is also equal to two right angles. It is enough to prove this for a right-angled triangle, since any triangle can be divided into two right-angled triangles. Let ABC be any right-angled triangle. If then the sum of the angles of any one triangle is equal to two right angles, we can construct (by the preceding Prop.) an isosceles right-angled triangle with the same property and with its perpendicular sides, greater than those of ABC. Let A'BC' be such a triangle, and let it be applied to ABC, as in the figure. Applying then Prop. 111. above, we deduce first that the sum of the three angles of the triangle ABC is equal to two right angles, and next, for the same reason, that the sum of the three angles of the original triangle A BC is equal to two right angles. VI. If in any one triangle the sum of the three angles is less than two right angles, the sum of the three angles of any other triangle is also less than two right angles. This follows from the preceding theorem. (It will be observed that the last two theorems are included among those of Saccheri, which contain however in addition the corresponding theorem touching the case where the sum of the angles is greater than two right angles.) We come now to the bearing of these propositions upon Euclid's Postulate 5 ; and the next theorem is VII. If the sum of the three angles of a triangle is equal to two right angles, through any point in a plane there (an only be drawn one parallel to a given straight line. r. Post. 5] NOTE ON POSTULATE 5 219 For the proof of this we require the following Lemma. // is always possible, through a faint P, to draw a straight line which shall make, with a ghten straight line (r), an angle less than any assigned angle. Let Q be the foot of the perpendicular from /'upon r. Let a segment QR be taken on r, on either side of Q, such that QR is equal to PQ. Join PR, and mark off the segment RR' equal to PR ; join PR'. If tu represents the angle QPR or the angle QRP, each of the equal angles RPR', RR'P is not greater than 10/ 1. Continuing the construction, we obtain, after the requisite number of operations, a triangle PR,-, R n in which each of the equal angles is equal to or less than ya\yth Tok tt>£\uvs) and again, b 14, of "the notion (or consciousness) that they are pleased at his good fortune." It is true that Plato and Aristotle do not use the word in a technical sense ; but neither was there apparently in Aristotle's time any fixed technical term for what we call "axioms," since he speaks of them variously as "the so-called axioms in mathematics," "the so- called common axioms," " the common (things) " (to. jcotpd), and even " the common opinions " (kmvoi bofru). I see therefore no reason why Euclid should not himself have given a technical sense to " Common Notions," which is at least a distinct improvement upon " common opinions." (3) The use of fcroux in Proclus' quotation from Apollonius seems to me to be an unfortunate, rather than a fortunate, coincidence from Tannery's point of view, for it is there used precisely in the old sense of the notion of an object (in that case a line). No doubt it is difficult to feel certain that Euclid did himself use the term Common Notions, seeing that Proclus' commentary generally speaks of Axioms, But even Proclus (p. 194, 8), after explaining the meaning of the word "axiom," first as used by the Stoics, and secondly as used by "Aristotle and »** BOOK I [!. C. N. i the geometers," goes on to say : " For in their view (that of Aristotle and the geometers) axiom and common notion are the same thing." This, as it seems to me, may be a sort of apology for using the word " axiom " exclusively in what has gone before, as if Proclus had suddenly bethought himself that he had described both Aristotle and the geometers as using the one term " axiom," whereas he should have said that Aristotle spoke of " axioms," while "the geometers" (in fact Euclid), though meaning the same thing, called them Common Notions, It may be for a like reason that in another passage (p. 76, 16), after quoting Aristotle's view of an "axiom," as distinct from a postulate and a hypothesis, he proceeds : " For it is not by virtue of a common notion that, without being taught, we preconceive the circle to be such and such a figure." If this view of the two passages just quoted is correct, it would strengthen rather than weaken the case for the genuineness of Common Notions as the Euclidean term. Again, it is clear from Aristotle's allusions to the " common axioms in mathematics " that more than one axiom of this kind had a place in the text- books of his day ; and as he constantly quotes the particular axiom that, if equals be taken from equals, the remainders are equals which is Euclid's Common Notion 3, it would seem that at least the first three Common Notions were adopted by Euclid from earlier textbooks. It is ? besides, scarcely credible that, if the Common Notions which Apollonius tried to prove had not been introduced earlier (e.g. by Euclid), they would then have been interpolated as axioms and not as propositions to be proved. The line taken by Apollonius is much better explained on the assumption that he was directly attacking axioms which he found already admitted into the Elements. Proclus, who recognised the five Common Notions given in the text, warns us, not only against the error of unnecessarily multiplying the axioms, but against the contrary error of reducing their number unduly (p. 196, 15), "as Heron does in enunciating three only; for it is also an axiom that the whole is greater than the part, and indeed the geometer employs this in many places for his demonstrations, and again that things which coincide are equal." Thus Heron recognised the first three of the Common Notions ; and this fact, together with Aristotle's allusions to "common axioms" (in the plural), and in particular to our Common Notion 3, may satisfy us that at least the first three Common Notions were contained in the Elements as they left Euclid's hands. Common Notion i. Things which are equal to the same thing are also equal to one another. Aristotle throughout emphasises the fact that axioms are self-evident truths, which it is impossible to demonstrate. If, he says, any one should attempt to prove them, it could only be through ignorance. Aristotle therefore would undoubtedly have agreed in Proclus' strictures on Apollonius for attempting to prove the axioms. Proclus gives (p. 194, 25), as a specimen of these attempted proofs by Apollonius, that of the first of the Common Notions. " Let A be equal to B, and the latter to C; I say that A is also equal to C. For, since A is equal to JB, it A B occupies the same space with it ; and since B is equal to C, it occupies the same space with it. Therefore A also occupies the same space with C." Proclus rightly remarks (p. 194, 23) that "the middle term is no more I. C.N 1—3] NOTES ON COMMON NOTIONS 1—3 *»3 intelligible {better known, yvmpifuar€pov} than the conclusion, if it is not actually more disputable." Again (p. 195, 6), the proof assumes two things, (1) that things which "occupy the same space" upyjj(,uvra tir oAAipXa ura aAAr/Xoi? cotcV. Things which coincide with one another are equal to one anotJur. The word itjmp/Aiiltiv, as a geometrical term, has a different meaning according as it is used in the active or in the passive. In the passive, <£af>/u>£«rltu, it means "to be applied to" without any implication that the applied figure will exactly fit, or coincide with, the figure to which it is applied ; on the other hand the active 'v^ap^t^w is used intransitively and means " to i. C. N. 4] NOTES ON COMMON NOTIONS 2—4 225 lit exactly," " to coincide with." In Euclid and Archimedes t<#>a/)^o{<«' is constructed with «ri and the accusative, in Pappus with the dative. On Common Notion 4 Tannery observes that it is incontestably geometrical in character, and should therefore have been excluded from the Common Notions; again, it is difficult to see why it is not accompanied by its converse, at all events for straight lines (and, it might be added, angles also), which Euclid makes use of in I. 4. As it is, says Tannery, we have here a definition of geometrical equality more or less sufficient, but not a real axiom. It is true that Froclus seems to recognise this Common Notion and the next as proper axioms in the passage (p. 196, 15 — 21) where he says that we should not cut down the axioms to the minimum, as Heron does in giving only three axioms; but the statement seems to rest, not upon authority, but upon an assumption that Euclid would state explicitly at the beginning all axioms subsequently used and not reducible to others unquestionably included. Now in 1. 4 this Common Notion is not quoted ; it is simply inferred that " the base BC will coincide with EF, and will be equal to it." The position is therefore the same as it is in regard to the statement in the same proposition that, "if... the base BC does not coincide with EF, two straight lines will enclose a spate : which is impossible " ; and, if we do not admit that Euclid had the axiom that " two straight lines cannot enclose a space," neither need we infer that he had Common Notion 4. I am therefore inclined to think that the latter is more likely than not to be an interpolation. It seems clear that the Common Notion, as here formulated, is intended to assert that superposition is a legitimate way of proving the equality of two figures which have the necessary parts respectively equal, or, in other words, to serve as an axiom of congruence. The phraseology of the propositions, e.g. 1. 4 and 1. 3, in which Euclid employs the method indicated, leaves no room for doubt that he regarded one figure as actually moved and placed upon the other. Thus in 1. 4 he says, " The triangle ABC being applied (t<£apfu>{o(ia>ov) to the triangle DEF, and the point A being placed {rJltjitvov) upon the point D, and the straight line AB on DE, the point B will also coincide with E because AB is equal to DE"; and in 1. 8, "If the sides BA, AC do not coincide with ED, DF, but fall beside them (take a different position, TapaAAafowrtv), then " etc. At the same time, it is clear that Euclid disliked the method and avoided it wherever he could, e.g. in 1. 26, where he proves the equality of two triangles which have two angles respectively equal to two angles and one side of the one equal to the corresponding side of the other. It looks as though he found the method handed down by tradition (we can hardly suppose that, if Thales proved that the diameter of a circle divides it into two equal parts, he would do so by any other method than that of superposition), and followed it, in -the few cases where he does so, only because he had not been able to see his way to a satisfactory substitute. But seeing how much of the Elements depends on 1. 4, directly or indirectly, the method can hardly be regarded as being, in Euclid, of only subordinate importance , on the contrary, it is fundamental. Nor, as a matter of fact, do we find in the ancient geometers any expression of doubt as to the legitimacy of the method. Archimedes uses it to prove that any spheroidal figure cut by a plane through the centre is divided into two equal parts in respect of both its surface and its volume; he also postulates in Equilibrium of Planes 1. that "when equal and similar plane figures coincide if applied to one another, their centres of gravity coincide also." Killing {Einfuhrung in die Grundlagen der Geometric, 11. pp. 4, 5) 2 26 BOOK I [i. C. M 4 contrasts the attitude of the Greek geometers with that of the philosophers, who, he says, appear to have agreed in banishing motion from geometry altogether. In support of this he refers to the view frequently expressed by Aristotle that mathematics has to do with immovable objects {wivr/rd), and that only where astronomy is admitted as part of mathematical science is motion mentioned as a subject for mathematics. Cf. Mttaph. 989 b 32 "For mathe- matical objects are among things which exist apart from motion, except such as relate to astronomy"; Metapk. rofi4 a 30 "Physics deals with things which have in themselves the principle of motion ; mathematics is a theoretical science and one concerned with things which are stationary (^mfovto) but not separable" (sc. from matter); in Pkysies if. z, 193 b 34 he speaks of the subjects of mathematics as "in thought separable from motion." But I doubt whether in Aristotle's use of the words " immovable," " with- out motion " etc. as applied to the subjects of mathematics there is any implication such as Killing supposes. We arrive at mathematical concepts by abstraction from material objects; and just as we, in thought, eliminate the matter, so according to Aristotle we eliminate the attributes of matter as such, e.g. qualitative change and motion. It does not appear to me that the use of " immovable " in the passages referred to means more than this. I do not think that Aristotle would have regarded it as illegitimate to move a geometrical figure from one position to another ; and I infer this from a passage in De caelo lit. 1 where he is criticising "those who make up every body that has an origin by putting together plants, and resolve it again into planes." The reference must be to the Timaeus (54 b sqq.) where Plato evolves the four elements in this way. He begins with a right-angled triangle in which the hypotenuse is double of the smaller side; six of these put together in the proper way produce one equilateral triangle. Making solid angles with {a) three, (6) four, and (c) five of these equilateral triangles respectively, and taking the requisite number of these solid angles, namely four of (a), six of (&) and twelve of (c) respectively, and putting them together so as to form regular solids, he obtains (a) a tetrahedron, (/J) an octahedron, (v) an icosahedron respectively. For the fourth element (earth), four isosceles right-angled triangles are first put together so as to form a square, and then six of these squares are put together to form a cube. Now, says Aristotle (299 b 23), "it is absurd that planes should only admit of being put together so as to touch in a line; for just as a line and a line are put together in both ways, lengthwise and breadthwise, so must a plane and a plane. A line can be combined with a line in the sense of being a line superposed, and not added"; the inference being that a plane can be superposed on &plane. Now this is precisely the sort of motion in question here; and Aristotle, so far from denying its permissibility, seems to blame Plato for not using it. Cf. also Physits v. 4, 228 b 25, where Aristotle speaks of "the spiral or other magnitude in which any part will not coincide with any other part," an where superposition is obviously contemplated. Motion without deformation. It is well known that Helmholtz maintained that geometry requires us to assume the actual existence of rigid bodies and their free mobility in space, whence he inferred that geometry is dependent on mechanics. Veronese exposed the fallacy in this {Fondamenti ii geometria, pp. xxxv — xxxvi, 239 — 240 note, 615 — 7), his argument being as follows. Since geometry is concerned with empty space, which is immovable, it would be at least strange if it was necessary to have recourse to the real motion of bodies for a definition, ]. C. N. 4] NOTE ON COMMON NOTION 4 xa-j and for the proof of the properties, of immovable space. We must distinguish the intuitive principle of motion in itself from that of motion without deforma- tion. Every point of a figure which moves is transferred to another point in space. " Without deformation " means that the mutual relations between the points of the figure do not change, but the relations between them and other figures do change (for if they did not, the figure could not move). Now consider what we mean by saying that, when the figure A has moved from the position A, to the position A it the relations between the points of A in the position A, are unaltered from what they were in the position A u are the same in fact as if A had not moved but remained at A t . We can only say that, judging of the figure (or the body with its physical qualities eliminated) by the impressions it produces in us during its movement, the impressions produced in us in the two different positions (which are in time distinct) are equal. In fact, we are making use of the notion of equality between two distinct figures. Thus, if we say that two bodies are equal when they can be superposed by means of movement without deformation, we are com- mitting a petitio principii. The notion of the equality of spaces is really prior to that of rigid bodies or of motion without deformation. HelmholU supported his view by reference to the process of measurement in which the measure must be, at least approximately, a rigid body, but the existence of a rigid body as a standard to measure by, and the question how we discover two equal spaces to be equal, are matters of no concern to the geometer. The method of superposition, depending on motion without deformation, is only of use as a practical test ; it has nothing to do with the theory of geometry. Compare an acute observation of Schopenhauer {Die Welt als (Vi/le, 2 ed- 1844, 11. p. 130) which was a criticism in advance of Helmholtz' theory : "I am surprised that, instead of the eleventh axiom [the Parallel-Postulate], the eighth is not rather attacked : ' Figures which coincide (sich decken) are equal to one another.' For coincidence (das Sichdecken) is either mere tautology, or something entirely empirical, which belongs, not to pure intuition (Anschauung), but to external sensuous experience. It presupposes in fact the mobility of figures ; but that which is movable in space is matter and nothing else. Thus this appeal to coincidence means leaving pure space, the sole element of geometry, in order to pass over to the material and empirical." Mr Bertrand Russell observes {Encyclopaedia Britannica, Suppl. Vol. 4, 1Q02, Art. " Geometry, non-Euclidean ") that the apparent use of motion here is deceptive ; what in geometry is called a motion is merely the transference of our attention from one figure to another. Actual superposition, which is nominally employed by Euclid, is not required; all that is required is the transference of our attention from the original figure to a new one defined by the position of some of its elements and by certain properties which it shares with the original figure. If the method of superposition is given up as a means of defining theoreti- cally the equality of two figures, some other definition of equality is necessary. But such a definition can be evolved out of empirical or practical observation of the result of superposing two material representations of figures. This is done by Veronese {Elementi di geometria, 1904) and Ingrami {Element i di geometria, 1904). Ingrami says, namely (p. 66); " If a sheet of paper be folded double, and a triangle be drawn Upon it and then cut out, we obtain two triangles superposed which we in practice call equal. If points A, B, C, D ... be marked on one of the triangles, then, when we place this triangle upon the other (so as to coincide with it), we see 1*8 BOOK I [i. C. If. 4 chat each of the particular points taken on the first is superposed on one particular point of the second in such a way that the segments AB, AC, AD, BC, BD, CD, ... ace respectively superposed on as many segments in the second triangle and are therefore equal to them respectively. In this way we justify the following "Definition of equality. " Any two figures whatever will be called equal when to the points of one the points of the other can be made to correspond univocally [i.e. every one point in one to one distinct point in the other and vice versa] in such a way that the segments which join the points, two and two, in one figure are respectively equal to the segments which join, two and two, the corresponding points in the other." Ingram! has of course previously postulated as known the signification of the phrase equal (reciiUneat) segments, of which we get a practical notion when we can place one upon the other or can place a third movable segment successively on both. New systems of Congruence-Postulates. In the fourth Article of Questioni riguardanti ie matematiehe etementari, I., pp. 93 — 122, a review is given of three different systems : (i) that of Pasch in Vorlesungen titer neuere Geometrie, 1882, p. 101 sqq., (3) that of Veronese according to the Fondamenti di geometria, 1891, and the Ekmcnti taken together, (3) that of Hilbert (see Grundlagen der Geometric, 1903, pp. 7—15). These systems differ in the particular conceptions taken by the three authors as primary, (t) Pasch considers as primary the notion of congruence or equality between any figures which are made up of a finite number 0/ points only. The definitions of congruent segments and of congruent angles have to be deduced in the way shown on pp. 102 — 103 of the Article referred to, after which Eucl. 1. 4 follows immediately, and Eucl. 1. 26 (1) and 1. 8 by a method recalling that in Eucl. 1. 7, 8. (2) Veronese takes as primary the conception of congruence between segments (rectilineal). The transition to congruent angles, and thence to triangles is made by means of the following postulate: "Let AB, ^Cand AB, A'C be two pairs of straight lines intersecting at A, A', and let there be determined upon them the congruent segments AB, A'ff and the congruent segments AC, A'C ; then, if BC, BC are congTuent, the two pairs of straight lines are con- gruent." fe icAoA w greater than the part. Prod us includes this " axiom " on the same ground as the preceding one. I think however there is force in the objection which Tannery takes to it, namely that it replaces a different expression in Eucl. f. 6, where it is stated that " the triangle DBC will be equal to the triangle A CB, the /ess to the greater: which is absurd." The axiom appears to be an abstraction or generalisation substituted for an immediate inference from a geometrical figure, but it takes the form of a sort of definition of whole and part. The probabilities seem to be against its being genuine, notwithstanding Proclus' approval of it Clavius added the axiom that the whole is the equal to the sum of its parts. Other Axioms introduced after Euclid's time. [9] Two straight lines do not enclose (or can tain) a space. Proclus {p. 196, 21) mentions this in illustration of the undue multiplication of axioms, and he points out, as an objection to it, that it belongs to the subject matter of geometry, whereas axioms are of a general character, and not peculiar to any one science. The real objection to the axiom is that it is unnecessary, since i.he fact which it states is included in the meaning of Postulate i. It was nr> doubt taken from the passage in 1. 4, "if. the base BC does not coincide with the base EF, two straight tines wilt enclose a space : which is impossible"; and we must certainly regard it as an interpolation, notwithstanding that two of the best mss. have it after Postulate 5, and one gives it as Common Notion 9. Pappus added some others which Proclus objects to (p. 198, 5) because they are either anticipated in the definitions or follow from them. (g) All the parts of a plane, or of a straight line, coincide with one another. \K) A point divides a line, a line a surface, and a surface a solid; on which Proclus remarks that everything is divided by the same things as those by which it is bounded. An-Nairizi {ed. Besthorn-Heiberg, p. 31, ed. Curtze, p. 38) in his version of this axiom, which be also attributes to Pappus, omits the reference to solids, but mentions planes as a particular case of surfaces. " (a) A Surface cuts a surface in a line ; fj8) If two surfaces which cut one another are plane, they cut one another in a straight line ; (y) A line cuts a line in a point (this last we need in the first proposition}." (A) Magnitudes are susceptible of the infinite {or unlimited) both by way of addition and by way of successive diminution, but in both cases potentially only (to mrttpov fv rots fnyi$ttrw iirriv got Tp xpoaGitrtt Kal rp iiittai0atpio , u, oW ifLtt Si imLrtpav). An-Nairizi's version of this refers to straight lines and plane surfaces only : "as regards the straight line and the plane surface, in consequence of their evenness, it is possible to produce them indefinitely. This "axiom" of Pappus, as quoted by Proclus, seems to be taken directly from the discussion of to avtipov in Aristotle, Physics ill. 5 — 8, even to the wording, for, while Aristotle uses the term division {huaiptatsi) most frequently as the antithesis of addition (avrBttrn), he occasionally speaks of subtraction (AAJHupttrn) and diminution (icaflaipHrii). Hankel (Zur Geschichte der Mathe- matik im Alterthum und Mittelalter, 1874, pp. 119 — 120) gave an admirable 1. Axx.] ADDITIONAL AXIOMS 233 summary of Aristotle's views on this subject ; and they are stated in greater detail in Gorland, AristoteUs und die Mathematik, Marburg, 1899, pp. 157— 183. The infinite or unlimited (airttpov) only exists potentially {Sura/i«), not in actuality (Ivtpytiq.). The infinite is so in virtue of its endlessly changing into something else, like day or the Olympic Games {Phys. m. 6, 206 a 15 — 2 S)- The infinite is manifested in different forms in time, in Man, and in the division of magnitudes. For, in general, the infinite consists in something new being continually taken, that something being itself always finite but always different. Therefore the infinite must not be regarded as a particular thing (toS« t»), as man, house, but as being always in course of becoming or decay, and, though finite at any moment, always different from moment to moment. But there is the distinction between the forms above referred to that, whereas in the case of magnitudes what is once taken remains, in the case of time and Man it passes or is destroyed but the succession is unbroken. The case of addition is in a sense the same as that of division ; in the finite magnitude the former takes place in the converse way to the latter ; for, as we see the finite magnitude divided ad infinitum, so we shall find that addition gives a sum tending to a definite limit. I mean that, in the case of a finite magnitude, you may take a definite fraction of it and add to it (continually) in the same ratio ; if now the successive added terms do not include one and the same magnitude whatever it is [i.e. if the successive terms diminish in geometrical progression], you will not come to the end of the finite magnitude, but, if the ratio is increased so that each term does include one and the same magnitude whatever it is, you will come to the end of the finite magnitude, for every finite magnitude is exhausted by continually taking from it any definite fraction whatever. Thus in no other sense does the infinite exist, but only in the sense just mentioned, that is, potentially and by way of diminution (106 a 25 — b 13). And in this sense you may have potentially infinite addition, the process being, as we say, in a manner, the same as with division ad infinitum : for in the case of addition you will always be able to find some* thing outside the total for the time being, but the total will never exceed every definite (or assigned) magnitude in the way that, in the direction of division, the result will pass every definite magnitude, that is, by becoming smaller than it. The infinite therefore cannot exist even potentially in the sense of exceeding every finite magnitude as the result of successive addition (206 b 16 — 22). It follows that the correct view of the infinite is the opposite of that commonly held : it is not that which has nothing outside it, but that which always has something outside it (206 b 33 — 207 a r). Contrasting the case of number and magnitude, Aristotle points out that (1) in number there is a limit in the direction of smallness, namely unity, but none in the other direction : a number may exceed any assigned number however great ; but (2) with magnitude the contrary is the case : you can find a magnitude smaller than any assigned magnitude, but in the other direction there is no such thing as an infinite magnitude (207 b r — 5). The tatter assertion he justified by the following argument. However large a thing can be potentially, it can be as large actually. But there is no magnitude perceptible to sense that is infinite. Therefore excess over every assigned magnitude is an impossibility j otherwise there would be something larger than the universe (oipavot) {207 b 17—21). Aristotle is aware that it is essentially of physical magnitudes that he is speaking. He had observed in an earlier passage (PAys, in, 5, 204 a 34) that it ts perhaps a more general inquiry that would be necessary to determine *34 BOOK ! [i. Axx. whether the infinite is possible in mathematics, and in the domain of thought and of things which have no magnitude ; but he excuses himself from entering upon this inquiry on the ground that his subject is physics and sensible objects. He returns however to the bearing of his conclusions on mathematics in m. 7, 207 b 2j : "my argument does not even rob mathematicians of their study, although it denies the existence of the infinite in the sense of actual existence as something increased to such an extent that it cannot be gone through (dSwfiTifrov) ; for, as it is, they do not even need the infinite or use it, but only require that the finite (straight line) shall be as long as they please; and another magnitude of any size whatever can be cut in the same ratio as the greatest magnitude. Hence it will make no difference to them for the purpose of demonstration." I^astly, if it should be urged that the infinite exists in thought, Aristotle replies that this does not involve its existence in fact. A thing is not greater than a certain size because it is conceived to be so, but because it is; and magnitude is not infinite in virtue of increase in thought (ao8 a 16 — zz). Hankel and Gorland do not quote the passage about an infinite series of magnitudes {206 b 3—13) included in the above paraphrase; but I have thought that mathematicians would be interested in the distinct expression of Aristotle's view that the existence of an infinite series the terms of which are magnitudes is impossible unless it is convergent, and (with reference to Riemann's developments) in the statement that it does not matter to geometry if the straight line is not infinite in length, provided that it is as long as we please. Aristotle's denial of even the potential existence of a sum of magnitudes which shall exceed every definite magnitude was, as he himself implies, in conflict with the lemma or assumption used by Eudoxus (as we infer from Archimedes) to prove the theorem about the volume of a pyramid. The lemma is thus stated by Archimedes (Quadrature of a parabola, preface): " The excess by which the greater of two unequal areas exceeds the less can, if it be continually added to itself, be made to exceed any assigned finite area." We can therefore well understand why, a century later, Archimedes felt it necessary to justify his own use of the lemma as he does in the same preface ; " The earlier geometers too have used this lemma : for it is by its help that they have proved that circles have to one another the duplicate ratio of their diameters, that spheres have to one another the triplicate ratio of their diameters, and so on. And, in the result, each of the said theorems has been accepted no less than those proved without the aid of this lemma." Principle of continuity. The use of actual construction as a method of proving the existence ot figures having certain properties is one of the characteristics of the Elements. Now constructions are effected by means of straight lines and circles drawn in accordance with Postulates 1 — 3 ; the essence of them is that such straight lines and circles determine by their intersections other points in addition to those given, and these points again are used to determine new lines, and so on. This being so, the existence of such points of intersection must be postulated or proved in the same way as that of the lines which determine them. Yet there is no postulate of thfs character expressed in Euclid except Post J. This postulate asserts that two straight lines meet if they satisfy a certain condition. The condition is of the nature of a Stop«r>«5s (discrimination, or condition of possibility) in a problem ; and, if the existence of the point of i. Axx.l PRINCIPLE OF CONTINUITY 335 intersection were not granted, the solutions of' problems in which the points of intersection of straight lines are used would not in general furnish the required proofs of the existence of the figures to be constructed. But, equally with the intersections of straight lines, the intersections of circle with straight line, and of circle with circle, are used in constructions. Hence, in addition to Postulate 5, we require postulates asserting the actual existence of points of intersection of circle with straight line and of circle with circle. In the very first proposition the vertex of the required equilateral triangle is determined as one of the intersections of two circles, and we need therefore to be assured that the circles will intersect. Euclid seems to assume it as obvious, although it is not so; and he makes a similar assumption in 1. 2 2. It is true that in the latter case Euclid adds to the enunciation that two of the given straight lines must be together greater than the third ; but there is nothing to show that, if this condition is satisfied, the construction is always possible. In 1. 12, in order to be sure that the circle with a given centre will intersect a given straight line, Euclid makes the circle pass through a point on the side of the line opposite to that where the centre is. It appears therefore as if, in this case, he based his inference in some way upon the definition of a circle combined with the fact that the point within it called the centre is on one side of the straight line and one point of the circumference on the other, and, in the case of two intersecting circles, upon similar con- siderations. But not even in Book hi., where there are several propositions about the relative positions of two circles, do we find any discussion of the conditions under which two circles have two, one, or no point common. The deficiency can only be made good by the Principle of Continuity. Killing {Einfiihrung in die Grundlagen der Geometric, 11. p 43) gives the following forms as sufficient for most purposes (a) Suppose a line belongs entirely to a figure which is divided into two parts; then, if the line has at least one point common with each part, it must also meet the boundary between the parts; or {b) If a point moves in a figure which is divided into two parts, and if it belongs at the beginning of the motion to one part and at the end of the motion to the other part, it must during the motion arrive at the boundary between the two parts. I n the Questioni riguardanti le matematuhe elemeniari, l.,Art.s,pp. 123—143, the principle of continuity is discussed with special reference to the Postulate of Dedektnd, and it is shown, first, how the Postulate may be led up to and, secondly, how it may be applied for the purposes of elementary geometry. Suppose that in a segment A B of a straight line a point C determines two segments AC, CB. If we consider the point Cas belonging to only one of the two segments A C, CB, we have a division of the segment AB into two parts with the following properties. 1. Every point of the segment AB belongs to one of the two parts. 2. The point A belongs to one of the two parts (which we will call the first) and the point B to the other; the point C may belong indifferently to one or the other of the two parts according as we choose to premise. 3. Every point of the first part precedes every point of the second in the order AB of the segment. (For generality we may also suppose the case in which the point C falls at A or at B. Considering C, in these cases respectively, as belonging to the first or second part, we still have a division into parts which have the properties above enunciated, one part being then a single point A or B.) J36 BOOK 1 [i. Axx. Now, considering carefully the inverse of the above proposition, we see that it agrees with the idea which we have of the continuity of the straight line. Consequently we are induced to admit as a postulate the following. If a segment of a straight lint AB is divided into two parts so that (i ) every point of the segment AB belongs to one of the parts, (2) the extremity A belongs to the first part and B to the second, and (3) any point whatever of the first part precedes any point whatever of the second part, in the order AB of the segment, there exists a point C of the segment AB {which may belong either to one part or to the other) such that every point of AB that precedes C belongs to the first part, and every point of AB that follows C belongs to the second part in the division originally assumed. (If one of the two parts consists of the single point A or B, the point C is the said extremity A or B of the segment.) This is the Postulate of Dedekind, which was enunciated by Dedekind himself in the following slightly different form (Stetigkcit unci irrationale Zahlen, 187s, new edition 1905, p. 11). " If all points of a straight line fall into hvo classes such that every point of the first class lies to the left of every point of the second class, there exists one and only one point which produces this division of all the points into two daises, this division of the straight line into two parts." The above enunciation may be said to correspond to the intuitive notion which we have that, if in a segment of a straight line two points start from the ends and describe the segment in opposite senses, they meet in a point. The point of meeting might be regarded as belonging to both parts, but for the present purpose we must regard it as belonging to one only and subtracted from the other part. Application of Dedckind's postulate to angles. If we consider an angle less than two right angles bounded by two rays a, b, and draw the straight line connecting A, a point on a, with B, a point on b, we see that all points on the finite segment AB correspond univocally to all the rays of the angle, the point corresponding to any ray being the point in which the ray cuts the segment AB ; and if a ray be supposed to move about the vertex of the angle from the position a to the position b, the corresponding points of the segment AB are seen to follow in the same order as the corresponding rays of the angle (ab). Consequently, if the angle (ab) is divided into two parts so that (1) each ray of the angle (ab) belongs to one of the two parts, (t) the outside ray a belongs to the first part and the ray b to the second, (3) any ray whatever of the first part precedes any ray whatever of the second part, the corresponding points of the segment AB determine tteo parts of the segments such that (1) every point of the segment AB belongs to one of the two parts, (2) the extremity A belongs to the first part and B to the second, (3) any point whatever of the first part precedes any point whatever of the second. But in that case there exists a point C of AB {which may belong to one or the other of the two parts) such that every point of AB that precedes C belongs to the first part and every point of AB that follows C belongs to the second part. I. Axx.] APPLICATIONS OF DEDEKIND'S POSTULATE ^37 Thus exactly the same thing holds of c, the ray corresponding to C, with reference to the division of the angle {ad) into two parts. It is not difficult to extend this to an angle (ai) which is either flat or greater than two right angles; this is done (Vitali, op. cit. pp. 126—127) by supposing the angle to be divided into two, {ad), {di>), each less than two right angles, and considering the three cases in which (1) the ray d is such that all the rays that precede it belong to the first patt and those which follow it to the second part, (2) the ray d is followed by some rays of the first part, (3) the ray d is preceded by some rays of the second part. Application to circular arcs. If we consider an arc AB of a circle with centre O, the points of the arc correspond uni vocally, and in the same order, to the rays from the point O passing through those points respectively, and the same argument by which we passed from a segment of a straight line to an angle can be used to make the transition from an angle to an arc. Intersections of a straight line with a circle. It is possible to use the Postulate of Dedekind to prove that If a straight line has one point inside and one point outside a circle, it has two points common with the circle. For this purpose it is necessary to assume (1) the proposition with reference to the perpendicular and obliques drawn from a given point to a given straight line, namely that of all straight lines drawn from a given point to a given straight line the perpendicular is the shortest, and of the rest (the obliques) that is the longer which has the longer projection upon the straight line, while those are equal the projections of which are equal, so that for any given length of projection there are two equal obliques and two only, one on each side of the perpendicular, and (2) the proposition that any side of a triangle is less than the sum of the other two. Consider the circle {(7) with centre 0, and a straight line {r) with one point A inside and one point B outside the circle. By the definition of the circle, if R is the radius, OAR. Draw OP perpendicular to the straight line r. Then OP< OA, so that OP is always less than R, and P is therefore within the circle C. Now let us fix our attention on the finite segment AB of the straight line r. It can be divided into two parts, (1) that containing all the points H for which OIf< R (i.e. points inside C), and (2) that containing all the points K for which OK £ R (points outside C or on the circumference of C). Thus, remembering that, of two obliques from a given point to a given straight line, that is greater the projection of which is greater, we can assert that all the points of the segment PB which precede a point inside C are inside C, and those which follow a point on the circumference of C or outside C are outside C. Hence, by the Postulate of Dedekind, there exists on the segment PB a 238 BOOK I [1. Axx. point M such that all the points which precede it belong to the first part and those which follow it to the second part. I say that M is common to the straight line r and the circle C, or OM=R. For suppose, e.g., that OM <.R. There will then exist a segment (or length) a less than the difference between R and OM. Consider the point Af, one of those which follow M, such that MAT is equal to a. Then, because any side of a triangle is less than the sum of the Other two, OM' < OM+ MM'. But OM+ MM' = OM+ R. Therefore OM mist be equal to R. It is immediately obvious that, corresponding to the point Mor\ the segment PS which is common to r and C, there is another point on r which has the same property, namely that which is symmetrical to M with respect to P. And the proposition is proved. Intersections of two circles. We can likewise use the Postulate of Dedekind to prove that If in a given plane a circle C has one paint X inside and one point Y outside another circle C ', the two circles intersect in two points. We must first prove the following Lemma. If O, 0' are the centres of two circles C, C, and R, R' their radii respectively, the straight line 00' meets the circle C in two points A, S, one of which is inside C" and the other outside it. Now one of these points must fall (r) on the prolongation of 00 beyond O or {2) on 00 itself or {3) on the prolongation of OO' beyond 0. (1) First, suppose A to lie on OO pro- duced. Then A0=AO+ 00' = JP+ OO" (a). But, in the triangle OO Y, 0YR\ OY=R, RR; and therefore lies outside C". (2) Secondly, suppose A to lie on 00. Then 00 = OA + A0 = R + A0 ...($). From the triangle 00 X we have O0t(a irXtupA ij Br * W to A). 13. the remainder AL...the remainder BG. Tbe Greek expressions ate X«r^ ^ A A and XorviJ r£ BH> and the literal translation would be 'Mi (or BG) rtmmning™ bat tbe shade of meaning conveyed by the position of the definite article can hardly be expressed in English. This proposition gives Proclus an opportunity, such as the Greek commentators revelled in, of distinguishing a multitude of cases. After explaining that those theorems and problems are said to have eases which have the same force, though admitting of a number of different figures, and preserve the same method of demonstration while admitting variations of position, and that cases reveal themselves in the etms/rue/tbn, he proceeds to distinguish the cases in this problem arising from the different positions which the given point may occupy relatively to the given straight line. It may be (he says) either (i) outside the line or (a) on the line, and, if (1), it may be either (a) on the line produced or (6) situated obliquely with regard to it ; if (a), it may be either (a) one of the extremities of the line or (A) an intermediate point on it. It will be seen that Proclus' anxiety to subdivide leads him to give a "case," (a) (a), which is useless, since in that "case" we are given what we are required to rind, and there is really no problem to solve. As Savile says, " qui quaerit ad punctum ponere rectam aequaiem rjj fiy rectae, quaerit quod datum est, quod nemo faceret nisi forte msaniat," Proclus gives the construction for (a) (i) following Euclid's way of taking G as the point in .which the circle with centre B intersects DB produced, and then proceeds to " cases," of which there are still more, which result from the different ways of drawing the equilateral triangle and of producing its sides. This last class of "cases" he subdivides into three according as AB is (1) equal to, {a) greater than or (3) less than BC, Here again "case "(i) serves no purpose, since, if AB is equal to BC, the problem is already solved. But Proclus' figures for the other two cases are worth giving, because in one of them the point G is on BD produced beyond D, and in the other it lies on BD itself and there is no need to produce any side of the equilateral triangle. A glance at these figures will show that, if they were used in the proposition, each of them would require a slight modification in the wording (r) of the construction, since BD is in one case produced beyond D instead of B and in the other case not produced at all, (a) of the proof, since BG, instead of being the difference between DG and DB, is in one case the sum of DG and DB and in the other the difference between DB and DG. 14<> BOOK I [i. a, 3 Modern editors generally seem to classify the cases according to the possible variations in the construction rather than according to differences in the data. Thus Lardner, Potts, and Todhunter distinguish eight cases due to the three possible alternatives, (1) that the given point may be joined to either end of the given straight line, {2) that the equilateral triangle may then be described on either side of the joining line, and {3) that the side of the equilateral triangle which is produced may be produced in either direction. (But it should have been observed that, where AB is greater than BC, the third alternative is between producing DB and not producing it at all.) Potts adds that, when the given point lies either on the line or on the line produced, the distinction which arises from joining the two ends of the line with the given point no longer exists, and there are only four cases of the problem (I think he should rather have said solutions). To distinguish a number of cases in this way was foreign to the really classical manner. Thus, as we shall see, Euclid's method is to give one case only, for choice the most difficult, leaving the reader to supply the rest for himself. Where there was a real distinction between cases, sufficient to necessitate a substantial difference in the proof, the practice was to give separate enunciations and proofs altogether as we may see, e.g., from the Cenia and the De section* rationis of Apollonius. Proclus alludes, in conclusion, to the error of those who proposed to solve 1. 2 by describing a circle with the given point as centre and with a distance equal to BC, which, as he says, is a petitio principii. De Morgan puts the matter very clearly {Supplementary Remarks on the first six Books* of Euclid's Elements in the Companion to the Almanac, 1849, p. 6). We should "insist," he says, "here upon the restrictions imposed by the first three postulates, which do not allow a circle to be drawn with a compass-carried distance; suppose the compasses to dose of themselves the moment they cease to touch the paper. These two propositions [1. 2, 3] extend the power of construction to what it would have been if all the usual power of the compasses had been assumed ; they are mysterious to all who do not see that postulate iii does not ask for every use of the compasses." Proposition 3. Given two unequal straight lines, to cut off from the greater a straight line equal to the less. c Let AB, C be the two given un- equal straight lines, and let AB be the greater of them. Thus it is required to cut off from AB the greater a straight line equal to C the less. At the point A let AD be placed equal to the straight line C ; [1. 2] and with centre A and distance AD let the circle DEF be described. [Post 3] I. 3, 4] PROPOSITIONS 2—4 *4T Now, since the point A is the centre of the circle DEF, AE is equal to AD. [Def. 15] But C is also equal to AD. Therefore each of the straight lines AE, C is equal to AD ; so that A E is also equal to C. [C.N. 1] Therefore, given the two straight lines AB, C, from AB the greater AE has been cut off equal to C the less. (Being) what it was required to do. P roc his contrives to make a number of "cases" out of this proposition also, and gives as many as eight figures. But he only produces this variety by practically incorporating the construction of the preceding proposition, instead of assuming it as we are entitled to do. If Prop, 2 is assumed, there is really only one " case " of the present proposition, for Potts distinction between two cases according to the particular extremity of the straight line from which the given length has to be cut off scarcely seems to be worth making. Proposition 4. If two triangles have the two sides equal to two sides respectively, and have the angles contained by the equal straight lines equal, they will also have the base equal to the base, the triangle will be equal to the triangle, and the remaining angles s will be equal to the remaining angles respectively, namely those which the equal sides subtend. Let ABC, DEF be two triangles having the two sides AB, AC equal to the two sides DE, DF respectively, namely AB to DE and AC to DF, and the angle BA C equal to the 10 angle EDF. I say that the base BC is also equal to the base EF, the triangle ABC will be equal to the triangle DEF, and the remaining angles will be equal to the remaining angles respectively, namely those which the equal sides subtend, that is is, the angle ABC to the angle DEF, and the angle ACB to the angle DEE. For, if the triangle ABC be applied to the triangle DEF, and if the point A be placed ao On the point D and the straight line AB on DE, then the point B will also coincide with E, because AB is equal to DE. 348 BOOK I fjfc 4 »j Again, AB coinciding with DE, the straight line AC will also coincide with DF, because the angle SAC is equal to the angle EDF; hence the point C will also coincide with the point F, because AC is again equal to DF. 30 But B also coincided with E ; hence the base BC will coincide with the base EF. [For if, when B coincides with E and C with F, the base BC does not coincide with the base EF, two straight lines will enclose a space : which is impossible. 35 Therefore the base BC will coincide with EF~\ and will be equal to it. \C.N. 4] Thus the whole triangle ABC will coincide with the whole triangle DEF, and will be equal to it. 40 And the remaining angles will also coincide with the remaining angles and will be equal to them, the angle ABC to the angle DEF, and the angle ACB to the angle DFE, Therefore etc. *S (Being) what it was required to prove. 1 — 3. It is a fact that Euclid's enunciations not infrequently leave something to be desired in point of clearness and precision. Here he speaks of the triangles having *' the angle equal to the angle, namely the angle contained by the equal straight lines " [rrtr yuriar r£ ywif tar}* §xv T h r v*& ruv law iijdtiutv wtpi.txt>M yr l v )i only one of the two angles being described in the latter expression (in the accusative], and a similar expression in the dative being left to be understood of the other angle. It is curious too that, after mentioning two "sides" he speaks of the angles contained by the equal "straight lints" not "sides. It may be that he wished to adhere scrupulously, at the outset, to the phraseology of the definitions, where the angle is the inclination to one another of two lines or straight lints. Similarly in the enunciation of I: £ he speaks of producing the equal " straight lines" as if to keep strictly to the wording of Postulate 1. t. respectively. I agree with Mr H. M. Taylor {Euclid, p. ix) that it is best to abandon the traditional translation of [ ' each to each," which would naturally seem to imply that all the four magnitudes are equal rather than (as the Greek itaripa. 1 tariff does) that one is equal to one and the other to the other. 3. the base. Here we have the word bast used for the first time in the Elements. Proclus explains it (p. 136, 13 — tj) as meaning (1), when no side of a triangle has been mentioned before, the side " which is on a level with the sight " (rV rpbi tq ftipet ntifiirjjr), and (1), when two sides have already been mentioned, the third side. Proclus thus avoids the mistake made by some modern editors who explain the term exclusively with reference to the case where two sides have been mentioned before. That this is an error is proved (t) by the occurrence of .the term in the enunciations of 1. 37 etc. about triangles on the same base and equal bases, (1) by the application of the same term to the bases of parallelograms in h 35 etc. The truth is that the use of the term must have been suggested by the practice of drawing the particular side horizontally, as it were, and the rest of the figure above it. The bait of a figure was therefore spoken of, primarily, in the same sense as the base of anything r. 4] PROPOSITION 4 *49 else. e.g. of a pedestal or column ; but 'vhen, as in ]. 5. two triangles were compared occupying other thun the norma! positions which gave rise to the name, and when two side' had been previously mentioned , the base was as Prod us says, necessarily the third side. 6. subtend. Owvrttteuf br6, " to stretch under," with accusative* 9. the angle BAC. The full Greek expression would be it irro rwp BA, AT vtpuxo^rn yaivtu, " the angle contained by the (straight lines) BA, AC." But it was a common practice of Greek geometers, e.g. of Archimedes and Apollonius (and Euclid too in Books x.— xiij.), to use the abbreviation at BAl' for at BA, AT, "the (straight lines} BA, AC." Thus, on TtpttX'tfni being dropped, the expression would become first ^ iiri rat BAr -yafta, then i M BAP yttrln, and finally i brb BAr, without ywla, as we regularly find it in Euclid. 17. if the triangle be applied to..., 13. coincide. The difference between the technical use of the passive t>apjt4f»T0ai " to be applitd (to)," and of the active t£np^f«> "to coincide (with} has been noticed above (note on Common tfetien 4, pp. 114 — j). J j. [For if, when B coincides... j'>. coincide with EF]. Heiberg (ParaHpumcna i» lid in Hermts, xxxvin., 1003, p. 56] has pointed out, as a conclusive reason for regarding these words as an early interpolation, that the text of an-NairM [Codex Ltidtnsis 300, 1, ed. Besthom- Heiberg, p. 55) does not give the words in this place but after the conclusion q.e.d., which shows that they constitute a scholium only. They were doubt less added b? some commentator who thought it necessary to explain the immediate inference that, since B coincides with E and C with F. the straight line BC coincides with the straight line EF, an inference which really follows from the definition of a straight tine and Post. 1 ; and no doubt the Postulate that "Two straight lines cannot enclose a space" (afterwards placed among the Common Notions) was interpolated at the same time. 44. Therefore etc. Where (as here) Euclid's conclusion merely repeats the enunciation word for word, I shall avoid the repetition and write " Therefore etc" simply. In the note on Common Notion 4 I have already mentioned that Euclid obviously used the method of superposition with reluctance, and I have given, after Veronese for the most part, the reason for holding that that method is not admissible as a theoretical means of proving equality, although it may be of use as a practical test, and may thus furnish an empirical basis on which to found a postulate. Mr Bertrand Russell observes {Principles of Mathematics I. p. 405) that Euclid would have done better to assume 1. 4 as an axiom, as is practically done by Hilbert (Grundtagen der Geometric, p. 9). It may be that Euclid himself was as well aware of the objections to the method as are his modem critics ; but at all events those objections were stated, with almost equal clearness, as early as the middle of the 16th century. Peletarius (Jacques Peletier) has a long note on this proposition (Jn Eudidis Elementa gtometrica demonttratwnunt libri sex, 1557), in which he observes that, if superposition of lines and figures could be assumed as a method of proof, the whole of geometry would be full of such proofs, that it could equally well have been used in 1. 2, 3 (thus in 1. * we could simply have supposed the line taken up and placed at the point), and that in short it is obvious how far removed the method is from the dignity of geometry. The theorem, he adds, is obvious in itself and does not require proof ; although it is introduced as a theorem, it would seem that Euclid intended it rather as a definition than a theorem, " for I cannot think that two angles are equal unless I have a conception of what equality of angles is." Why then did Euclid include the proposition among theorems, instead of placing it among the axioms ? Peletarius makes the best excuse he can, but concludes thus : " Huius itaque propositionis veritatem non aliunde quam a communi iudirio petemus : cogitabimusque figuras figuris superponere, Mechantcum quippiam esse : intelligere verb, id demum esse Mathematicum." Expressed in terms of the modern systems of Congruence-Axioms referred to in the note on Common Notion 4, what Euclid really assumes amounts to the following : (1) On the line DE, there is a point E, on either side of D, such that AB is equal to DE. *5<> BOOK I [1.4 (2) On either side of the ray DE there is a ray DF such that the angle EDF is equal to the angle BAC- It now follows that on DF there is a point ^ such that DF is equal to^C. And lastly (3), we require an axiom from which to infer that the two remaining angles of the triangles are respectively equal and that the bases are equal. I have shown above (pp. 229 — 230) that Hilbert has an axiom stating the equality of the remaining angles simply, but proves the equality of the bases. Another alternative is that of Pasch ( VorUsungen titer neuert Geometric, p. 109) who has the following "Gnmdsatz": If two figures AB and FGH are given (FGH not being contained in a straight length), and AB, FG are congruent, and if a plane surface be laid through A and B, we can specify in this plane surface, produced if necessary, two points C, D, neither more nor less, such that the figures ABC and ABD are congruent with the figure FGH t and the straight line CD has with the straight line AB or with AB produced one point common. I pass to two points of detail in Euclid's proof : (1) The inference that, since B coincides with E, and C with F, the bases of the triangles are wholly coincident rests, as expressly stated, on the impossibility of two straight tines enclosing a space, and therefore presents no difficulty. But (2) most editors seem to have failed to observe that at the very beginning of the proof a much more serious assumption is made without any explanation whatever, namely that, if A be placed on D, and AB on DE, the point B will coincide with £, because AB is equal to DE. That is, the converse of Common Notion 4 is assumed for straight lines. Proem s merely observes, with regard to the converse of this Common Notion, that it is only true in the case of things " of the same form " (i/toeiBi}), which he explains as meaning straight lines, arcs of one and the same circle, and angles " contained by lines similar and similarly situated" (p. 241, 3 — 8^. Savile however saw the difficulty and grappled with it in his note on the Common Notion. After stating that all straight lines with two points common are congruent between them (for otherwise two straight lines would enclose a space), he argues thus. Let there be two straight lines AB, DE, and let A be placed on D, and AB on DE. Then B will coincide with E. For, if not, let B fall somewhere short of E or beyond E ; and in either case it will follow that the less is equal to the greater, which is impossible. Savile seems to assume (and so apparently does Lardner who gives the same proof) that, if the straight lines be " applied," B will fall somewhere on DE or DE produced. But the ground for this assumption should surely be stated ; and it seems to me that it is necessary to use, not Postulate 1 alone, nor Postulate 2 alone, but both, for this purpose (in other words to assume, not only that two straight lines cannot enclose a space, but also that two straight lines cannot have a common segment). For the only safe course is to place A upon D and then turn AB about D until some point on AB intermediate between A and B coincides with some point on DE. In this position AB and DE have two points common. Then Postulate 1 enables us to infer that the straight lines coincide between the two common points, and Postulate 2 that they coincide beyond the second common point towards B and E. Thus the straight lines coincide throughout so far as both extend; and Savile's argument then proves that B coincides with E. I. 5] PROPOSITIONS 4, 5 151 Proposition 5. In isosceles triangles the angles at the base are equal to one another, and, if the equal straight lines be produced further, the angles under the base will be equal to one another. Let ABC be an isosceles triangle having the side AB 5 equal to the side AC; ' and let the straight lines BD, CE be produced further in a straight line with AB, AC. [Post. 2} 1 say that the angle ABC is equal to the angle ACS, and the angle CBD to the angle BCE. 10 Let a point F be taken at random on BD; from AE the greater let AG be cut off equal to AF the less ; [1. 3] and let the straight lines FC, GB be joined. [Post 1] is Then, since AF is equal to AG and AB to AC, the two sides FA, AC are equal to the two sides GA, AB, respectively ; and they contain a common angle, the angle FAG. 10 Therefore the base FC is equal to the base GB, and the triangle AFC is equal to the triangle AGB, and the remaining angles will be equal to the remaining angles respectively, namely those which the equal sides subtend, that is, the angle ACF to the angle ABG, 11 and the angle AFC to the angle AGB. [l 4] And, since the whole AF is equal to the whole AG, and in these AB is equal to AC, the remainder BF is equal to the remainder CG. But FC was also proved equal to GB ; v> therefore the two sides BF, FC are equal to the two sides CG, GB respectively ; and the angle BFC is equal to the angle CGB, while the base BC is common to them ; therefore the triangle BFC is also equal to the triangle CGB, (S and the remaining angles will be equal to the remaining 3S* BOOK I [l. 5 angles respectively, namely those which the equal sides subtend ; therefore the angle FBC is equal to the angle GCB, and the angle BCF to the angle CBG. 4° Accordingly, since the whole angle ABG was proved equal to the angle ACF, and in these the angle CBG is equal to the angle BCF, the remaining angle ABC is equal to the remaining angle ACB; 45 and they are at the base of the triangle ABC. But the angle FBC was also proved equal to the angle GCB ; and they are under the base. Therefore etc. Q. e. d. 1. the equal straight lines (meaning the equal sida). Cf. note on the similar expression in Prop. 4, lines 1, 3. 10, Let a point F be taken at random on BD, ettrj^tfu irl rQt BA Tv%6r miitur t& Z, where rvxhr a-rj^fiof means "a chance point/' 17, the two sides FA, AC are equal to the two aides OA, AB respectively, l(So at ZA, AT 0url toTi HA, AB loat flair iuaripa tzartpf. Here, and in numberless later passages, I have inserted the word " sides" for the reason given in the note on 1- r, line 10. It would have been permissible to supply either "straight lines'' or "sides"; but on the whole M sides " seems to be more in accordance with the phraseology of I. 4, 33. the base BC is common to them, i.e., apparently, common to the origin, as the arW-ur in pdtrti a&rww ttotrij can only refer to yuyLa and yuAa preceding. Simson wrote "and the base BC is common to the two triangles BFC, CGB , Todhunter left out these words as being of no use and tending to perplex a beginner. But Euclid evidently chose to quote the conclusion of 1. 4 exactly I the first phrase of that conclusion is that the bases (of the two triangles) are equal, and, as the equal bases are here the same base, Euclid naturally substitutes the worn "common" for '* equal." 48. As " (Being) what it was required to prove " {or " do ") is somewhat long, 1 shall henceforth write the time-honoured "Q. e. D. and "<}. F_ F." for irtp titt ititfit and ortp Wet rotfyrat. According to Proclus (p. 250, 20) the discoverer of the fact that in any isosceles triangle the angles at the base are equal was Thales, who however is said to have spoken of the angles as being similar, and not as being equal. (Cf. ArisL De caelo iv. 4, 311 b 34 n-pos ifiolas ymvlat dxuVcrtu dxpoVcvov where equal angles are meant.) A pre-Euclidean proof of I. 5. One of the most interesting of the passages in Aristotle indicating differences between Euclid's proofs and those with which Aristotle was familiar, in other words, those of the text-books immediately preceding Euclid's, has reference to the theorem of 1. 5. The passage {Anal. Prior. 1. 34, 41 b 13 — 21) is so important that I must quote it in Kill. Aristotle is illustrating the fact that in any syllogism one of the propositions must be affirmative and universal {uttftJAou}. "This," he says, "is better shown in the case of geometrical propositions " {b> toii StaypapiMurtv), e.g. the proposition that the angles at Ike bast of an isosceles triangle are equal. " For let A, B be drawn [i.e. joined] to the centre. I. s] PROPOSITION s *S3 "If, then, we assumed (i) that the angle AC [i.e. A + C] is equal to the angle BD [i.e. B + D\ without asserting generally that the angles of semicircles are equal, and again (a) that the angle C is equal to the angle D without making the further assumption that the two angles of all segments art equal, and if we then inferred, lastly, that, since the whole angles are equal, and equal angles are subtracted from them, the angles which remain, namely E, F, are equal, we should commit a petitio principii, unless we assumed [generally j that, when equals art subtracted from equals, tht remainders are equal.'' The language is noteworthy in some respects. (i ) A, B are said to be drawn (iyy/«Viii) to the centre (of the circle of which the two equal sides are radii) as if A, B were not the angular points but the sides or the radii themselves. (There is a parallel for this in Eucl. iv. 4.) (2) "The angle AC" is the angle which is the sum of A and C, and A means here the angle at A of the isosceles triangle shown in the figure, and afterwards spoken of by Aristotle as E, while C is the " mixed " angle between AB and the circumference of the smaller segment cut off by it. (3) The "angle of a. semicircle" (i.e. the "angle" between the diameter and the circumference, at the extremity of the diameter) and the " angle of a segment" appear in Euclid tn. 16 and 111. Def. 7 respectively, obviously as survivals from earlier text-books. But the most significant facts to be gathered from the extract are that in the text-books which preceded Euclid's " mixed " angles played a much more important part than they do with Euclid, and, in particular, that at least two propositions concerning such angles appeared quite at the beginning, namely the propositions that the (mixed) angles of semicircles art equal and that the two (mixed) angles of any segment of a circle art equal. The wording of the first of the two propositions is vague, but it does not necessarily mean more than that the two (mixed) angles in one semicircle are equal, and I know of no evidence going to show that it asserts that the angle of any one semicircle is equal to the angle of any other semicircle (of different size). It is quoted in the same form, " because the angles of semicircles are equal," in the Latin translation from the Arabic of Heron's Catopirica, Prop. 9 (Heron, Vol. 11., Teubner, p. 334), but it is only inferred that the different radii of one circle make equal "angles" with the circumference; and in the similar proposition of the Pseudo-Euclidean Catoptriea (Euclid, Vol. vn., p. 394) angles of the same sort in one circle are said to be equal " because they are (angles) of a semicircle." Therefore the first of the two propositions may be only a particular case of the second. But it is remarkable enough that the second proposition (that the two "angles of" any segment of a circle art equal) should, in earlier text-books, have been placed before the theorem of Eucl. 1. 5. We can hardly suppose it to have been proved otherwise than by the superposition of the semicircles into which the circle is divided by the diameter which bisects at right angles the base of the segment; and no doubt the proof would be closely connected with that of Thales' other proposition that any diameter of a circle bisects it, which must also (as Proclus indicates) have been proved by superposing one of the two parts upon the other. It is a natural inference from the passage of Aristotle that Euclid's proof of *54 BOOK I [i. s i. 5 was his own, and it would thus appear that his innovations as regards order of propositions and methods of proof began at the very threshold of the subject. Proof without producing the sides. In this proof, given by Proclus (pp. 148, 21—249, J 9)> & an ^ E are ta ^en on AS, AC, instead of on AB, AC produced, so that AD, AE&te equal. The method of proof is of course exactly like Euclid's, but it does not establish the equality of the angles beyond the base as well. Pappus* proof. Proclus (pp. 249, 20 — 250, 1 2) says that Pappus proved the theorem in a still shorter manner without the help of any construction whatever. This very interesting proof is given as follows : " Let ABC be an isosceles triangle, and AB equal to AC. Let us conceive this one triangle as two triangles, and let us argue in this way. Since AB is equal to AC, and AC to AB, the two sides AB, AC are equal to the two sides AC, AB. And the angle BA C is equal to the angle CAB, for it is the same. Therefore all the corresponding parts (in the triangles) are equal, namely BC to BC, the triangle ABC to the triangle ABC (i.e. ACB), the angle ABC to the angle ACB, and the angle ACB to the angle ABC, (for these are the angles subtended by the equal sides AB, A C. Therefore in isosceles triangles the angles at the base are equal." This will no doubt be recognised as the foundation of the alternative proof frequently given by modern editors, though they do not refer to Pappus. But they state the proof in a different form, the common method being to suppose the triangle to be taken up, turned over, and placed again upon itself, after which the same considerations of congruence as those used by Euclid in 1. 4 are used over again. There is the obvious difficulty that it supposes the triangle to be taken up and at the same time to remain where it is, (Cf. Dodgson's humorous remark upon this, Euclid and Ail modern Rivals, p. 47.) Whatever we may say in justification of the proceeding (e.g. that the triangle may be supposed to leave a tract), it is really equivalent to assuming the construction (hypothetical, if you will) of another triangle equal in all respects to the given triangle ; and such an assumption is not in accordance with Euclid's principles and practice. It seems to me that the form given to the proof by Pappus himself is by far the best, for the reasons (i) that it assumes no construction of a second triangle, real or hypothetical, (2) that it avoids the distinct awkwardness involved by a proof which, instead of merely quoting and applying the result of a previous proposition, repeats, with reference to a new set of data, the process by which that result was established. If it is asked how we are to realise Pappus' idea of two triangles, surely we may answer that we keep to one triangle and merely view it in two aspects. If it were a question of helping a beginner to understand this, we might say that one triangle is the triangle i. 5, 6] PROPOSITIONS 5, 6 355 looked at in front and that the other triangle is the same triangle looked at from behind ; but even this is not really necessary. Pappus' proof, of course, does not include the proof of the second part of the proposition about the angles under the base, and we should still have to establish this much in the same way as Euclid does. Purpose of the second part of the theorem. An interesting question arises as to the reason for Euclid's insertion of the second part, to which, it will be observed, the converse proposition 1. 6 has nothing corresponding. As a matter of fact, it is not necessary for any subsequent demonstration that is to be found in the original text of Euclid, but only for the interpolated second case of 1. 7 ; and it was perhaps not unnatural that the undoubted genuineness of the second part of 1. 5 convinced many editors that the second case of 1. 7 must necessarily be Euclid's also. Proctus' explanation, which must apparently be the right one, is that the second part of 1. 5 was inserted for the purpose of fore-arming the learner against a possible objection (frimturif), as it was technically called, which might be raised to 1. 7 as given in the text, with one case only. The objection would, as we have seen, take the specific ground that, as demonstrated, the theorem was not conclusive, since it did not cover all possible cases. From this point of view, the second part of 1. 5 is useful not only for 1. 7 but, according to Proclus, for 1. 9 also. Simson does not seem to have grasped Proclus' meaning, for he says : " And Proclus acknowledges, that the second part of Prop. 5 was added upon account of Prop. 7 but gives a ridiculous reason for it, 'that it might afford an answer to objections made against the 7th,' as if the case of the 7th which is left out were, as he expressly makes it, an objection against the proposition itself." Proposition 6. If in a triangle two angles be equal to one another, the sides which subtend the equal angles will also be equal to one another. Let ABC be a triangle having the angle ABC equal to the angle ACS', I say that the side AB is also equal to the side AC. For, if AB is unequal to AC, one of them is greater. Let AB be greater; and from AB the greater let DB be cut off equal to AC the less ; let DC be joined. Then, since DB is equal to AC, and BC is common, the two sides DB, BC are equal to the two sides AC, CB respectively ; *$6 BOOK I [i. 6 and the angle DBC is equal to the angle ACB ; therefore the base DC is equal to the base AB, and the triangle DBC will be equal to the triangle ACB, the less to the greater : which is absurd. Therefore AB is not unequal to AC; it is therefore equal to it. Therefore etc Q. E. D. Euclid assumes that, because D is between A and B, the triangle DBC is less than the triangle ABC. Some postulate is necessary to justify this tacit assumption; considering an angle less than two right angles, say the angle ACB in the figure of the proposition, as a cluster of rays issuing from C and bounded by the rays CA f CB, and joining AB (where A, B are any two points on CA, CB respectively), we see that to each successive ray taken in the direction from CA to CB there corresponds one point on AB in which the said ray intersects AB, and that all the points on AB taken in order from A to B correspond uni vocally to all the rays taken in order from CA to CB, each point namely to the ray intersecting AB in the point. We have here used, for the first time in the Elements, the method of redact in ad absurdum, as to which I would refer to the section above (pp. 136, 140) dealing with this among other technical terms. This proposition also, being the converse of the preceding proposition, brings us to the subject of Geometrical Conversion. This must of course be distinguished from the logical conversion of a proposition. Thus, from the proposition that alt isosceles triangles have the angles opposite to the equal sides equal, logical conversion would only enable us to conclude that some triangles with two angles equal are isosceles. Thus 1. 6 is the geometrical, but not the logical, converse of 1. 5. On the other hand, as De Morgan points out (Companion to the Almanac, 1849, p. 7), 1. 6 is a purely logical deduction from t. 5 and 1. 18 taken together, as is 1. 19 also. For the general argument see the note on 1. 19. For the present proposition it is enough to state the matter thus. Let X denote the class of triangles which have the two sides other than the base equal, Y the class of triangles which have the base angles equal ; then we may call aan-X the class of triangles having the sides other than the base unequal non- Y the class of triangles having the base angles unequal. Thus we have aii x is r, [1. 5] All aon-X is non-K; [1. 18) and it is a purely logical deduction that All Y is X. [1. 6] According to Proclus (p. 252, 5 sqq.) two forms of geometrical conversion were distinguished. (1) The leading form {rpo^yovji.iyin). the conversion par excellence (rj xvpum t. 6] PROPOSITION 6 457 awKrrpo^itf), is the complete or simple conversion in which the hypothesis and the conclusion of a theorem change places exactly, the conclusion of the theorem being the hypothesis of the converse theorem, which again establishes, as its conclusion, the hypothesis of the original theorem. The relation between the first part of 1. 5 and 1. 6 is of this character. In the former the hypothesis is that two sides of a triangle are equal and the conclusion is that the angles at the base are equal, while the converse {1. 6) starts from the hypothesis that two angles are equal and proves that the sides subtending them are equal. (2) The other form of conversion, which we may call partial, is seen in cases where a theorem starts from two or more hypotheses combined into one enunciation and leads to a certain conclusion, after which the converse theorem takes this conclusion in substitution for one of the hypotheses of the original theorem and from the said conclusion along with the rest of the original hypotheses obtains, as its conclusion, the omitted hypothesis of the original theorem, r, 8 is in this sense a converse proposition to 1. 4 ; for 1. 4 takes as hypotheses ( 1 ) that two sides in two triangles are respectively equal, (a) that the included angles are equal, and proves (3) that the bases are equal, while 1. 8 takes {1) and (3) as hypotheses and proves (2) as its conclusion. It is clear that a conversion of the leading type must be unique, while there may be many partial conversions of a theorem according to the number of hypotheses from which it starts. Further, of convertible theorems, those which took as their hypothesis the genus and proved a, property were distinguished as the leading theorems (rponfyovptva), while those which started from the property as hypothesis and described, as the conclusion, the genus possessing that property were the converse theorems. 1. 5 is thus the leading theorem and [. 6 its converse, since the genus is in this case taken to be the isosceles triangle. Converse of second part of I. 5. Why, asks Proclus, did not Euclid convert the second part of I. 5 as well ? He suggests, properly enough, two reasons: (1) that the second part of 1. 5 itself is not wanted for any proof occurring in the original text, but is only put in to enable objections to the existing form of later propositions to be met, whereas the converse is not even wanted for this purpose ; (2) that the converse could be deduced from t. 6, if wanted, at any time after we have passed 1. 13, which can be used to prove that, if the angles formed by producing two sides of a triangle beyond the base are equal, the base angles themselves are equal. Proclus adds a proof of the converse of the second part of 1. 5. i.e. of the proposition that, if the angles formed by producing two sides of_ a triangle beyond the base are equal, the triangle a is isosceles ; but it runs to some length and then only „ A effects a reduction to the theorem of 1. 6 as we have it. AA As the result of this should hardly be assumed, a better / \\ proof would be an independent one adapting Euclid's / ^ own method in 1. 6. Thus, with the construction of 1. 5, l"~^~~^\ we first prove by means of 1. 4 that the triangles BFC, /-^"^*^~^\ CGB are equal in all respects, and therefore that FC is y~ "^^ equal to GB, and the angle BFC equal to the angle CGB. D g Then we have to prove that AF, AG are equal. If they are not, let AF be the greater, and from FA cut off FH equal to GA. Join CH. i S 8 BOOK I [i.«,T Then we have, in the two triangles HFC, AGB, two sides HF, FC equal to two sides AG, GB and the angle HFC equal to the angle AGB. Therefore (l 4) the triangles HFC, AGB are equal. But the triangles BFC, CGB are also equal Therefore {if we take away these equals respectively) the triangles HBC, ACB are equal: which is impossible. Therefore AF, AG are not unequal. Hence AF'm equal to AG and, if we subtract the equals BF, CG respec- tively, AB is equal to A C. This proof is found in the commentary of an-Nairlzi (ed. Besthom-Heiberg, p. 61 ; ed. Curtze, p. 50). Alternative proofs of I. 6. Todhunter points out that I. 6, not being wanted till II. 4, could be postponed till later and proved by means of i. 26. Bisect the angle BAC by a straight line meeting the base at D. Then the triangles ABD, A CD are equal in all respects. Another method depending on 1. 26 is given by an-Nairlzi after that proposition. Measure equal lengths BD, CE along the sides BA, CA. Join BE, CD. Then [1. 4] the triangles DBC, ECB are equal in all respects ; therefore EB, DC are equal, and the angles BEC, CDB are equal. The supplements of the latter angles are equal [1. 13], and hence the triangles ABE, A CD have two angles equal respectively and the side BE equal to the side CD. Therefore [1. 26] AB is equal to AC. P HO POSITION 7. Given two straight lines constructed on a straight line [from its extremities) and meeting in a point, there cannot be constructed on the same straight line {from its extremities), and on the same side of it, two other straight lines meeting in 5 another point and equal to the former two respectively, namely each to that which has the same extremity with it. For, if possible, given two straight lines AC, CB con- structed on the straight line AB and meeting at the point C, let two other straight lines 10 AD, DB be constructed on the same straight line AB, on the same side of it, meeting in another point D and equal to the former two respectively, namely each to that which has the same extremity with it, so that CA is 15 equal to DA which has the same extremity A with it, and I. 7] PROPOSITIONS 6, 7 259 CB to DB which has the same extremity B with it ; and let CD be joined. Then, since AC is equal to AD, the angle A CD is also equal to the angle ADC; [1. 5] 20 therefore the angle ADC is greater than the angle DCB ; therefore the angle CDB is much greater than the angle DCB. Again, since CB is equal to DB, the angle CDB is also equal to the angle DCB. 25 But it was also proved much greater than it : which is impossible. Therefore etc. Q. E. D. 1 — 6. In an English translation of the enunciation of this proposition it is absolutely necessary, in order to make it intelligible, to insert some words which are not in the Greek. The reason is partly that the Greek enunciation is itself very elliptical, and partly that some words used in it conveyed more meaning than the corresponding words in English do. Particularly is this the case with oC evaratrboorrat iri "there shall not be constructed upon," since evrUraaSat. is the regular word for constructing a triangle in particular. Tbus a Greek would easily understand avaraB-fyrarrat iri as meaning the construction of two lines forming a triangle on a given straight line as base; whereaa.ro "construct two straight lines on a straight line " is not in English sufficiently definite unless we explain that they are drawn from the ends of the straight line to meet at a point. I have had the less hesitation in putting in the words "from its extremities" because they are actually used by Euclid in the somewhat similar enunciation of 1. at. How impossible a literal translation into English is, if it is to convey the meaning of the enunciation intelligibly, will be clear from the following attempt to render literally: "On the same straight line there shall not be constructed two other straight lines equal, each to each, to the same two straight lines, (terminating) at different points on the same side, having the same extremities as the original straight lines " (irl riji a*rj)i eiBctai S60 raft airaii tiStlmt AXXor ivo evSttat focu inar4pa ixarlpa ov ewTaBfaairrai vpdr dWt^j rtai AXX^I OTjfuU^ i-wl f A ovrd pift r« a*"k ripara (x'oeai roll it Apx*)' tiltlaa). The reason why Euclid allowed himself to use, in this enunciation, language apparently so obscure is no doubt that the phraseology was traditional and therefore, vague as it was, had a conventional meaning which the contemporary geometer well understood. This is proved, I think, by the occurrence in Aristotle (Meteorologica 111, 5, 376 a j sqq.) of the very same, evidently technical, expressions. Aristotle is there alluding to the theorem given by Eutocius from Apollonius' Plane Loci to the effect that, if H, K be two fixed points and M such a variable point that the ratio of MH to MK is a given ratio (not one of equality), the locus of M is a circle. (For an account of this theorem see note on vt. 3 below.) Now Aristotle says "The lines drawn up from H, K in this ratio cannot be constructed to two different points of the semicircle A " (of ttr iri tCh HK inayb/urai ypawial tr roirip rif \t>yy uLr ffvffTad^aovTat rov £q> y A i^uxurXfou rpbt AWo teal A\\o a-rjfLtiov). If a paraphrase is allowed instead of a translation adhering as closely as possible to the original, Simson's is the best that could be found, since the fact that the straight lines form triangles on the same base is really conveyed in the Greek. Simson's enunciation is, Upon the same base, and on the same side of i(, there cannot lie two triangles line have their sides which are terminated in one extremity of the iate equal to one another, and liiewue these which art terminated at the other extremity. Th. Taylor (the translator of Proclus) attacks Simson's alteration as "indiscreet" and as detracting from the beauty and accuracy of Euclid's enunciation which are enlarged upon by Proclus in his commentary. Yet, when Taylor says "Whatever difficulty learners may find in conceiving this proposition abstractedly is easily removed by its exposition in the figure," he really gives his case away. The fact is that Taylor, always enthusiastic over his author, was nettled by Simson's slighting remarks on Proclus' comments on the proposition. Simson had said, with reference to Proclus' explanation of the bearing of the second part of 1. % on 1. 7, that it was not "worth while j6o BOOK t [i. 7 to relate his [rifles at full length," to which Taylor retorts "But Mr Simson was no philosopher ; and therefore the greatest part of these Commentaries must be considered by him as trifles, from the want of a philosophic genius to comprehend their meaning, and a taste superior to that of a mere mathematician, to discover their beauty and elegance." 10. It would be natural to insert here the step "but the angle ACD is greater than the angle BCD. [C. N. 5.3" tl, much greater, literally ''greater by much" (roXX^i fielfav). Simson and those who follow him translate : " much mart then is the angle BDC greater than the angle BCD," but the Greek for this would have to be toXXv (or iroXii) jiaX \i r tm...JU(JW>. troXX^ ^aXXof, however, though used by Apotlonius, is not, apparently, found in Euclid or Archimedes. Just as in I. 6 we need a Postulate to justify theoretically the statement that CD falls within the angle ACB, so that the triangle DBC is less than the triangle ABC, so here we need Postulates which shall satisfy us as to the relative positions of CA, CB, CD on the one hand and of DC, DA, DB on the other, in order that we may be able to infer that the angle BDC is greater than the angle ADC, and the angle ACD greater than the angle BCD, De Morgan {sp. cit. p. 7) observes that 1. 7 would be made easy to beginners if they were first familiarised, as a common notion, with " if two magnitudes be equal, any magnitude greater than the one is greater than any magnitude less than the other." I doubt however whether a beginner would follow this easily ; perhaps it would be more easily apprehended tn the form "if any magnitude A is greater than a magnitude B, the magnitude A is greater than any magnitude equal to B, and (a fortiori) greater than any magnitude less than B." It has been mentioned already (note on 1. 5) that the second case of 1. 7 given by Simson and in our text-books generally is not in the original text (the omission being in accordance with Euclid's general practice of giving only one case, and that the most difficult, and leaving the others to be worked out by the reader for himself). The second case is given by Proclus as the answer to a possible objection to Euclid's proposition, which should assert that the proposition is not proved to be universally true, since the proof given does not cover all possible cases. Here the objector is supposed to contend that what Euclid declares to be impossible may still be possible if one pair of lines lie wholly within the other pair of lines; and the second part of 1. 5 enables the objection to be refuted. If possible, let AD, DB be entirely within the triangle formed by AC, CB with AB, and let AC be equal to AD and BC to BD. Join CD, and produce AC, AD to E and F. Then, since AC is equal to AD, o( the triangle ACD is isosceles, and the angles ECD, FDC under the base are equal. But the angle ECD is greater than the angle BCD , therefore the angle FDC is also greater than the angle BCD. Therefore the angle BDC is greater by far than the angle BCD. Again, since DB is equal to CB, the angles at the base of the triangle BDC are equal, [1. 5] that is, the angle BDC is equal to the angle BCD. Therefore the same angle BDC is both greater than and equal to the angle BCD: which is impossible. The case in which D falls on AC or BC does not require proof, i. 7. 8] PROPOSITIONS 7. 8 *6i I have already referred (note on 1. r) to the mistake made by those editors who regard I. 7 as being of no use except to prove 1. 8. What 1. 7 proves is that if, in addition to the base of a triangle, the length of the side terminating at each extremity of the base is given, only one triangle satisfying these conditions can be constructed on one and the same side of the given base. Hence not only does 1. 7 enable us to prove 1. 8, but it supplements 1. 1 and 1. 22 by showing that the constructions of those propositions give one triangle only on one and the same side of the base. But for [. 7 this could not be proved except by anticipating in, 10, of which therefore 1. 7 is the equivalent for Book 1. purposes. Dodgson (Etttfid and his modern Rivals, pp. 194 — 5) puts it in another way. " It [l. 7] shows that, of airplane figures that can be made by hingeing rods together, the /Aree-sided ones (and these only) are rigid (which is another way of stating the fact that there cannot be two such figures on the same base). This is analogous to the fact, in relation to solids contained by plane surfaces hinged together, that any such solid is rigid, there being no maximum number of sides. And there is a close analogy between 1. 7, 8 and in. 23, 24. These analogies give to geometry much of its beauty, and I think that they ought not to be lost sight of." It will therefore be apparent how ill-advised are those editors who eliminate 1. 7 altogether and rely on Philo's proof for 1. 8. Proclus, it may be added, gives (pp. 2 68, 19 — 269, 10) another explanation of the retention of I. 7, notwithstanding that it was apparently only required for 1. 8. It was said that astronomers used it to prove that three successive eclipses could not occur at equal intervals of time, i.e. that the third could not follow the second at the same interval as the second followed the first ; and it was argued that Euclid had an eye to this astronomical application of the proposition. But, as we have seen, there are other grounds for retaining the proposition which are quite sufficient of themselves. Proposition 8. If two triangles have the two sides equal to two sides respectively, and have also the base equal to the base, they will also have the angles equal which are contained by the equal straight lines. s Let ABC, DBF be two triangles having the two sides AB, AC equal to the two sides DE, DF respectively, namely AB to DE, and AC to DF\ and let them have the base BC equal 10 to the base EF ; I say that the angle BAC is also equal to the angle EDF. For, if the triangle ABC be applied to the triangle DEF, and if the point B be placed on 15 the point E and the straight line BC on EF, the point C will also coincide with F, because BC is equal to EF. 2 62 BOOK I [l8 Then, BC coinciding with EF, BA, AC will also coincide with ED, DF; ao for, if the base BC coincides with the base EF, and the sides BA, AC do not coincide with ED, DF but fall beside them as EG, GF, then, given two straight lines constructed on a straight line (from its extremities) and meeting in a point, there will »5 have been constructed on the same straight line (from its extremities), and on the same side of It, two other straight lines meeting in another point and equal to the former two respectively, namely each to that which has the same extremity with it 30 But they cannot be so constructed. [1. 7} Therefore it is not possible that, if the base BC be applied to the base EF, the sides BA, AC should not coincide with ED, DF; they will therefore coincide, 35 so that the angle BAC will also coincide with the angle EDF t and will be equal to it. If therefore etc. q. e. d. 19. BA, AC. The text has here " BA, CA." SI, fall be aide them. The Greek has the future, i-apoXXdfoLvf. TapaXk&m* means "to pass by without touching," "to miss" or "to deviate." As pointed out above (p. 157) 1, 3 is a par tied converse of t. 4. It is to be observed that in I. 8 Euclid is satisfied with proving the equality of the vertical angles and does not, as in 1. 4, add that the triangles are equal, and the remaining angles are equal respectively. The reason is no doubt (as pointed out by Proclus and by Savile after him) that, when once the vertical angles are proved equal, the rest follows from 1. 4, and there is no object in proving again what has been proved already. Aristotle has an allusion to the theorem of this proposition in Me&orokgka in. 3, 373 a 5 — 16. He is speaking of the rainbow and observes that, if equal rays be reflected from one and the same point to one and the same point, the points at which reflection takes place are on the circumference of a circle. "For let the broken lines ACB, AFB, ADB be all reflected from the point A to the point B (in such a way that) AC, AF, AD are all equal to one another, and the lines {terminating) at B, i.e. CB, FB, DB, are likewise all equal ; and let AEB be joined. It follows that the triangles art equal; for they are upon the equal (base) AEB." Heiberg {Mathtmatisehes tu Aristolelts, p. 18) thinks that the form of the conclusion quoted is an indication that in the corresponding proposition tc. Eucl. 1. 8, as it lay before Aristotle, it was maintained that the triangles were equal, and not only the angles, and "we see here therefore, in a clear example, how the stones of the ancient fabric were recut for the rigid structure of his i 8] PROPOSITION 8 a6j Elements. " I do not, however, think that this inference from Aristotle's language as to the form of the pre- Euclidean proposition is safe. Thus if we, nowadays, were arguing from the data in the passage of Aristotle, we should doubtless infer directly that the triangles are equal in all respects, quoting I 8 alone. Besides, Aristotle's language is rather careless, as the next sentences of the same passage show. "Let perpendiculars," he says, " be drawn to AEB from the angles, CE from C, FE from ^and DE from D. These, then, are equal ; for they are all in equal triangles, and in one plane ; for all of them are perpendicular to AEB, and they meet at one point E. There- fore the (line) drawn (through C, F, D) will be a circle, and its centre (will be) E." Aristotle should obviously have proved that the three perpendiculars will meet at one point E cm AEB before he spoke of drawing the perpendiculars CE, FE, DE. This of course follows from their being "in equal triangles" {by means of EucL i. 26); and then, from the fact that tbe perpendiculars meet at one point on AB, it can be inferred that all three are in one plane. Philo's proof of I. 8. This alternative proof avoids the use of 1. 7, and it is elegant ; but it is inconvenient in one respect, since three cases have to be distinguished. Proctus gives the proof in the following order (pp. 266, 15 — 168, 14). I-et ABC, DEF be two triangles having the sides AB, A C equal to the sides DE, DE respectively, and the base BC equal to the base EF. Let the triangle ABC be applied to the triangle DEF, so that B is placed on E and BC on EF, but so that A falls on the opposite side of EF from D, taking the position G. Then C will coincide with F, since BC is equal to EF. Now FG will either be in a straight line with DF, or make an angle with it, and in the tatter case the angle will either be interior (™ro to Ivtos) to the figure or exterior (no.™ to Item). I. Let FG be in a straight line with DF. Then, since DE is equal to EG-, and DFG is a straight line, DEG is an isosceles triangle, and the angle at D is equal to the angle at G. ['■ 5]. II. Let DF, FG form an angle interior to the figure. Let DG be joined. Then, since DE, EG are equal, the angle EDG is equal to the angle EGD. Again, since DF is equal to FG, the angle FDG is equal to the angle FGD. Therefore, by addition, the whole angle EDF is equal to the whole angle EGF. A i6 4 BOOK I [i. 8, 9 III. Let DF, FG form an angle ex/trier to the figure. Let DG be joined. The proof proceeds as in the last case, except that subtraction takes the place of addition, and the remaining angle EDF is equal to the remaining angle ECF. Therefore in all three cases the angle EDF is equal to the angle EGF, that is, to the angle BAC. It wrill be observed that, in accordance with the practice of the Greek geometers in not recognising as an "angle" any angle not less than two right angles, the re-entrant angle of the quadrilateral JDEGF'm. ignored and the angle DFG is said to be outside the figure. Proposition 9. To bisect a given rectilineal angle. Let the angle BAC be the given rectilineal angle. Thus it is required to bisect it. Let a point D be taken at random on AB ; let AE be cut off from AC equal to AD ; [1. 3] let DE be joined, and on DE let the equilateral triangle DEF be constructed ; let AF be joined. I say that the angle BAC has been bisected by the straight line AF. For, since AD is equal to AE, and AF is common, the two sides DA, AF are equal to the two sides EA, AF respectively. And the base DF is equal to the base EF; therefore the angle DAF is equal to the angle EAF. [.. 8] Therefore the given rectilineal angle BAC has been bisected by the straight line AF. q, e. f. It will be observed from the translation of this proposition that Euclid does not say, in his description of the construction, that the equilateral triangle should be constructed on the side of HE opposite to A ; he leaves this to be inferred from his figure. There is no particular value in Proclus' explanation as to how we should proceed in case any one should assert that he could not recognise the existence of any space below DE. He supposes, then, the equilateral triangle described on the side of DE towards A, and hence has to consider three cases according as the vertex of the equilateral triangle falls on A, above A or below it. The second and third cases do not d'ffer >.«] PROPOSITIONS 8, 9 265 substantially from Euclid's. In the first case, where ADE is the. equilateral triangle constructed on DE, take any point F or\ AD, and from AE cut off AG equal to AF. Join DG, EF meeting in H\ and join AH. Then AH is the bisector required. Proclus also answers the possible objection that might be raised to Euclid's proof on the ground that it assumes that, if the equilateral triangle be described on the side of DE opposite to A, its vertex .f will lie within the angle BAC. The objector is supposed to argue that this is not necessary, but that F might fall either on one of the lines forming the angle or outside it altogether. The two cases are disposed of thus. Suppose Fxx> fall as shown in the two figures below respectively. Then, since FD is equal to FE, the angle FDE is equal to the angle FED. Therefore the angle CED is greater than the angle FDE ; and, in the second figure, a fortiori, the angle CED is greater than the angle BDE. But, since ADE is an isosceles triangle, and the equal sides are produced, the angles under the base are equal, i.e., the angle CED is equal to the angle BDE. But the angle CED was proved greater : which is impossible. Here then is the second case in which, in Proclus' view, the second part of i. 5 is useful for refuting objections. On this proposition Proclus takes occasion (p. 27 r, rj — 19) to emphasize the fact that the given angle must be rectilineal, since the bisection of any sort of angle (including angles made by curves with one another or with straight lines) is not matter for an elementary treatise, besides which it is questionable whether such bisection is always possible. "Thus it is difficult to say whether it is possible to bisect the so-called horn-like angle " (formed by the circumference of a circle and a tangent to it). Trisection of an angle. Further it is here that Proclus gives us his valuable historical note about the trisection of any acute angle, which (as well as the division of an angle in any given ratio) requires resort to other curves than circles, i.e. curves of the species which, after Geminus, he calls "mixed." "This," he says (p. 372, 1 — 12), "is shown by those who have set themselves the task of trisecting such a given rectilineal angle. For Nicomedes trisected any rectilineal angle by means of the conchoidal lines, the origin, order, and properties of which he has handed down to us, being himself the discoverer of their peculiarity. Others have done the same thing by means of the quadratrices of Hippias and Nicomedes, thereby again using 'mixed' curves. But others, starting from the Archimedean spirals, cut a given rectilineal angle in a given ratio." i66 BOOK I [1.9 (a) T riscction by means of the conchoid. I have already spoken of the conchoid of Nicomedes {note on Def. t, pp. 160 — i); it remains to show how it could be used for trisecting an angle. Pappus explains this (iv. pp. 274 — 5) as follows. Let ABC be the given acute angle, and from any point A in AB draw A C perpendicular to BC. B O Complete the parallelogram FBCA and produce FA to a point E such that, if BE be joined, BE intercepts between AC and AE a length DE equal to twice AB. I say that the angle EBC is one-third of the angle ABC. For, joining A to G, the middle point of DE, we have the three straight lines AG, DG, EG equal, and the angle AGO is double of the angle A ED or EBC. But DE is double of AB ; therefore AG, which is equal to DG, is equal to AB. Hence the angle AGD is equal to the angle ABG. Therefore the angle ABD is also double of the angle EBC; so that the angle EBC is one-third of the angle ABC. So far Pappus, who reduces the construction to the drawing of BE so that DE shall be equal to twice AB. This is what the conchoid constructed with B as pole, AC 'as directrix, and distance equal to twice AB enables us to do ; for that conchoid cuts AE in the required point E. (6) Use of the quadratrix. The plural quadratrices in the above passage is a Hellenism for the singular quadratrix, which was a curve discovered by Hippias of El is about 420 B.C. According to Prod us (p. 356, 11) Hippias proved its properties; and we are told (1) in the passage quoted above that Nicomedes also investigated it and that it was used for trisecting an angle, and (2) by Pappus (iv. pp. 350, 33— 353, 4) that it was used by Dinostratus and Nicomedes and some more recent writers for squaring the circle, whence its name. It is described thus (Pappus iv. p. 352). Suppose that ABCD is a square and BED a quadrant of a circle with centre A. Suppose (1) that a radius of the circle moves uniformly about A from the position AB to the position AD, and (*) that in the same time the line BC moves uniformly, always parallel to itself, and with its extremity B moving along BA, from the position BC to the position AD. Then the radius AE and the moving line BC determine at any instant by their intersection a point F, The locus of F'\s the quadratrix. I. 9, 10] PROPOSITIONS o, 10 167 The property of the curve is that, if F is any point, the arc BED is to the arc ED as AB is to FH. In other words, if ^ is the angle FAD, p the radius vector AFa.nd a the side of the square, (p sin $)/« = $/Jt. Now the angle EAD can not only be trisected but divided in any given ratio by means of the quadratrix (Pappus iv. p. 386). For let FJfbe divided at JCin the given ratio. Draw KL parallel to AD, meeting the curve in L ; join AL and produce it to meet the circle in N. Then the angles EAN, NAD are in the ratio of FK to KH, as is easily proved. (e) Use of the spiral of Archimedes. The trisect ion of an angle, or the division of an angle in any ratio, by means or the spiral of Archimedes is of course an equally simple matter. Suppose any angle included between the two radii vectores OA and OB of the spiral, and let it be required to cut the angle AOB in a given ratio. Since the radius vector increases proportionally with the angle described by the vector which generates the curve (reckoned from the original position of the vector coinciding with the initial line to the particular position assumed), we have only to take the radius vector OB (the greater of the two OA, OB), mark off OC along it equal to OA, cut CB in the given ratio (at D say}, and then draw the circle with centre and radius OD cutting the spiral in E. Then OE will divide the angle AOB in the required manner. Proposition 10. To bisect a given finite straight line. Let AB be the given finite straight line. Thus it is required to bisect the finite straight line AB. Let the equilateral triangle ABC be constructed on it, [1. r] and let the angle ACB be bisected by the straight line CD ; ft, 9] I say that the straight line AB has been bisected at the point D. . For, since AC is equal to CB, and CD is common, the two sides A C, CD are equal to the two sides BC, CD respectively ; and the angle A CD is equal to the angle BCD ; therefore the base AD is equal to the base BD. [1. 4] Therefore the given finite straight line AB has been bisected at D. q. e, f. 268 BOOK I [>■ A poll on i us, we are told (P rod us, pp. 279, 16—280, 4), bisected a straight line AB by a construction tike that of 1. 1. With centres A, B, and radii AB, BA respec- tively, two circles are described, intersecting in C, p. Joining CD, AC, CB, AD, DB, Apoi- lonius proves in two steps that CD bisects AB. (1) Since, in the triangles A CD, BCD, two sides AC, CD are equal to two sides BC, CD, and the bases AD, BD are equal, the angle A CD is equal to the angle BCD. [1.8] (2) The latter angles being equal, and AC being equal to CB, while CE is common, the equality of AE, EB follows by I. 4. The objection to this proof is that, instead of assuming the bisection of the angle ACB, as already effected by 1. 9, Apollo nius goes a step further back and embodies a construction for bisecting the angle. That is, he unnecessarily does over again what has been done before, which is open to objection from a theoretical point of view. Proclus (pp. 277, 25 — 279, 4) warns us against being moved by this proposition to conclude that geometers assumed, as a preliminary hypothesis, that a line is not made up of indivisible parts (1$ AjttpAr). This might be argued thus. If a line is made up of indivisibles, there must be in a finite line either an odd or an even number of them. If the number were odd, it would be necessary in order to bisect the line to bisect an indivisible (the odd one). In that case therefore it would not be possible to bisect a straight line, if it is a magnitude made up of indivisibles. But, if it is not so made up, the straight line can be divided ad infinitum or without limit (hf irtipov faatpttTai). Hence it was argued (^acrtV), says Proclus, that the divisibility of magnitudes without limit was admitted and assumed as a geometrical principle. To this he replies, following Geminus, that geometers did indeed assume, by way of a common notion, that a continuous magnitude, i.e. a magnitude consisting of parts connected together (owrjfijtow), is divisible (SuuptTov). But infinite divisibility was not assumed by them ; it was proved by means of the first principles applicable to the case. "For when," he says, " they prove that the incommensurable exists among magnitudes, and that it is not all things that are commensurable with one another, what else will any one say that they prove but that every magnitude can be divided for ever, and that we shall never arrive at the indivisible, that is, the least common measure of the magnitudes? This then is matter of demonstration, whereas it is an axiom that everything continuous is divisible, so that a finite continuous line is divisible. The writer of the Elements bisects a finite straight line, starting from the latter notion, and not from any assumption that it is divisible without limit" Proclus adds that the proposition may also serve to refute Xenocrates' theory of indivisible lines (di-ojiu* ypapjiat). The argument given by Proclus to disprove the existence of indivisible lines is substantially that used by Aristotle as regards magnitudes generally (cf Physkt vi. 1, 231 a 21 sqq. and especially vl. 2, 133 b 15 — 32). J. n] PROPOSITIONS io, u 269 Proposition ii. To draw a straight line at right angles to a given straight line from a given point on it. Let AB be the given straight line, and C the given point on it. s Thus it is required to draw from the point C a straight line at right angles to the straight line AB. Let a point D be taken at ran- dom on AC; 10 let CE be made equal to CD ; [1. 3] on DE let the equilateral triangle FDE be constructed, [1. i] and let FC be joined ; I say that the straight line FC has been drawn at right is angles to the given straight line AB from C the given point on it. For, since DC is equal to CE, and CF is common, the two sides DC, CF are equal to the two sides EC, 20 CF respectively ; and the base DF is equal to the base FE ; therefore the angle DCF is equal to the angle ECF\ M] and they are adjacent angles. But, when a straight line set up on a straight line makes »s the adjacent angles equal to one another, each of the equal angles is right ; [Def. 10] therefore each of the angles DCF, FCE is right. Therefore the straight line CF has been drawn at right angles to the given straight line AB from the given point 30 C on it. Q. E. F. 10. let CB be made equal to CD. The verb is mbrSu which, as welt as the othet parts of Ktitiai, a constant iy used lor the passive of rWiftu " lo plats " ; and the latter word it constantly used in the sense of making, e.g., one straight line equal to another straight line. 1 )e Morgan remarks that this proposition, which is " to bisect the angle made by a straight line and its continuation " [i.e. a flat angle], should be a particular case of 1. 9, the constructions being the same. Thjs is certainly J7o BOOK I ['■ worth noting, though I doubt the advantage of rearranging the propositions in consequence. Apollonius gave a construction for this proposition (see P rod us, p. 282, 8) differing from Euclid's in much the same way as his construction for bisecting a straight line differed from that of 1. 10. Instead of assuming an equilateral triangle drawn without repeating the process of 1. 1, Apollonius takes D and E equidistant from C as in Euclid, and then draws circles in the manner of 1. 1 meeting at F This necessitates proving again that DF\s equal to FE\ whereas Euclid's assumption of the construction of 1. 1 in the words " let the equilateral triangle FDE be constructed " enables him to dispense with the drawing of circles and with the proof that DF is equal to FE at the same time. While however the substitution of Apollonius' constructions for 1. 10 and 1 1 would show faulty arrangement in a theoretical treatise like Euclid's, they are entirely suitable for what we call practical geometry, and such may have been Apollonius' object in these constructions and in his alternative for '- »3- . . . . Proclus gives a construction for drawing a straight line at right angles to another straight line but from one end of it, instead of from an intermediate point on it, it being supposed (for the sake of argument) that we are not permitted to produce the straight line. In the commentary of an-Nairizi (ed. Besthorn-Heiberg, pp. 73 — 4; ed. Curtze, pp. 54 — 5) this construction is attributed to Heron. Let it be required to draw from A a straight line at right angles to AB. On AB take any point C, and in the manner of the proposition draw CE at right angles to AB. from CE cut off CD equal to AC, bisect the angle ACE by the straight line CF\ [1. 9] and draw DF at right angles to CE meeting CF in F. Join FA. Then the angle FAC will be a right angle. For, since, in the triangles ACF, DCF, the two sides AC, CF are equal to the two sides DC, CF respectively, and the included angles ACF, DCFare equal, the triangles are equal in all respects. [1, 4] Therefore the angle at A is equal to the angle at D, and is accordingly a right angle. E F D A ( a PROPOSITION 12. To a given infinite straight tine, from a given point which is not on it, to draw a perpendicular straight line. Let AB be the given infinite straight line, and C the given point which is not on it ; I. u] PROPOSITIONS ii, 12 371 Sthus it is required to draw to the given infinite straight line AB, from the given point C which is not on it, a per- pendicular straight line. For let a point D be taken 10 at random on the other side of the straight line AB, and with centre C and distance CD let the circle EFG be described ; [Post. 3] let the straight line EG ij be bisected at H, [1. 10] and let the straight lines CG, CH, CE be joined. [Post. 1] I say that CH has been drawn perpendicular to the given infinite straight line AB from the given point C which is not on it. jo For, since GH is equal to HE, and HC is common, the two sides GH, HC are equal to the two sides EH, HC respectively ; and the base CG is equal to the base CE ; 35 therefore the angle CHG is equal to the angle EHC. [1. 8] And they are adjacent angles. But, when a straight line set up on a straight line makes the adjacent angles equal to one another, each of the equal angles is right, and the straight line standing on the other is jo called a perpendicular to that on which it stands. [Def. 10] Therefore CH has been drawn perpendicular to the given infinite straight line AB from the given point C which is not on it Q. E. F. 1. a perpendicular straight line, tiitrer ti$ttv ypapph'- This is the full expression for a ptrpendkutar, KiBtrm meaning >tt f) tries to refute this objection, and it is interesting to follow his argument, though it will easily be seen to be inconclusive. He takes in order three possible suppositions. 1. May not the circle meet AB in a third point K between the middle point of GE and either extremity of it, taking the form drawn in the figure appended ? Suppose this possible. Bisect GE in H. Join CH, and produce it to meet the circle in L. Join CG, CK, CE. Then, since CG is equal to CE, and CH is common, while the base GH is equal to the base HE, the angles CHG, CHE are equal and, since they are adjacent, they are both right. Again, since CG is equal to CE, the angles at G and E are equal. Lastly, since CK is equal to CG and also to CE, the angles CGK, CKG are equal, as also are the angles CKE, CEK. Since the angles CGK, CEK are equal, it follows that the angles CKG, CKE are equal and therefore both right. Therefore the angle CKH'\% equal to the angle CHK, and CH is equal to CK. I. 12] PROPOSITION 12 273 But CK is equal to CL, by the definition of the circle ; therefore CH is equal to CL : which is impossible. Thus Proclus; but why should not the circle meet AB in H as well as .X? 2. May not the circle meet AB in // the middle point of GE and take the form shown in the second figure? In that case, says Proclus, join CG, CH, CE as before. Then bisect ME at K, join CK and produce it to meet the circumference at L. Now, since HK is equal to KE, CK is common, and the base CH is equal to the base CE, the angles at K are equal and therefore both right angles. Therefore the angle CHK is equal to the angle CKH, whence CK is equal to CH and therefore to CL : which is impossible. So Proclus ; but why should not the circle meet AB in A* as well as Hf 3, May not the circle meet AB in two points besides G, E and pass, between those two points, to the side of A3 towards C, as in the next figure ? Here again, by the same method, Proclus proves that, K, L being the other two points in which the circle cuts AB, CK is equal to CH, and, since the circle cuts CH'm M, CM is equal to CK and therefore to CH: which is impossible. But, again, why should the circle not cut AB in the point H a& well? In fact, Proclus' cases are not mutually exclusive, and his method of proof only enables us to show that, if the circle meets AB in one more point besides G, E, it must meet it in more points still. We can always find a new point of intersection by bisecting the distance separating any two points of intersection, and so, applying the method ad infinitum, we should have to conclude ultimately that the circle with radius CH (or CG) coincides with AB. It would follow that a circle with centre C and radius greater than CH would not meet AB at all. Also, since all straight lines from C to points on AB would be equal in length, there would be an infinite number of perpendiculars from C on AB. Is this under any circumstances possible ? It is not possible in Euclidean space, but it is possible, under the Riemann hypothesis (where a straight line is a " closed series " and returns on itself), in the case where C is the pole of the straight line AB. It is natural therefore that, for a proof that in Euclidean space there is only one perpendicular from a point to a straight line, we have to wait until 1. 16, the precise proposition which under the Riemann hypothesis is only valid with a certain restriction and not universally. There is no difficulty involved by waiting until 1. 16, since t. 12 is not used before that proposition is reached; and we are only in the same position as when, in order to satisfy ourselves of the number of possible solutions of 1. r, we have to wait till 1. 7. But if we wish, after all, to prove the truth of the assumption without recourse to any later proposition than 1. 12, we can do so by means of this same invaluable 1. 7. »74 BOOK I [l, 12 If the circle intersects AB as before in G, E, let H be the middle point of GE, and suppose, if possible, that the circle also intersects AB in any other point K on AH. From H, on the side of AB opposite to C, draw HL at right angles to AB, and make HL equal to HC. Join CG, EG, CK, LK. Now, in the triangles CHG, EHG, CHis equal to /.//, and HG is common. Also the angles CHG, EHG, being both right, are equal. Therefore the base CG is equal to the base EG. Similarly we prove that CK is equal to LK. But, by hypothesis, since K is on the circle, CK is equal to CG. Therefore CG, CK, LG, LK are all equal. Now the next proposition, i. 13, will tell us that CH, HL are in a straight line; but we will not assume this. Join CE, Then on the same base CE and on the same side of it we have two pairs of straight lines drawn from C, L to G and K such that CG is equal to CK and EG to LK But this is impossible [1. 7 \ Therefore the circle cannot cut BA or BA produced in any point other than G on that side of CL on which G is. Similarly it cannot cut AB or AB produced at any point other than E on the other side of CL. The only possibility le/t therefore is that the circle might cut AB in the same point as that in which CL cuts it. But this is shown to be impossible by an adaptation of the proof of I. 7, For the assumption is that there may be some point M on CL such thai CM is equal to CG and LM to LG. If possible, let this be the case, and produce CG to N. Then, since CM is equal to CG, the angle NGM is equal to the angle GML [1. 5, part 2]. Therefore the angle GML is greater than the angle MGL. Again, since LG k equal to LM, the angle GML is equal to the angle MGL. But it was also greater : which is impossible. Hence the circle in the original figure cannot cut AB in the point in which CL cuts it. Therefore the circle cannot cut AB in any point whatever except G and E. [This proof of course does not prove that CK is less than CG, but only that it is not equal to it. The proposition that, of the obliques drawn from C to AB, that is less the foot of which is nearer to H can only be proved later. The proof by 1. 7 also fails, under the Riemann hypothesis, if C, L are the poles of the straight line AB, since the broken lines CGL, CKL etc. become equal straight lines, all perpendicular to AB.] Proclus rightly adds (p. z8q, 18 sqq.) that it is not mcessary to take D on the side of AB away from A if an objector " says that there is no space on 1. is, 13] PROPOSITIONS is, 13 »75 that side." If it is not desired to trespass on that side of AB, we can take D anywhere on AB and describe the arc of a circle between D and the point where it meets AB again, drawing the arc on the side of AB on which C is. If it should happen that the selected point D is such that the circle only meets AB in oik point (D itself), we have only to describe the circle with CD as radius, then, if E be a point on this circle, take Fa point further from C than E is, and describe with CF as radius the circular arc meeting AB in two points. Proposition 13. If a straight line set up on a straight line make angles, it wilt make either two right angles or angles equal to two right angles. For let any straight line AB set up on the straight line s CD make the angles CBA, ABD j I say that the angles CBA, ABD are either two right angles or equal to two right angles. Now, if the angle CBA is equal to 10 the angle ABD, they are two right angles. [Def. 10] But, if not, let BE be drawn from the point B at right angles to CD ; [1. n] therefore the angles CBE, EBD are two right angles. is Then, since the angle CBE is equal to the two angles CBA, ABE, let the angle EBD be added to each ; therefore the angles CBE, EBD are equal to the three angles CBA, ABE, EBD. [C X a] M Again, since the angle DBA is equal to the two angles DBE, EBA, let the angle ABC be added to each ; therefore the angles DBA. ABC are equal to the three angles DBE, EBA, ABC. [ax*] ij But the angles CBE, EBD were also proved equal to the same three angles ; and things which are equal to the same thing are also equal to one another ; [C. X 1] therefore the angles CBE, EBD are also equal to the 3a angles DBA, ABC. 376 BOOK I [i. 13, 14 But the angles CBE, EBD are two right angles ; therefore the angles DBA, ABC are also equal to two right angles. Therefore etc. Q. E, D, 17. let the angle EBD be added to each, literally "let the angle EBD be added (so as to be) common, " *a*fy rpoandeBu % irrh EBA. Similarly tow)) d^pjsrtfw is used of subtracting a straight line or angle from each of two others. "Let the common angle EBD be added is clearly an inaccurate translation, for the angle is not common before it 13 added, i.e. the mur^i is proleptie. "Let the common angle be tuitraclid" as a translation of rar^ d therefore the remaining angle CEA is equal to the remaining angle BED. [C. If. 3] Similarly it can be proved that the angles CEB, DEA are also equal. Therefore etc. Q. E, D. *5 [Porism. From this it is manifest that, if two straight lines cut one another, they will make the angles at the point of section equal to four right angles.] 1. the vertical angles. The difference between adjacent angles (ai ifeffit fut/ku) and vertical fugles (td Kara xopva -yuvLar Jj licifuw xXcupi. iinoTtltti. In order to keep the proper order in English we must use the passive of the verb in I. 19. Aristotle quotes the result of 1. to, using the exact wording, vk& fkp t)jv ^ttflfw ywrtay bwoTttt/ci {Mtteorologica tit. 5, 376 a 11). " In order to assist the student in remembering which of these two propositions [1. 18, 19] is demonstrated directly and which indirectly, it may be observed that the order is similar to that in I. 5 and 1. 6" (Todhunter). An alternative proof of L 18 given by Porpnyry (see Ptoclus, pp. 315, \\ — 316, 13} is interesting. It starts by supposing a length equal to AB cut off from the other end of AC; that is, CD and not AD is made equal to AB. Produce AB to E so that BE is equal to AD, and join EC. Then, since AB is equal to CD, and BE to AD, AE is equal to AC, aS* BOOK I [i. 18, 19 Therefore the angle A EC is equal to the angle ACE. Now the angle ABC is greater than the angle A EC, [t. 16] and therefore greater than the angle ACE. Hence, a fortiori, the angle ABC is greater than the angle ACB. Proposition 19. In any triangle the greater angle is subtended by the greater side. Let ABC be a triangle having the angle ABC greater than the angle BCA ; I say that the side AC is also greater than the side AB. For, if not, AC is either equal to AB or less. Now AC is not equal to AB ; for then the angle ABC would also have been equal to the angle A CB ; [u 5] but it is not ; therefore AC is not equal to AB. Neither is AC less than AB, for then the angle ABC would also have been less than the angle ACB; [t 18] but it is not ; therefore AC is not less than AB. And it was proved that it is not equal either. Therefore AC is greater than AB, Therefore etc. q. e. d. This proposition, like t. 6, can be proved by merely logical deduction from 1. 5 and 1. 18 taken together, as pointed out by De Morgan. The general form of the argument used by De Morgan is given in his Formal Logic (1847), p. 25, thus : "Hypothesis. Let there be any number of propositions or assertions — three for instance, X, Y and Z — of which it is the property that one or the other must be true, and one only. Let there be three other propositions P, Q and P of which it is also the property that one, and one only, must be true. Let it be a connexion of those assertions that : when X is true, P is true, when Kis true, Q is true, when Z is true, R is true. Constquenu : then it follows that, when P is true, X is true, when Q is true, Y is true, when JR is true, Z is true" [1 ■5] [<• !8] [' .6] [«■ , 9 ] i. 19] PROPOSITIONS 18, 19 185 To apply this to the case before us, let us denote the sides of the triangle ABC by a, 6, c, and the angles opposite to these sides by A, B, C respectively, and suppose that a is the base. Then we have the three propositions, when b is equal to c, B is equal to C, when b is greater than c, B is greater than C, 1 when 6 is less than c, B is less than C, f and it follows logically that, when B is equal to C, b is equal to c, when B is greater than C, b is greater than c, \ when i? is less than C, b is less than c. J Reductio ad absurdum by exhaustion. Here, says Proclus (p. 318, 16 — 33), Euclid proves the impossibility "by means of division" (« fkaipurtan). This means simply the separation of different hypotheses, each of which is inconsistent with the truth of the theorem to be proved, and which therefore must be successively shown to be impossible. If a straight line is not greater than a straight tine, it must be either equal to it or less ; thus in a reductio ad absurdum intended to prove such a theorem as 1. 19 it is necessary to dispose successively of two hypotheses inconsistent with the truth of the theorem. Alternative (direct) proof. Proclus gives a direct proof (pp. 319—321) which an-NairizI also has and attributes to Heron. It requires a lemma and is consequently open to the slight objection of separating a theorem from its converse. But the lemma and proof are worth giving. Lemma. If an angle of a triangle be bisected and the straight line bisecting it meet the base and divide it into unequal parts, the sides containing the angle will be unequal, and the greater will be that which meets the greater segment of the base, and the less that which meets the lest. Let AD, the bisector of the angle A of the triangle ABC, meet BC in D, making CD greater than BD. I say that AC is greater than AB. Produce AD to £ so that DE is equal to AD. And, since DC is greater than BD, cut off DF equal to BD. Join BFsmd produce it to G. Then, since the two sides AD, DB are equal to the two sides ED, DF, and the vertical angles at D are equal, AB is equal to EF, and the angle DEF to the angle BAD, i.e. to the angle DAG (by hypothesis). Therefore AG is equal to EG, and therefore greater than EF, or AB, Hence, a fortiori, AC is greater than AB. *86 BOOK I U. 19, 30 Proof of I. 19. Let ABC be a triangle in which the angle ABC is greater than the angle ACB. * Bisect BC at D, join AD, and produce it to B so that DE is equal to ^Z>. Join BE. Then the iwo sides BD t DE are equal to the two sides CD, DA, and the vertical angles at D are equal ; therefore BE is equal to AC, and the angle DBE to the angle at C. But the angle at C is less than the angle ABC ; therefore the angle DBE is less than the angle ABD. Hence, if BF bisect the angle ABE, BF meets AE between A and D. Therefore EF is greater than FA. It follows, by the lemma, that BE is greater than BA, that is, AC is greater than >4.#. Proposition 20. /« a«y triangle twu sides taken together in any manner are greater than the remaining one. For let ABC be a triangle ; I say that in the triangle ABC two sides taken together in any manner are greater than the remaining one, namely BA, AC greater than BC, AB, BC greater than A C, BC, CA greater than AB. For let BA be drawn through to the point D s let DA be made equal to CA, and let DC be joined. Then, since DA is equal to AC, the angle ADC is also equal to the angle ACD; [1. 5 ] therefore the angle BCD is greater than the angle ADC. [C. JV. 5] And, since DCB is a triangle having the angle BCD greater than the angle BDC, and the greater angle is subtended by the greater side, [l 19] therefore DB is greater than BC. i. 20] PROPOSITIONS 19, 20 287 ButZM is equal to AC; therefore BA, AC are greater than BC. Similarly we can prove that AB, BC are also greater than CA, and BC, CA than AB. Therefore etc. Q. E. D, It was the habit of the Epicureans, says Prod us (p. 322), to ridicule this theorem as being evident even to an ass and requiring no proof, and their allegation that the theorem was "known" (yywpijiov) even to an ass was based on the fact that, if fodder is placed at one angular point and the ass at another, he does not, in order to get to his food, traverse the two sides of the triangle but only the one side separating them (an argument which makes Savile exclaim that its authors were "digni ipsi, qui cum Asino foenum essent," p. 78). Proclus replies truly that a mere perception of the truth of the theorem is a different thing from a scientific proof of it and a knowledge of the reason why it is true. Moreover, as Sim son says, the number of axioms should not be increased without necessity. Alternative Proofs. Heron and Porphyry, we are told (Proclus, pp. 323 — 6), proved this theorem in different ways as follows, without producing one of the sides. First proof. Let ABC be the triangle, and let it be required to prove that the sides BA, AC are greater than BC. Bisect the angle BAC by AD meeting BC inD. Then, in the triangle ABD, the exterior angle ADC is greater than the interior and opposite angle BAD, [1. 16] that is, greater than the angle DAC. Therefore the side AC is greater than the side CD, [1. 19] Similarly we can prove that AB is greater than BD. Hence, by addition, BA, AC are greater than BC. Second proof. This, like the first proof, is direct. There are several cases to be considered. (1) If the triangle is equilateral, the truth of the proposition is obvious. (3) If the triangle is isosceles, the proposition needs no proof in the case (a) where each of the equal sides is greater than the base. (#) If the base is greater than either of the other sides, we have to prove that the sum of the two equal sides is greater than the base. Let BC be the base in such a triangle. Cut off from BC a length BD equal to AB, and join AD. Then, in the triangle ADB, the exterior angle ADC is greater than the interior and opposite angle BAD. [1. 1 61 Similarly, in the triangle ADC, the exterior angle ADB is greater than the interior and opposite angle CAD. *88 BOOK I [i. to By addition, the tyro angles BDA, ADC are together greater than the two angles BA D, DA C (or the whole angle BA C). Subtracting the equal angles BDA, BAD, we have the angle ADC greater than the angle CAD. It follows that AC is greater than CD; [i. 19] and, adding the equals AB, BD respectively, we have BA, AC together greater than BC. (3) If the triangle be scalene, we can arrange the sides in order of length. Suppose BC is the greatest, AB the intermediate and AC the least side. Then it is obvious that AB, BC are together greater than AC, and BC, CA together greater than AB. It only remains therefore to prove that CA, AB are together greater than BC. We cut off from BC a length BD equal to the adjacent side, join AD, and proceed exactly as in the above case of the isosceles triangle. Thirdprwf. This proof is by reduetio ad ahsurdum. Suppose that BC is the greatest side and, as before, we have to prove that BA, AC are greater than BC. If they are not, they must be either equal to A or less than BC. (1) Suppose BA, AC ire together equal to BC. From BC cut off BD equal to BA, and join AD. It follows from the hypothesis that DC is equal to AC. Then, since BA is equal to BD, the angle BDA is equal to the angle BAD. Similarly, since AC is equal to CD, the angle CDA is equal to the angle CAD. By addition, the angles BDA, ADC are together equal to the whole angle BAC. That is, the angle BAC is equal to two right angles : which is impossible. {2) Suppose BA, AC ate together less than BC. From BC cut off BD equal to BA, and from CB cut off CE equal to CA. Join AD, AE. In this case, we prove in the same way that the angle BDA is equal to the angle BAD, and the angle CEA to the angle CAE. By addition, the sum of the angles BDA, AEC is equal to the sum of the angles BAD, CAE. Now, by 1. 16, the angle BDA is greater than the angle DAC, and therefore, a fortiori, greater than the angle EA C. Similarly the angle AEC is greater than the angle BAD. Hence the sum of the angles BDA, AEC is greater than the sum of the angles BAD, EAC. But the former sum was also equal to the latter : which is impossible, li ai] PROPOSITIONS 20, 21 289 Proposition 21. If on one of the sides of a triangle, from its extremities, there be constructed two straight lines meeting within the triangle, the straight lines so constructed will be less than the remaining two sides of the triangle, but will contain a greater s angle. On BC, one of the sides of the triangle ABC, from its extremities B, C, let the two straight lines BD, DC be con- structed meeting within the triangle ; I say that BD, DC are less than the remaining two sides 10 of the triangle BA, AC, but contain an angle BDC greater than the angle BAC. For let BD be drawn through to E. Then, since in any triangle two sides are greater than the remaining 15 one, [1. zo] therefore, in the triangle ABE, the two sides A B, AE are greater than BE. Let EC be added to each ; therefore BA, AC art greater than BE, EC. 20 Again, since, in the triangle CED, the two sides CE, ED are greater than CD, let DB be added to each ; therefore CE, EB are greater than CD, DB. But BA, AC were proved greater than BE, EC; 2$ therefore BA, AC are much greater than BD, DC. Again, since in any triangle the exterior angle is greater than the interior and opposite angle, [1- 16] therefore, in the triangle CDE, the exterior angle BDC is greater than the angle CED. 30 For the same reason, moreover, in the triangle ABE also, the exterior angle CEB is greater than the angle BAC. But the angle BDC was proved greater than the angle CEB ; therefore the angle BDC is much greater than the angle BAC. 3 5 Therefore etc. q. e. d, 1* be con strutted. ..meeting within the triangle. The word n meeting" is not in the Greek, where the words are irrht rwmrt&rty. evrLrrwiBtu is the word used of con- structing two straight lines to a point (cf- [■ 7) or so as to form a triangle ; but it is necessary in English to indicate that they mat. 3. the straight lines so constructed. Observe the elegant brevity of the Creek al 29° BOOK I [i. at The editors generally call attention to the fact that the lines drawn within the triangle in this proposition must be drawn, as the enunciation says, from the ends of the side ; otherwise it is not necessary that their sum should be less than that of the remaining sides of the triangle. Proclus (p. 3*7, II sqq.) gives a simple illustration. Let ABC be a right-angled triangle. Take any point D on BC, join DA, and cut off from it DE equal to AB. Bisect AE at F, and join FC. Then shall CF, FD be together greater than CA, AB. For CF, FE are equal to CF, FA, and therefore greater than CA. Add the equals ED, AB respectively ; therefore CF, FD are together greater than CA, AB. Pappus gives the same proposition as that just proved, but follows it up by a number of others more elaborate in character, selected apparently from " the so-called paradoxes " of one Erycinus (Pappus, m. p. 106 sqq.). Thus he proves the following : 1. In any triangle, except an equilateral triangle or an isosceles triangle with base less than one of the other sides, it is possible to construct on the base and within the triangle two straight lines the sum of which is equal to the sum of the other two sides of the triangle. 2. In any triangle in which it is possible to construct two straight lines on the base which are equal to the sum of the other two sides of the triangle it is also possible to construct two others the sum of which is greater than that sum. 3. Under the same conditions, if the base is greater than either of the other two sides, two straight tines can be constructed in the manner described which are respectively greater than the other two sides of the triangle ; and the lines may be constructed so as to be respectively equal to the two sides, if one of those two sides is less than the other and each of them less than the base. 4. The lines may be so constructed that their sum will bear to the sum of the two sides of the triangle any ratio less than 2:1. As a specimen of the proofs we will give that of the proposition which has been numbered (1) for the case where the triangle is isosceles (Pappus, in. pp. 108 — 110)1 I. »i] PROPOSITION 21 391 Let ABC be an isosceles triangle in which the base AC is greater than either of the equal sides .<4.5, BC. With centre v4 and radius AB describe a circle meeting j4Cin D. Draw any radius AEFsuch that it meets BC in a point F outside the circle. Take any point G on EF, and through it draw GZf parallel to AC. Take any point .AT on GH, and draw KL parallel to FA meeting AC in L. From jSCcut off BN equal to EG. Thus AG, or LK, is equal to the sum of AB, BN, and CWis less than LK. Now GF, Fffare together greater than GH, and CH, UK together greater than CK, Therefore, by addition, CF, FG, HK are together greater than CK, HG. Subtracting HK from each side, we see that CF, FG are together greater than CK, KG ; therefore, if we add AG to each, AF, FCaie together greater than AG, GK, KC. And AB, BC are together greater than AF, EC. [1. 31] Therefore AB, BC are together greater than A G, GK, KC. But, by construction, AB, BN are together equal to AG ; therefore, by subtraction, NC is greater than GK, KC, and a fortiori greater than KC. Take on KC produced a point .A/' such that KM is equal to NC; with centre K and radius KM describe a circle meeting CL in 0, and join KO. Then shall LK, KO be equal to AB, BC. For, by construction, LK is equal to the sum of AB, BN, and KO is equal to NC; therefore LK, KO are together equal to AB, BC. It is after 1. at that (as remarked by De Morgan) the important proposition about the perpendicular and obliques drawn from a point to a straight line of unlimited length is best introduced : Of all straight lines that can be drawn to a given straight line of unlimited length from a given point without it : (a) the perpendicular is the shortest ; (b) of the obliques, that is the greater the fool of which is further from the perpendicular ; (e) given one oblique, only one other can be found of the same length, namely that the foot of which is equally distant with the foot of the given one from the perpendicular, but on the other side of it. Let A be the given point, BC the given straight line ; let AD be the perpendicular from A on BC, and AE, AF any two obliques of which AF makes the greater angle with AD. Produce AD to A', making A'D equal to AD, and join A'E, A F. Then the triangles ADE, A'DE are equal in all respects ; and so are the triangles ADF, A'DF. Now (1) in the triangle AEA' the two sides AE, EA' are-greater than AA' [1. 20I, that is, twice AE is greater than twice AD. a 9 * BOOK I [l. a i, Therefore AE is greater than AD. (a) Since AE, A'E are drawn to E, a point within the triangle A FA', AE, EA' are together greater than AE, EA\ [l *i] or twice AE is greater than twice AE.. Therefore AE is greater than AE. (3) Along DB measure off DG equal to DF, and join AG. The triangles AGD, AFD are then equal in all respects, so that the angles GAD, FAD are equal, and AG is equal to AF. Proposition 22. Out of three straight lines, which are equal to three given straight lines, to construct a triangle : thus it is necessary that two of the straight lines taken together in any manner should be greater than the remaining one. [1. *o] Let the three given straight lines be A, B, C, and of these let two taken together in any manner be greater than the remaining one, namely A, B greater than C, A, C greater than B, and B, C greater than A ; thus it is required to construct a triangle out of straight lines equal to A, B, C. a- 8- c- Let there be set out a straight line DB, terminated at D but of infinite length in the direction of B, and let DF be made equal to A, FG equal to B, and GH equal to C. [1. 3] With centre F and distance FD let the circle DKL be described ; again, with centre G and distance GH let the circle KLH be described ; and let KF, KG be joined ; I say that the triangle KFG has been constructed out of three straight lines, equal to A, B, C. i. 11] PROPOSITIONS a i, a* »93 For, since the point F is the centre of the circle DKS,, FD is equal to FK. But FD is equal to A ; therefore KF is also equal to A. Again, since the point G is the centre of the circle LKH, GH is equal to GK. But GH is equal to C; therefore KG is also equal to C. And FG is also equal to B ; therefore the three straight lines KF, FG, GK are equal to the three straight lines A, B, C. Therefore out of the three straight lines KF, FG, GK, which are equal to the three given straight lines A, B, C, the triangle KFG has been constructed. 6 q. e. F. i — *> This is the first cast in the Elements of a Satparfiii to a problem in the sense of a statement of the conditions or limits of the possibility of a solution. The criterion is of course supplied by the preceding proposition. j. thu» It is necessary. This is usually translated (e.g. by Williamson and Simson) "But it is necessary," which is however inaccurate, since the Greek is not Sti W but &ri ti). The words are the same as those used to introduce the ttopurpit in the other sense of the " definition " or " particular statement " of a construction to be effected. Hence, as in the latter case we say " thus it is required " (e.g. to bisect the finite straight line AS, I. 10}, we should here translate " thus it is necessary. 4* To this enunciation alt the M5S. and Bocthius add, after the Sioptefifa, the words "because in any triangle two sides taken together in any manner are greater than the remaining one." But this explanation has the appearance of a gloss, and it is omitted hy Proclus and Campanus. Moreover there is no corresponding addition to the Supwitit of vi, 18. It was early observed that Euclid assumes, without giving any reason, that the circles drawn as described will meet if the condition that any two of the straight lines A, S, C are together greater than the third be fulfilled. Prod us (p. 33 r, S sqq.) argues the matter by means of redudio ad abrurdum, but does not exhaust the possible hypotheses inconsistent with the contention. He says the circles must do one of three things, (1) cut one another, (1) touch one another, (3} stand apart {&«rrarai) from one another. He then considers the hypotheses (a) of their touching externally, (t>) of their being separated from one another by a space. He should have considered also the hypothesis (r) of one circle touching the other internally or lying entirely within the other without touching. These three hypotheses being successively disproved, it follows that the circles must meet (this is the line taken by Camerer and Todhunter). Simson says : " Some authors blame Euclid because he does not demonstrate that the two circles made use of in the construction of this problem must cut one another : but this is very plain from the determination he has given, namely, that any two of the straight lines DF, FG, GH must be greater than the third, For who is so dull, though only beginning to learn the Elements, as not to perceive that the circle described from the centre F, at the distance FD, must meet FH betwixt F and H, because FD is less than FH; and that, for the like reason, the circle described from the ig4 BOOK I [i. 22, 33 centre G at the distance GH must meet DG betwixt D and G ; and that these circles must meet one another, because FD and GH are together greater than FG." We have in fact only to satisfy ourselves that one of the circles, e.g. that with centre G, has at least one point of its circumference outside the other circle and also at least one point of its circumference inside the same circle ; and this is best shown with reference to the points in which the first circle cuts the straight line DE. For (i) FH, being equal to the sum of B and C, is greater than A t i.e. than the radius of the circle with centre F, and therefore His outside that circle. (2) If GAf be measured along GF equal to GH or C, then, since GM is either (a) less or {p} greater than GF, Jfwill fall either {a) between G and F or (6) beyond F towards D ; in the first case (a) the sum of FM and C is equal to FG and therefore less than the sum of A and C, so that FM is less than A or FD ; in the second case {/>) the sum of MF and FG, i.e. the sum of MF and B, is equal to GAf or C, and therefore less than the sum of ^ and B, so that MF is less than A or FD ; hence in either case M falls within the circle with centre F. It being now proved that the circumference of the circle with centre G has at least one point outside, and at least one point inside, the circle with centre F, we have only to invoke the Principle of Continuity, as we have to do in 1. 1 (cf. the note on that proposition, p. 242, where the necessary postulate is stated in the form suggested by Killing). That the construction of the proposition gives only two points of intersection between the circles, and therefore only two triangles satisfying the condition, one on each side of FG, is clear from I. 7, which, as before pointed out, takes the place, in Book 1., of 111. 10 proving that two circles cannot intersect in more points than two. Proposition 23. On a given straight line and at a point on it to construct a rectilineal angle equal to a given rectilineal angle. Let AB be the "given straight line, A the point on it, and the angle DCE the given rectilineal angle; thus it is required to construct on the given straight line AB, and at the point A on it, a rectilineal angle equal to the given rectilineal angle DCE. On the straight lines CD, CE respectively let the points D, E be taken at random ; let DE be joined, and out of three straight lines which are equal to the three li »3] PROPOSITIONS 22, 23 395 straight lines CD, DE, CE let the triangle AFG he. con- structed in such a way that CD is equal to AF, CE to AG, and further DE to FG. [1. 22] Then, since the two sides DC, CE are equal to the two sides FA., AG respectively, and the base DE is equal to the base FG, the angle DCE is equal to the angle FAG. [1. 8] Therefore on the given straight line A3, and at the point A on it, the rectilinealangle FAG has been constructed equal to the given rectilineal angle DCE. e Q. E. F. This problem was, according to Eudemus (see Proclus, p. 333, 5), "rather the discovery of Oenopides," from which we must apparently infer, not that Oenopides was the first to find any solution of it, but that it was he who dis- covered the particular solution given by Euclid. (Cf. Bretschneider, p. 65.) The editors do not seem to have noticed the fact that the construction of the triangle assumed in this proposition is not exactly the construction given in 1. 22. We have here to construct a triangle on a certain finite straight line AG as base; in 1. 11 we have only to construct a triangle with sides of given length without any restriction as to how it is to be placed. Thus in 1. ■• we set out any tine whatever and measure successively three lengths along it beginning from the given extremity, and what we must regard as the base is the intermediate length, not the length beginning at the given extremity, of the straight line arbitrarily set out. Here the base is a given straight line abutting at a given point Thus the construction has to be modified somewhat from h B that of the preceding proposition. We must measure AG along AB so that AG is equal to CE (or CD), and GH along GB equal to DE; and then we must produce BA, in the opposite direction, to F, so that AF'is equal to CD (or CE, if AG has been made equal to CD). Then, by drawing circles (1) with centre A and radius AF, (2) with centre G and radius GH, we determine K, one of their points of intersection, and we prove that the triangle KAG is equal in all respects to the triangle DCE, and then that the angle at A is equal to the angle DCE. I think that Proclus must (though he does not say so) have felt the same difficulty with regard to the use in 1. 33 of the result of 1. 22, and that this is probably the reason why he gives over again the construction which I have given above, with the remark (p. 334, 6) that "you may obtain the construction of the triangle in a more instructive manner (StSatricaA.iKioTi^oi') as follows," Proclus objects to the procedure of Apollonius in constructing an angle under the same conditions, and certainly, if he quotes Apollonius correctly, the tatter's exposition must have been somewhat slipshod. 296 BOOK I {]. *3, «4 "He takes an angle CDE at random," says Prod us (p. 335> '9 s Vl-)> ' ,and a straight line AB, and with centre D and distance CD describes the circumference CE, and in the same way with centre A and distance AB the circumference FB. Then, cutting off FB equal to CE, he joins AF. And he declares that the angles A, D standing on equal circumferences are equal." In the first place, as Prod us remarks, it should be premised that AB is equal to CD in order that the circles may be equal; and the use of Book lit. for such an elementary construction is objectionable. The omission to state that AB must be taken equal to CD was no doubt a slip, if it occurred. And, as regards the equal angles "standing on equal circum- ferences," it would seem possible that Apollonius said this in explanation, for the sake of brevity, rather than by way of proof. It seems to me probable that his construction was only given from the point of view of practical, not theoretical, geometry. It really comes to the same thing as Euclid's except that DC is taken equal to DE. For cutting off the arc BF equal to the arc CE can only be meant in the sense of measuring the chord CE, say, with a pair of compasses, and then drawing a circle with centre B and radius equal to the chord CE. Apollonius' direction was therefore probably intended as a practical short cut, avoiding the actual drawing of the chords CE, BF, which, as well as a proof of the equality in all respects of the triangles CDE, BAF, would be required to establish theoretically the correct- ness of the construction. Proposition 24. If two triangles have the two sides equal to two sides respectively, but have the one of the angles contained by the equal straight lines greater than the other, they will also have the base greater than the base. 5 Let ABC, DEF be two triangles having the two sides AB, A C equal to the two sides DE, DF respectively, namely AB to DE, and A C to DF, and let the angle at A be greater than the angle at D ; I say that the base BC is also greater than the base £F, 10 For, since the angle BAC is greater than the angle EDF, let there be constructed, on the straight line DE, and at the point D on it, the angle EDG jj equal to the angle BAC; [1. 13] let DG be made equal to either of the two straight lines AC, DF, and let EG, FG be joined. i. 24] PROPOSITIONS 23, 24 *97 Then, since AB is equal to DE, and AC to DG, *>the two sides BA, AC are equal to the two sides ED, DG, respectively ; and the angle BAC is equal to the angle EDG ; therefore the base BC is equal to the base EG. [1. 4] Again, since DF is equal to DG, 25 the angle DGF is also equal to the angle DFG ; [l 5] therefore the angle DFG is greater than the angle EGF. Therefore the angle EFG is much greater than the angle EGF. And, since EFG is a triangle having the angle EFG 30 greater than the angle EGF, and the greater angle is subtended by the greater side, [*. 19] the side EG is also greater than EF. But EG is equal to BC. Therefore BC is also greater than EF. 3S Therefore etc. Q. E. D. 10. I have naturally left out the well-known words added by Simson in order to avoid the necessity of considering three cases : " Of the two sides DE, DF let DE be the side which is not greater than the other." I doubt whether Euclid could have been induced to insert the words himself, even if it had been represented to him that their omission meant leaving two possible cases out of consideration. His habit and that of the great Greek geometers was, not to set out all possible cases, but to give as a rule one case, generally the most difficult, as here, and to leave the others to the reader to work out for himself. We have already seen one instance in 1. 7. Proclus of course gives the other two cases which arise if we do not first provide that DE is not greater than DF. (1) In the first case G may fall on EF produced, and it is then obvious that EG is greater than EF. (2) In the second case EG may fall below EF. If so, by 1. 21, DF, FE are together less than DG, GE. But DF is equal to DG ; there- fore EF is less than EG, i.e. than BC. These two cases are therefore decidedly simpler than the case taken by Euclid as typical, and could well be left to the ingenuity of the learner. If however after all we prefer to insert Simson's words and avoid the latter BOOK I [i-*4 two cases, the proof is not complete unless we show that, with his assumption, /"must, in the figure of the proposition, fall below EG. De Morgan would make the following proposition precede: Every straight tine drawn from the vertex of a triangle to the base is less than the greater of the two sides, or than either if they are equal, and he would prove it by means of the proposition relating to perpendicular and obliques given above, p. 291. But it is easy to prove directly that F falls below EG, if DE is not greater than DG, by the method employed by Pfleiderer, Lardner, and Todhunter. Let DF, produced if necessary, meet EG in H. Then the angle DUG is greater than the angle DEG\ ["• >6] and the angle DEG is not less than the angle DGE ; [.. .8] therefore the angle DUG is greater than the angle DGH. Hence DH is less than DG, [1. 19] and therefore DH is less than DF. Alternative proof. Lastly, the modern alternative proof is worth giving. A Let DHhisect the angle FDG (after the triangle DEG has been made equal in all respects to the triangle A BC t as in the proposition), and let DH meet EG in H. Join HF. Then, in the triangles FDH, GDH, the two sides FD, DH are equal to the two sides GD, DH, and the included angles FDH, GDH at equal ; therefore the base HF is equal to the base HG Accordingly EG is equal to the sum of EH, HF; and EH, HF are together greater than EF; [1. 20] therefore EG, or BC, is greater than EF. Proclus (p. 339, 1 1 sqq.) answers by anticipation the possible question that might occur to any one on this proposition, viz. why does Euclid not compare the areas of the triangles as he does in 1. 4 ? He observes that inequality of the areas does not follow from the inequality of the angles contained by the equal sides, and that Euclid leaves out all reference to the question both for this reason and because the areas cannot be compared without the help of the theory of parallels. " But if," says Proclus, " we must anticipate what is to come and make our comparison of the areas at once, we assert that (1) if the angles A, D — supposing that our argument proceeds with reference to the figure in the proposition — are {together) equal to two right angles, the triangles L 34, 25] PROPOSITIONS 14, 25 299 a« primed equal, {2) if greater than tins right angles, that triangle which has the greater angle is less, and (3) 4] but it is not ; therefore the angle BA C is not equal to the angle EDF. Neither again is the angle BA C less than the angle EDF; for then the base BC would also have been less than the base EF, [ft 24] but it is not ; therefore the angle BA C is not less than the angle EDF. But it was proved that it is not equal either ; therefore the angle BAC is greater than the angle EDF. Therefore etc. Q. E. D. 3 oo BOOK I [i. 15 De Morgan points out that this proposition (as also i. 8) is a purely logical consequence of i. 4 and 1. 14 in the same way as 1. r 9 and 1. 6 are purely logical consequences of 1. 18 and 1. 5. If d, 6, e denote the sides, A, B, C the angles opposite to them in a triangle ABC, and a', b', /, A', E, C the sides and opposite angles respectively in a triangle A'HC, 1. 4 and I. 14 tell us that, i, e being respectively equal to i\ /, ( 1 ) if A is equal to A\ then a is equal to a', (z) if A is less than A', then a is less than «', (3) if A is greater than A', then a is greater than a' ; and it follows logically that, (1) if a is equal to a, the angle A is equal to the angle A', [1. 8] (3) if a is less than a, A is less than A', \ (3) if a is greater than d, A is greater than A'. } I * a 5J Two alternative proofs of this theorem are given by Proclus (pp. 345 — 7), and they are both interesting. Moreover both are direct. I. Proof by Menelaus of Alexandria. Let ABC, DEB" be two triangles having the two sides BA, AC equal to the two sides ED, DF, but the base BC greater than the base EF. Then shall the angle at A be greater than the angle at D. From BC cut off BG equal to EF. At B, on the straight line BC, make the angle GBH (on the side of BG remote from A) equal to the angle FED. Make BH equal to DE ; join HG, and produce it to meet A C in K. Join AH. Then, since the two sides GB, BH are equal to the two sides FE, ED respectively, and the angles contained by them are equal, HG is equal to ZVor AC, and the angle BHG is equal to the angle EDF. Now HK is greater than HG or AC, and a fortiori greater than AK; therefore the angle KAH\% greater than the angle KHA. And, since AB is equal to BH, the angle BAH is equal to the angle BHA. Therefore, by addition, the whole angle BA C is greater than the whole angle BHG, that is, greater than the angle EDF. i. «5, a6] PROPOSITIONS 35, 36 301 II, Heron's proof. Let the triangles be given as before. Since BC is greater than EF, produce EF to G so that EG is equal to EC. Produce ED to If so that DH is equal to DF. The circle with centre D and radius DFwiW then pass through H. Let it be described, as FKH. Now, since BA, AC are together greater than £C, and Atf, ^C are equal to ,££>, Z>Zf respectively, while JC is equal to EG, EH is greater than EG. Therefore the circle with centre E and radius EG will cut Elf, and therefore will cut the circle already drawn. Let it cut that circle in K, and join DK, KB. Then, since D is the centre of the circle FKH, DK is equal to Z>^or AC Similarly, since £ is the centre of the circle KG, EK is equal to EG or BC, And DE is equal to A B. Therefore the two sides BA, AC are equal to the two sides ED, DK respectively; and the base BC is equal to the base EK; therefore the angle BAC is equal to the angle EDK Therefore the angle BA C is greater than the angle EDF. Proposition 26, If two triangles have the two angles equal to two angles respectively, and one side equal to one side, namely, either the side adjoining the equal angles, or that subtending erne 0/ the equal angles, they wilt also have the remaining sides equal to s the remaining sides and the remaining angle to the remaining angle. 3°* BOOK I [1. 26 Let ABC, DEF be two triangles having the two angles ABC, BCA equal to the two angles DEF, EFD respectively, namely the angle ABC to the angle DEF, and the angle 10 BCA to the angle EFD ; and let them also have one side equal to one side, first that adjoining the equal angles, namely BC to EF; I say that they will also have the remaining sides equal to the remaining sides respectively, namely AB to DE and >s AC to DF, and the remaining angle to the remaining angle, namely the angle BA C to the angle EDF. For, if AB is unequal to DE, one of them Is greater. Let AB be greater, and let BG be made equal to DE ; and let GC be joined. 20 Then, since BG is equal to DE, and BC to EF, the two sides GB, BC are equal to the two sides DE, EF respectively; and the angle GBC is equal to the angle DEF ; therefore the base GC is equal to the base DF, 25 and the triangle GBC is equal to the triangle DEF, and the remaining angles will be equal to the remaining angles, namely those which the equal sides subtend ; [1. 4] therefore the angle GCB is equal to the angle DFE. But the angle DFE is by hypothesis equal to the angle BCA; 3 o therefore the angle BCG is equal to the angle BCA, the less to the greater : which is impossible. Therefore AB is not unequal to DE, and is therefore equal to it. But BC is also equal to EF; 35 therefore the two sides AB, BC are equal to the two sides DE, EF respectively, and the angle ABC is equal to the angle DEF; therefore the base A C is equal to the base DF, and the remaining angle BAC is equal to the remaining 40 angle EDF. [l 4] I. *6] PROPOSITION a* 303 Again, let sides subtending equal angles be equal, as AB to DE; I say again that the remaining sides will be equal to the remaining sides, namely AC to DF and BC to EF, and 45 further the remaining angle BAC is equal to the remaining angle EDF. For, if BC is unequal to EF, one of them is greater. Let BC be greater, if possible, and let BH be made equal to EF; let AH be joined. 50 Then, since BH is equal to EF, and AB to DE, the two sides AB, BH are equal to the two sides DE, EF respectively, and they contain equal angles ; therefore the base AH is equal to the base DF, and the triangle ABH is equal to the triangle DEF, ss and the remaining angles will be equal to the remaining angles, namely those which the equal sides subtend ; [1. 4] therefore the angle BHA is equal to the angle EFD. But the angle EFD is equal to the angle BCA ; therefore, in the triangle AHC, the exterior angle BHA is 60 equal to the interior and opposite angle BCA : which is impossible. [1. 16] Therefore BC is not unequal to EF, and is therefore equal to it. But AB is also equal to DE ; 65 therefore the two sides AB, BC are equal to the two sides DE, EF respectively, and they contain equal angles ; therefore the base AC is equal to the base DF, the triangle ABC equal to the triangle DEF, and the remaining angle BAC equal to the remaining angle r>EDF, [1.4] Therefore etc. Q, E. D. 1 — 3. the aide adjoining the equal angles, rhtupir Hjr rpot roTi but ywrtati. 39. la by hypothesis equal. vr6Ktir at tnj, according to the elegant Greek idiom. ifbtettiat is used for the passive of irrvrlffitfn, as Keit*eu is used for the passive of riBrtfu, and to with the other compounds. Cf. Trpxruiitidat, " to be added." The alternative method of proving this proposition, viz. by applying one triangle to the other, was very early discovered, at least so far as regards the case where the equal sides are adjacent to the equal angles in each. An-Nairizi gives it for this case, observing that the proof is one which he had found, but of which he did not know the author. 3o 4 BOOK I [i 3 6 Proclus has the following interesting note {p. 35*, 13 — 18): "Eudemus in his geometrical history refers this theorem to Thales. For he says that, in the method by which they say that Thales proved the distance of ships in the sea, it was necessary to make use this theorem." As, unfortunately, this information is not sufficient of itself to enable us to determine how Thales solved this problem, there is considerable room for conjecture as to bis method. The suggestions of Bretschneider and Cantor agree in the assumption that the necessary observations were probably made from the top of some tower or structure of known height, and that a right-angled triangle was used in which the tower was the perpendicular, and the line connecting the bottom of the tower and the ship was the base, as in the annexed figure, where AB is the tower and C the ship. Bretschneider (Die Geometrie and die Geometer vor Eukleides, § 30) says that it was only necessary for the observer to observe the angle CAB, and then the triangle would be completely determined by means of this angle and the known length AB. As Bretschneider says that the result would be obtained "in a moment " by this method, it is not clear in what sense he supposes Thales to have "observed" the angle SAC. Cantor is more definite (GescA. d. Math. i ( , p. 145), for he says that the problem was nearly related to that of finding the Seqt from given sides. By the Seqt in the Papyrus Rhind is meant the ratio to one another of certain lines in pyramids ot obelisks. Eisenlohr and Cantor took the one word to be equivalent, sometimes to the cosine of the angle made by the edge of the pyramid with the coterminous diagonal of the base, sometimes to the tangent of the angle of slope of the faces of the pyramid. Ft is now certain that it meant one thing, viz. the ratio of half the side of the base to the height of the pyramid, i.e. the cotangent of the angle of slope. The calculation of the Seqt thus implying a sort of theory of simi- larity, or even of trigonometry, the suggestion of Cantor is apparently that the Seqt in this case would be found from a smalt right-angled triangle ADE having a common angle A with ABC as shown in the figure, and that the ascertained value of the Stqt with the length AB would determine BC. This amounts to the use of the property of similar triangles ; and Bretschneider's suggestion must apparently come to the same thing, since, even if Thales measured the angle in our sense (e.g. by its ratio to a right angle), he would, in the absence of something corresponding to a table of trigonometrical ratios, have gained nothing and would have had to work out the proportions all the same. Max C P, Schmidt also (Kulturhistorische Beiirage zur Kenntnis des griechisehcn und romischen Alter turns, 1906, p. 32) similarly supposes Thales to have had a right angle made of wood or bronze with the legs graduated, to have placed it in the position ADE (A being the position of his eye}, and then to have read off the lengths AD, DE respectively, and worked out the length of BC by the rule of three. How then does the supposed use of similar triangles and their property square with Eudemus' remark about 1. 26 ? As it stands, it asserts the equality of two triangles which have two angles and one side respectively equal, and the theorem can only be brought into relation with the above explanations by taking it as asserting that, if two angles and one side of one triangle are given, the triangle is completely determined. But, if Thales i. 26] PROPOSITION 26 305 practically used proportions, as supposed, 1. 26 is surely not at all the theorem which this procedure would naturally suggest as underlying it and being "necessarily used"; the use of proportions or of similar but not equal triangles would surely have taken attention altogether away from 1. 26 and fixed it on vi. 4. For this reason I think Tannery is on the right road when he tries to find a solution using 1. 26 as it stands, and withal as primitive as any recorded solution of such a problem. His suggestion (La Gtemitrit gretqut, pp. 90—1) is based on the fiuminis varatio of the Roman agrimensor Marcus Junius Nipsus and is as follows. To find the distance from a point A to an inaccessible point B. From A measure along a straight line at right angles to AB a length AC and bisect it at D. From C draw C£ at right angles to CA on the side of it remote from B, and let £ be the point on it which is in a straight line with B and D. Then, by 1. 26, CE is obviously equal to AB. As regards the equality of angles, it is to be observed that those at D are equal because they are vertically opposite, and, curiously enough, Thales is expressly credited with the discovery of the equality of such angles. The only objection which I can see to Tannery's solution is that it would require, in the case of the ship, a certain extent of free and level ground for the construction and measurements. I suggest therefore that the following may have been Thales' method. Assuming that he was on the top of a tower, he had only to use a rough instrument made of a straight stick and a cross-piece fastened to it so as to be capable of turning about the fastening (say a nail) so that it could form any angle with the stick and would remain where it was put. Then the natural thing would be to fix the stick upright (by means of a plumb-line) and direct the cross-piece towards the ship. Next, leaving the cross-piece at the angle so found, the stick could be turned round, still remaining vertical, until the cross-piece pointed to some visible object on the shore, when the object could be mentally noted and the distance from the bottom of the tower to it could be subsequently measured. This would, by I. 36, give the distance from the bottom of the tower to the ship. This solution has the advantage of corresponding better to the simpler and more probable version of Thales' method of measuring the height of the pyramids; Diogenes Laertius says namely (i. 27, p. 6, ed. Cobet) on the authority of Hieronymus of Rhodes (b.c 293 — 230), that he waited for this purpose until the moment when our shadows are of thesamt length as ourselves. Recapitulation of congruence theorems. Proclus, like other commentators, gives at this point (p. 347, 20 sqq.) a summary of the cases in which the equality of two triangles in all respects can be established. We may, he says, seek the conditions of such equality by successively considering as hypotheses the equality (1) of sides alone, (2) of angles alone, (3) of sides and angles combined. Taking (1) first, we can only establish the equality of the triangles in all respects if all three sides are respectively equal; we cannot establish the equality of the triangles by any hypothesis of class (2), not even the hypothesis that all the three angles are respectively equal ; among the hypotheses of class (3), the equality of one jo6 BOOK I [i. 16 side and one angle in each triangle is not enough, the equality (a) of one side and all three angles is more than enough, as is also the equality {&) of two sides and two or three angles, and (c) of three sides and one or two angles. The only hypotheses therefore to be examined from this point of view are the equality of (a) three sides [Eucl. 1. 8]. (jB) two sides and one angle [1. 4 proves one case of this, where the angle is that contained by the sides which are by hypothesis equal]. (y) one side and two angles [1. *6 covers all cases]. It is curious that Proclus makes no allusion to what we call the ambiguous case, that case namely of (£') in which it is an angle opposite to one of the two specified sides in one triangle which is equal to the angle opposite to the equal side in the other triangle. Camerer indeed attributes to Proclus the observation that in this case the equality of the triangles cannot, unless some other condition is added, be asserted generally ; but it would appear that Camerer was probably misled by a figure {Proclus, p. 351) which looks like a figure of the ambiguous case but is really only used to show that in 1. 26 the equal sides most be corresponding sides, i.e., they must be either adjacent to the equal angles in each triangle, or opposite to corresponding equal angles, and that, e.g., one of the equal sides must not be adjacent to the two angles in one triangle, while the side equal to it in the other triangle is opposite to one of the two corresponding angles in that triangle. The ambiguous case. If (wo triangles have two sides equal to two sides respectively, and if /he angles opposite to one pair of equal sides be also equal, then wilt the angles opposite the other pair of equal sides be either equal or supplementary ; and, in the former ease, the triangles will be equal in all respects. Let ABC, DEFbt two triangles such that AB is equal to DE, and AC to DF, while the angle ABC is equal to the angle DEF^ it is required to prove that the angles ACB, DFE are either equal or supplementary. A A ? Now (1), if the angle BAC be equal to the angle EDF, it follows, since the two sides AB, A Cute equal to the two sides DE, DF respectively, that the triangles ABC, DEFaxe equal in all respects, [l. 4] and the angles A CB, DFE are equal. (a) If the angles BAC, EDF be not equal, make the angle EDO {on the same side of ED as the angle EDF) equal to the angle BA C. Let EF, produced if necessary, meet DG in G. Then, in the triangles ABC, DEG, the two angles BAC, ABC are equal to the two angles EDG, DEG respectively, and the side AB is equal to the side DE; i. 16, 17] PROPOSITIONS 26, 37 307 therefore the triangles ABC, DEG ate equal in all respects, [1. a 6] so that the side AC is equal to the side DG, and the angle ACS is equal to the angle DGE. Again, since AC is equal to DF&s well as to DG, DF'vl equal to DG, and therefore the angles DFG, DGFaie equal. But the angle DFE is supplementary to the angle DFG; and the angle DGF was proved equal to the angle ACS; therefore the angle DFE is supplementary to the angle ACB. If it is desired to avoid the ambiguity and secure that the triangles may be congruent, we can introduce the necessary conditions into trie enunciation, on the analogy of Eucl. vi. 7. If two triangles have two sides of the one equal to two sides of the other respectively, and the angles opposite to a pair of equal sides equal, then, if tie angles opposite to the other pair of equal sides are both acute, or both obtuse, or if one of them is a right angle, the two triangles art equal in all respects. The proof of the three cases (by reductio ad absurdum) was given by Todhunter- Proposition 27. If a straight line failing on two straight lines make the alternate angles equal to one another, the straight lines will be parallel to one another. For let the straight line EF failing on the two straight s lines AB, CD make the alternate angles AEF, EFD equal to one another ; I say that AB is parallel to CD. For, if not, AB, CD when pro- duced will meet either in the direction 10 of B, D or towards A, C. Let them be produced and meet, in the direction of B, D, at G. Then, in the triangle GEF, the exterior angle A EF is equal to the interior and opposite is angle EFG : which is impossible. [1. 16' Therefore AB, CD when produced will not meet in the direction of B, D. Similarly it can be proved that neither will they meet 10 towards A, C. 3 o8 BOOK I [i. »7 But straight lines which do not meet in either direction are parallel ; [Def. 23] therefore AB is parallel to CD. . Therefore etc. Q. E. D. 1. filling on two straight lines, ds tfa tiSiUit t/irtrrvuita, the phrase being the s u that used in Post, 5, meaning a transversa!. 1. the alternate angles, a! rfrsXXif yteriai. Produs (p. 357, 9) explains that Euclid uses the word alternate (or, more exactly, alternattfy, kaXXdf) in two conneiions, (1) of a certain transformation of a proportion, as in Book V, and the arithmetical Books, (1} as here, of certain of the angles formed by parallels with a straight line crossing them, Attentate angles are, according to Euclid as interpreted by Produs, those which are not on the same side of the transversal, and are not adjacent, but are separated by the transversal, both being within the parallels but one "above" and the other "below. - ' The meaning is natural enough if we imagine the four internal angles to be taken in cyclic order and alternate angles to be any two of them not successive but separated by one angle of the four. 9. in the direction of B, D or towards A, C, literally "towards the parts B, D or towards A, C," iri rk B, &> Mm 4 M *i A, T. With this proposition begins the second section of the first Book. Up to this point Euclid has dealt mainly with triangles, their construction and their properties in the sense of the relation of their parts, the sides and angles, to one another, and the comparison of different triangles in respect of their parts, and of their area in the particular cases where they are congruent. The second section leads up Co the third, in which we pass to relations between the areas of triangles, parallelograms and squares, the special feature being a new conception of equality of areas, equality not dependent on amgrvenu. This whole subject requires the use of parallels. Consequently the second section beginning at 1. 27 establishes the theory of parallels, introduces the cognate matter of the equality of the sum of the angles of a triangle to two right angles (1. 32), and ends with two propositions forming the transition to the third section, namely 1, 33, 34, which introduce the parallelo- gram for the first time, Aristotle on parallels. We have already seen reason to believe that Euclid's personal contribution to the subject was nothing less than the formulation of the famous Postulate 5 (see the notes on that Postulate and on Def. 13), since Aristotle indicates that the then current theory of parallels contained a petitio principii, and presumably therefore it was Euclid who saw the defect and proposed the remedy. But it is clear that the propositions 1. 27, 18 were contained in earlier text-books. They were familiar to Aristotle, as we may judge from two interesting passages. (1) In Anal. Pott. 1. 5 he is explaining that a scientific demonstration must not only prove a fact of every individual of a class (mri rarro?) but must' prove it primarily and generally true (rpaTur xaBokov) of the whok of the class as one ; it will not do to prove it first of one part, then of another part, and so on, until the class is exhausted. He illustrates this {74 a 13 — 16) by a reference to parallels : " If then one were to show that right (angles) do not meet, the proof of this might be thought to depend on the fact that this is true of all (pairs of actual) right angles. But this is not so, inasmuch as the result does not follow because (the two angles are) equal (to two right I. 27, 28] PROPOSITIONS 27, 28 309 angles) in the particular way [i.e. because each is a right angle], but by virtue of their being equal {to two right angles) in any way whatever [i.e. because the sum only needs to be equal to two right angles, and the angles themselves may vary as much as we please subject to this]," (2) The second passage has already been quoted in the note on Def. 23 : " there is nothing surprising in different hypotheses leading to the same false conclusion ; e.g. the conclusion that parallels meet might equally be drawn from either of the assumptions (a) that the interior (angle) is greater than the exterior or (b) that the sum of the angles of a mangle is greater than two right angles" (Anal, Prior, n. 17, 66 a n — 15). I do not quite concur in the interpretation which Heiberg places upon these passages (Mathematisehes su Aristoteks, pp. 18 — 19), He says, first, that the allusion to the "interior angle" being "greater than the exterior" in the second passage shows that the reference in the first passage must be to Eucl. 1. 28 and not to 1. 27, and he therefore takes the words Sri wSl lv) ^nd (AG ff.GIfC); (b) two external angles, viz. the pairs (EGB, DHJF) and (EGA, CHF) ; (e) one external and one internal angle, viz. the pairs (EGB, GffD\ (FHD, HGB), (EGA, GHC) and (EMC, HGA). i. 28, i<>] PROPOSITIONS 28, 29 311 And (2) the possible pairs on different sides of the transversal may consist respectively of (a) two internal angles, viz. the pairs (AGH, GHD) and {CHG, HGB); {b) two external angles, vi*. the pairs {AGE, DHF) and {EGB, CHF); if) one external and one internal, viz. the pairs {AGE, GHD), {EGB, GHC), (FHC, HGB) aird {FHD, HGA). The angles are equal in the pairs (i) {e), {2) (a) and (2) (/>), and the sum is equal to two right angles in the case of the pairs (1 ) (a), (1) {6) and (2) {c\ For his criteria Euclid selects the cases (2) (o) [i. 27] and (1) {(), {i)(jn) [1. 28], leaving out the other three, which are of course equivalent but are not quite so easily expressed. From Proclus' note on 1. 28 (p. 361) we iearn that one Aigeias {? Aineias) of Hierapolis wrote an epitome or abridgment of the Elements. This seems to be the only mention of this editor and his work; and they are only mentioned as having combined Eucl. 1. 27, 28 into one proposition. To do this, or to make the three hypotheses the subject of three separate theorems, would, Proclus thinks, have been more natural than to deal with them, as Euclid does, in two propositions. Proclus has no suggestion for explaining Euclid's arrangement unless the ground were that 1. 27 deals with angles on different sides, 1. 28 with angles on one and the same side, of the transversal. But may not the reason have been one of convenience, namely that the criterion of 1. 27 is that actually used to prove parallelism, and is moreover the basis of the construction of parallels in I. 31, while 1. 28 only reduces the other two hypotheses to that of 1. 27, so that precision of reference, as well as clearness of exposition, is better secured by the arrangement adopted ? Proposition 29. A straight line falling on parallel straight lines makes the alternate angles equal to one another, the exterior angle equal to the interior and opposite angle, and the interior angles on the same side equal to two right angles. 5 For let the straight line EF fall on the parallel straight lines AB, CD; I say that it makes the alternate angles AGH, GHD equal, the exterior angle EGB equal to the interior and opposite angle GHD, and the interior angles on the same 10 side, namely BGH, GHD, equal to two right angles. For, if the angle AGH is unequal to the angle GHD, one of them is greater. Let the angle AGH be greater. is .Let the angle BGH be added to each ; therefore the angles A GH, BGH are greater than the angles BGH, GHD. 3'a BOOK I [i. *9 *> But the angles A GH, BGH are equal to two right angles; p. 13] therefore the angles BGH, GHD are less than two right angles. But straight lines produced indefinitely from angles less than two right angles meet ; [Post. 5J 35 therefore AB, CD, if produced indefinitely, will meet; but they do not meet, because they are by hypothesis parallel. Therefore the angle AGH is not unequal to the angle GHD, and is therefore equal to it. 30 Again, the angle AGH is equal to the angle EGB ; [* 15] therefore the angle EGB is also equal to the angle GHD. [C H. 1] Let the angle BGH be added to each ; therefore the angles EGB, BGH are equal to the 3S angles BGH, GHD. [C. N. j] But the angles EGB, BGH are equal to two right angles ; therefore the angles BGH, GHD are also equal to two right angles. Therefore etc q. e. d. 13. straight lines produced indefinitely from angles less than two right angles, of H aV' {\avirbrwy ff Sio ofifltvy fK^aWliuf *a. til drttpv vvilvIttww, a variation from the more explicit language of Postulate j. A good deal is left to be understood, namely that the straight lines begin from points at which they meet a transversal, and make with it internal angles on the same aide the sum of which is less than two right angles. 16. because they are by hypothesis parallel, literally " because they are supposed parallel," &d t6 rapoXK^Xavt afrai itTrtmtiaBai. Proof by "Playfair's" axiom. If, instead of Postulate 5, it is preferred to use " Playfair's " axiom in the proof of this proposition, we proceed thus. To prove that the alternate angles AGH, GHD are equal. If they are not equal, draw another straight line KL through G making the angle KGH equal to the angle GHD. Then, since the angles KGH, GHDaie equal, KL is parallel to CD. [1. 17] Therefore two straight lines KL, AB intersecting at G are bath parallel to the straight line CD : which is impossible {by the axiom). Therefore the angle AGH cannot but be equal to the angle GHD. The rest of the proposition follows as in Euclid. t. i 9 ] PROPOSITION 29 313 Proof of Euclid's Postulate 5 from "PI ay fair's" axiom. ■Let AB, CD make with the transversal EF the angles AEF, EFC :ogether less than two right angles. To prove that AS, CD meet towards A, C. _ B Through E draw GH making with -E-Fthe angle Q-^^^^^^r^rr.^ GEF equal (and alternate) to the angle EFD. \ Thus GH is parallel to CD, [1. 27] \ Then (1) AB must meet CZ>in one direction or c F the other. For, if it does not, AB must be parallel to CD; hence we have two straight lines AB, GH intersecting at E and both parallel to CD : which is impossible. Therefore AB, CD must meet. (2) Since AB, CD meet, they must- form a triangle with EF. But in any triangle any two angles are together less than two right angles. Therefore the angles AEF, EFC (which are less than two right angles), and not the angles BEF, EFD (which are together greater than two right angles, by 1. 13), are the angles of the triangle; that is, EA, FC meet in the direction of A, C, or on the side of EF on which are the angles together less than two right angles. The usual course in modern text-books which use " Play fair's " axiom in lieu of Euclid's Postulate is apparently to prove i a 9 by means of the axiom, and then Euclid's Postulate by means of 1. 29, De Morgan would introduce the proof of Postulate 5 by means of " Playfair's " axiom before 1. 29, and would therefore apparently prove I. ag as Euclid does, without any change. As between Euclid's Postulate 5 and " Playfair's " axiom, it would appear that the, tendency in modern text-books is rather in favour of the latter. Thus, to take a few noteworthy foreign writers, we find that Rausenberger stands almost alone in using Euclid's Postulate, while Hilbert, Henrict and Trentlein, Rouch^ and De Comberousse, Enriques and Amaldi all use " Playfair's " axiom. Yet the case for preferring Euclid's Postulate is argued with some force by Dodgson (Euclid and his modem Rivals, pp. 44 — fi). He maintains (i) that " Playfair's " axiom in fact involves Euclid's Postulate, but at the same time involves more than the latter, so that, to that extent, it is a needless strain on the faith of the learner. This is shown as follows. Given AB, CD making with EF l\\z angles AEF, EFC together less than two right angles, draw GH through E so that the angles GEF, EFC are together equal to two right angles. Then, by 1, 28, GH, CD are "separational," We see then that any lines which have the property (a) that they make with a transversal angles less than two right angles have also the property (£) that one of them intersects a straight line which is "separational" from the other. Now Playfair's axiom asserts that the lines which have property (p) meet if produced : for, if they did not, we should have two intersecting straight lines both " separational " from a third, which is impossible. We then argue that lines having property (a) meet because lines having property (a) are lines having property (ji). But we do not know, until we have proved 1. 29, that all pairs of lines having property (fi) have also property 3"4 BOOK I [(. *o, 30 (a). For anything we know to the contrary, class (0) may be greater than class (a). Hence, if you assert anything of class (fi), the logical effect is more extensive than if you assert it of class (o) ; for you assert it, not only of that portion of class 0) which is known to be included in class (a), but also of the unknown (but possibly existing) portion which is not so included. (a) Euclid's Postulate puts before the beginner clear and positive con- ceptions, a pair of straight lines, a transversal, and two angles together less than two right angles, whereas "PlayfairV' axiom requires him to realise a pair of straight lines which never meet though produced to infinity : a negative conception which does not convey to the mind any clear notion of the relative position of the lines. And (p, 68) Euclid's Postulate gives a direct criterion for judging that two straight lines meet, a criterion which is constantly required, e.g. in I. 44. It is true that the Postulate can be deduced from " Playfair's " axiom, but editors frequently omit to deduce it, and then tacitly assume it afterwards : which is the least justifiable course of all. Proposition 30. Straight lines parallel to the same straight line are also parallel to one another. Let each of the straight lines AB, CD be parallel to EF\ I say that AB is also parallel to CD. 5 For let the straight line GK fall upon them. Then, since the straight line GK has fallen on the parallel straight lines AB, EF, to the angle AGK is equal to the angle GHF. [1. *a] Again, since the straight line GK has fallen on the parallel straight lines EF, CD, the angle GHF is equal to the angle GKD. [i- 29] 1 j But the angle A GK was also proved equal to the angle GHF; therefore the angle AGK is also equal to the angle GKD; \C.N.i\ and they are alternate. 30 Therefore AB is parallel to CD. Q. E. D. JO. The usual comlusiim in general terms ("Therefore etc.") repenting the oouociaiion ts, curiously enough, wanting at the end of this proposition. The proposition is, as De Morgan points out, the iogieai equivalent of "PlayfaiiV' axiom. Thus, if X denote "pairs of straight lines intersecting one '■ 3°. 30 PROPOSITIONS 29—31 3«5 another," Y" pairs of straight lines parallel to one and the same straight line," we have No X is Y, and it follows logically that No Y is X. De Morgan adds that a proposition is much wanted about parallels (or perpendiculars) to two straight lines respectively making the same angles with one another as the latter do. The proposition may be enunciated thus : If the sides of one angle be respectively (1) parallel or (2) perpendicular to the sides of another angle, the two angles are either equal Or supplementary. (1) Let DE be parallel to AB and GEF parallel toBC. To prove that the angles ABC, DEG are equal and the angles ABC, DEF supplementary. Produce DE to meet BC in H. Then [1. 29] the angle DEG is equal to the angle DMC, and the angle ABC is equal to the angle DHC. Therefore the angle DEG is equal to the angle ABC; whence also the angle DEF is supplementary to the angle ABC- it) Let ED be perpendicular to AB, and GEF perpendicular to BC. To prove that the angles ABC, DEG are equal, and the angles ABC, DEF supplementary. Draw ED' at right angles to ED on the side of it opposite to B, and draw EG' at right angles to EFoa the side of it opposite to B. Then, since the angles BDE, DED, being right angles, are equal, ED is parallel to BA. [1. 27] Similarly EG' is parallel to BC. Therefore [Part (1)] the angle DEG' is equal to the angle ABC. But, the right angle DED being equal to the right angle GEG, if the common angle GED be subtracted, the angle DEG is equal to the angle D"EG '. Therefore the angle DEG is equal to the angle ABC; and hence the angle DEF is supplementary to the angle ABC. Proposition 31. Through a given point to draw a straight line parallel to a given straight line. Let A be the given point, and BC the given straight line ; thus it is required to draw through the point A a straight line parallel to the straight line BC. 3»& BOOK 1 \i. JIf 3 j Let a point D be taken at random on BC, and let AD be joined ; on the straight line DA, and at the point A on it, let the g £ angle DAE be constructed equal / to the angle W/?C [i. 23] ; and let the / straight line AF be produced in a e d c straight line with EA. Then, since the straight line AD falling on the two straight lines BC, EF has made the alternate angles EAD, ADC equal to one another, therefore EAF is parallel to BC. [i. 17] Therefore through the given point A the straight line EAF has been drawn parallel to the given straight line BC. Q. E. F. Proclus rightly remarks (p. 376, 14 — 20) that, as it is implied in 1. iz that only one perpendicular can be drawn to a straight line from an external point, so here it is implied that only one straight line can be drawn through a point parallel to a given straight line. The construction, be il observed, depends only upon I. 27, and might therefore have come directly after that proposition. Why then did Euclid postpone it until after 1, 29 and 1. 30? Presumably because he considered it necessary, before giving the construction, to place beyond all doubt the fact that only one such parallel can be drawn. Proclus infers this fact from 1. 30 ; for, he says, if two straight lines could be drawn through one and the same point parallel to the same straight line, the two straight lines would be parallel, though intersecting at the given point : which is impossible. I think it is a fair inference that Euclid would have considered it necessary to justify the assumption that only one parallel can be drawn by some such argument, and that he deliberately determined that his own assumption was more appropriate to be made the subject of a Postulate than the assumption of the uniqueness of the parallel. Proposition 32. In any triangle, if one of the sides be produced, the exterior angle is equal to the two interior and opposite angles, and the three interior angles of the triangle are equal to two right angles. Let ABC be a triangle, and let one side of it BC be produced to D ; I say that the exterior angle A CD is equal to the two interior and opposite angles CAB, ABC, and the three interior angles of the triangle ABC, BCA, CAB are equal to two right angles. I. 3 a] PROPOSITIONS 31, 31 317 For let CE be drawn through the point C parallel to the straight line AB. [i- 3'] Then, since AB is parallel to CE, and AC has fallen upon them, the alternate angles EAC, ACE are equal to one another. [1. 19] Again, since AB is parallel to CE, and the straight line BD has fallen upon them, the exterior angle ECD is equal to the interior and opposite angle ABC. [1. 29] But the angle ACE was also proved equal to the angle BAC; therefore the whole angle A CD is equal to the two interior and opposite angles EAC, ABC. Let the angle A CB be added to each ; therefore the angles A CD, ACE are equal to the three angles ABC, BCA, CAB. But the angles A CD, ACE are equal to two right angles; , D- 13] therefore the angles ABC, BCA t CAB are also equal to two right angles. Therefore etc. q. E. D. This theorem was discovered in the very early stages of Greek geometry. What we know of the history of it is gathered from three allusions found in Eutocius, Proclus and Diogenes Laertius respectively. 1. Eutocius at the beginning of his commentary on the Conies of Apollonius (ed. Heiberg, Vol. u- p. r7o) quotes Geminus as saying that "the ancients (oE J^^uum) investigated the theorem of the two right angles in each individual species of triangle, first in the equilateral, again in the isosceles, and afterwards in the scalene triangle, and later geometers demonstrated the general theorem to the effect that in any triangle the three interior angles are equal to two right angles." a. Now, according to Proclus (p. 379, 2 — 5), Eudemus the Peripatetic refers the discovery of this theorem to the Pythagoreans and gives what he affirms to be their demonstration of it This demonstration will be given below, but it should be remarked that it is general, and therefore that the "later geometers" spoken of by Geminus were presumably the Pythagoreans, whence it appears that the "ancients" contrasted with them must have belonged to .the time of Thales, if they were not his Egyptian instructors. 3. That the truth of the theorem was known to Thales might also be inferred from the statement of Pamphile (quoted by Diogenes Laertius, 1. 34 — Si p 6, ed. Cobet) that " he, having leamt geometry from the 318 BOOK I [1. 32 Egyptians, was the first to inscribe a right-angled triangle in a circle and sacrificed an ox" (on the strength of it) ; in other words, he discovered that the angle in a semicircle is a right angle. No doubt, when this fact was once discovered (empirically, say), the consideration of the two isosceles triangles having the centre for vertex and the sides of the right angle for bases respectively, with the help of the theorem of Eucl. t. 5, also known to Thales, would easily lead to the conclusion that the sum of the angles of a right-angled triangle is equal to two right angles, and it could be readily inferred that the angles of any triangle were likewise equal to two right angles (by resolving it into two right-angled triangles). But it is not easy to see how the property of the angle in a semicircle could he pruned except (in the reverse order) by means of the equality of the sum of the angles of a right-angled triangle to two right angles ; and hence it is most natural to suppose, with Cantor, that Thales proved it (if he did prove it) practically as Euclid does in lit. 31, i.e. by means of 1. 31 as applied to right-angled triangles at all events. If the theorem of 1. 32 was proved before Thales' time, or by Thales himself, by the stages indicated in the note of Geminus, we may be satisfied that the reconstruction of the argument of the older proof by Hankel (pp. 96 — -7) and Cantor (i,, pp. 143 — 4) is not far wrong. First, it must have been observed that six angles equal to an angle of an equilateral triangle would, if placed adjacent to one another round a common vertex, fill up the whole space round that vertex. It is true that Proclus attributes to the Pythagoreans the general theorem that only three kinds of regular polygons, the equilateral triangle, the square and the regular hexagon, can fill up the entire space round a point, but the practical knowledge that equilateral triangles have this property could hardly have escaped the Egyptians, whether they made floors with tiles in the form of equilateral triangles or regular hexagons (Allman, Greek Geometry from Tkalef to Euclid, p. 11) or joined the ends of adjacent radii of a figure like the six-spoked wheel, which was their common form of wheel from the time of Ramses II. of the nineteenth Dynasty, say 1 300 B.C. (Cantor, i lt p, 109). It would then be clear that six angles equal to an angle of an equilateral triangle are equal to four right angles, and therefore that the three angles of an equilateral triangle are equal to two right angles. (It would be as clear or clearer, from observation of a square divided into two triangles by a diagonal, that an isosceles right-angled triangle has each of its equal angles equal to half a right angle, so that an isosceles right-angled triangle must have the sum of its angles equal to two right angles.) Next, with regard to the equilateral triangle, it could not fail to be observed that, if AD were drawn from the vertex A perpendicular to the base BC, each of the two right-angled triangles so formed would have the sum of its angles equal to two right j angles; and this would be confirmed by completing the J rectangle ADCE, when it would be seen that the rectangle / (with its angles equal to four right angles) was divided by / its diagonal into two equal triangles, each of which had B the sum of its angles equal to two right angles. Next it would be inferred, as the result of drawing the diagonal of any rectangle and observing the equality of the triangles forming the two halves, that the sum of the angles of any right-angled triangle is equal to two right angles, and henqe (the two congruent right-angled triangles being then placed so as to form one isosceles triangle) that the same is true of any isosceles triangle. Only the last step remained, namely that of observing that any triangle could be regarded as the half of a rectangle (drawn as indicated in the next figure), or L 3»] PROPOSITION 3* 319 simply that any triangle could be divided into two right-angled triangles, whence it would be inferred that in general the sum of the angles of any triangle is equal to two P y^T >s sL right angles. i / I TS^. j Such would be the probabilities if we could j / ^^ j absolutely rely upon the statements attributed to '•,/_ i ^vi Pamphile and Geminus respectively. But in fact there is considerable ground for doubt in both cases. 1. Pamphiie's stoty of the sacrifice of an ox by Thales for joy at his discovery that the angle in a semicircle is a right angle is too suspiciously like the similar story told with reference to Pythagoras and his discovery of the theorem of Eucl. 1. 47 (Prod us, p. 426, 6—9). And, as if this were not enough, Diogenes Laertius immediately adds that ''others, among whom is Apollodorus the calculator (0 \oyum«K), say it was Pythagoras " (sc. who " inscribed the right-angled triangle in a circle "). Now Pamphile lived in the reign of Nero (a.d. 54 — 68) and therefore some 700 years after the birth of Thales (about 640 B.C.). I do not know on what Max Schmidt bases his statement {Kullurhistorische Beitrage tur Kcnntnis des grieehischm uttd romiuhen AUertums, 190C, p. 31) that "other, much older, sources name Pythagoras as the discoverer of the said proposition," because nothing more seems to be known of Apollodorus than what is stated here by Diogenes Laertius. But it would at least appear that Apollodorus was only one of several authorities who attributed the proposition to Pythagoras^ while Parrfphile is alone mentioned as referring it to Thales. Again, the connexion of Pythagoras with the investigation of the right-angled triangle makes it a priori more likely that it would be he who would discover its relation to a semicircle. On the whole, therefore, the attribution to Thales would seem to be more than doubtful. 2. As regards Geminus' account of the three stages through which the proof of the theorem of 1. 32 passed, we note, first, that it is certainly not confirmed by Eudemus, who referred to the Pythagoreans the discovery of the theorem that the sum of the angles of any triangle is equal to two right angles and says nothing about any gradual stages by which it was proved. Secondly, it must be admitted, I think, that in the evolution of the proof as reconstructed by Hank el the middle stage is rather artificial and unnecessary, since, once it is proved that any right-angled triangle has the sum of its angles equal to two right angles, it is just as easy to pass at once to any scalene triangle (which is decomposable into two unequal right-angled triangles) as to the isosceles triangle made up of two congruent right-angled triangles. Thirdly, as Heiberg has recently pointed out (Mathcmatisehes su Aristoteles, p. ao), it is quite possible that the statement of Geminus from beginning to end is simply due to a misapprehension of a passage of Aristotle (Anal. Post. 1. 5, 74 a 25). Aristotle is illustrating his contention that a property is not scientifically proved to belong to a class of things unless it is proved to belong primarily (xpwrav) and generally (kuAEXm) to the whole of the class. His first illustration relates to parallels making with a transversal angles on the same side together equal to two right angles, and has been quoted above in the note on 1. 27 (pp. 308 — 9). His second illustration refers to the transformation of a proportion aliernando, which (he says) "used at one time to be proved separately " for numbers, lines, solids, and times, although it admits of being proved of all at once by one demonstration. The third illustration is: "For the same reason, even if one should prove (0O8' ok t« &ti(g) with reference to 3*0 BOOK I [i. 32 each (sort of) triangle, the equilateral, scalene and isosceles, separately, that each has its angles equal to two right angles, either by one proof 01 by different proofs, he does not yet know that the triangle, i.e. the triangle in general, has its angles equal to two right angles, except in a sophistical sense, even though there exists no triangle other than triangles of the kinds mentioned For he knows it, not qua\ triangle, nor of tvtty triangle, except in a numerical sense (not apiBy£v)\ he does not know it nationally (nwr tl&ot) of every triangle, even though there be actually no triangle which he does not know." The difference between the phrase " used at one time to be proved " in the second illustration and " if any one should prove " in the third appears to indicate that, while the former referred to a historical fact, the latter does not; the reference to a person who should prove the theorem of 1. 3* for the three kinds of triangle separately, and then claim that he had proved it generally, states a purely hypothetical case, a mere illustration. Vet, coming after the historical fact stated in the preceding illustration, it might not unnaturally give the impression, at first sight, that it was historical too. On the whole, therefore, it would seem that we cannot safely go behind the dictum of Eudemus that the discovery and proof of the theorem of i. 32 in all its generality were Pythagorean. This does not however preclude its having been discovered by stages such as those above set out after Hankel and Cantor. Nor need it be doubted that Thales and even his Egyptian instructors had advanced some way on the same road, so far at all events as to see that in an equilateral triangle, and in an isosceles right-angled triangle, the sum of the angles is equal to two right angles. The Pythagorean proof. This proof, handed down by Eudemus (Proclus, p. 379,-2 — 15), is no less elegant than that given by Euclid, and is a natural development from the last figure in the recon- structed argument of Hankel. It would be seen, after the theory of parallels was added to geometry, that the actual drawing of the perpendicular and the complete rectangle on BC as base was unnecessary, and that the parallel to BC through A was all that was required. Let ABC be a triangle, and through A draw DE parallel to BC. [1. 31] Then, since BC, DE are parallel, the alternate angles DAB, ABC are equal, [1. 39] and so are the alternate angles EAC, ACB also. Therefore the angles ABC, ACS are together equal to the angles DAB, EAC. Add to each the angle BAC; therefore the sum of the angles ABC, ACB, BAC is equal to the sum of the angles DAB, BAC, CAE, that is, to two right angles. Euclid's proof pre- Euclidean. The theorem of 1. 32 is Aristotle's favourite illustration when he wishes to refer to some truth generally acknowledged, and so often does it occur that it is often indicated by two or three words in themselves hardly intelligible, e.g. to Suo-If opSatt (Anal. Post. 1. 24, 85 b 5) and tijnipjjfi Tcayri rpiytivip to &!o (ibid. 85 b 11). One passage (MttapA. 1051 a 14) makes it clear, as Heiberg {op. at. i. 31] PROPOSITION 3* 321 p. 19) acutely observes, that in the proof as Aristotle knew it Euclid's construction was used. " Why does the triangle make up two right angles ? Because the angles about one point are equal to two right angles. If then the parallel to the side had been drawn up (civ^kto), the fact would at once have been clear from merely looking at the figure." The words " the angles about one point " would equally fit the Pythagorean construction, but " drawn upwards " applied to the parallel to a side can only indicate Euclid's. Attempts at proof independently of parallels. The most indefatigable worker on these lines was Legendre, and a sketch of his work has been given in the note on Postulate 5 above. One other attempted proof needs to be mentioned here because it has found much favour. I allude to Thibaut's method. This appeared in Thibaut's Grundriss der reinen Mathematik, Gottingen {2 ed. 1809, 3 ed. 1818), and is to the following effect. Suppose CB produced to D, and let BD (produced to any necessary extent either way) revolve in one direction (say clockwise) first about B into the position BA, then about A into the position of AC produced both ways, and lastly about C into the position CB produced both ways. The argument then is that the straight line BD has revolved through the sum of the three exterior angles of the triangle. But, since it has at the end of the revolution assumed a position in the same straight line with its original position, it must have revolved through four wight angles. Therefore the sum of the three exterior angles is equal to four right angles ; from which it follows that the sum of the three angles of the triangle is equal to two right angles. But it is to be observed that the straight line BD revolves about different points in it, so that there is translation combined with rotatory motion, and it is necessary to assume as an axiom that the two motions are independent, and therefore that the translation may be neglected. Schumacher (letter to Gauss of 3 May, 1831) tried to represent the rotatory motion graphically in a second figure as mere motion round a point ; but Gauss (letter of 17 May, 1831) pointed out in reply that he really assumed, without proving it, a proposition to the effect that " If two straight tines (1) and (2) which cut one another make angles A , A" with & straight line (3) cutting both of them, and if a straight line {4) in the same plane is likewise cut by (1) at an angle A', then {4) will be cut by (2) at the angle A". But this proposition not only needs proof, but we may say that it is, in essence, the very proposition to be proved " (see Engel and Stackel, Die Theorit der ParalkUinien von Euktid it's auf Gauss, 1895, p. 230). How easy it is to be deluded in this way is plainly shown by Proclus' attempt on the same lines. He says (p. 384, 13 — 21) that the truth of the theorem is borne in upon us by the help of " common notions " only. For, if we conceive a straight line with two perpendiculars drawn to it at its ex- tremities, and if we then suppose the perpendiculars to (revolve about their feet and) approach one another, so as to form a triangle, we see that, BOOK I [I. 3 », 33 to the extent to which they converge, they diminish the right, angles which they made with the straight line, so that the amount taken from the right angles is also the amount added to the vertical angle of the triangle, and the three angles are necessarily made equal to two right angles," But a moment's reflection shows that, so far from being founded on mere " common notions," the supposed proof assumes, to begin with, that, if the perpendiculars ap- proach one another ever so little, they will then form a triangle immediately, LO., it assumes Postulate 5 itself; and the fact about the vertical angle can only be seen by means of the equality of the alternate angles exhibited by drawing a perpendicular from the vertex of the triangle to the base, i.e. % parallel to either of the original perpendiculars. Extension to polygons. The two important corollaries added to 1. 32 in Simson's edition are given by Proclus ; but Proclus' proof of the first is different from, and perhaps somewhat simpler than, Simson's. 1. The turn of the interior angles of a convex rectilineal figure is equal to twice as many right angles as the figure has sides, less four. For let one angular point A be joined to all the other angular points with which it is not con- nected already. The figure is then divided into triangles, and mere inspection shows (t) that the number of triangles is two less than the number of sides in the figure, (3) that the sura of the angles of all the triangles is equal to the sum of all the interior angles of the figure. Since then the sum of the angles of each triangle is equal to two nght angles the sum of the interior angles of the figure is equal to i(n — 2) right angles, i.e. (zn - 4) right angles, where n is the number of sides in the figure. 3. The exterior angles of any convex rectilineal figure an together equal to four right angles. For the interior and exterior angles together are equal to 27i right angles, where n is the number of sides. And the interior angles are together equal to (a«-4> right angles. Therefore the exterior angles are together equal to four right angles. This last property is already quoted by Aristotle as true of all rectilineal figures in two passages (Anal. Post. 1. 24, 85 b 38 and 11. 17, 99 a 10). Proposition 33. The straight lines joining equal and parallel straight lines {at the extremities which are) in the same directions {respectively) are themselves also equal and parallel. Let AB, CD be equal and parallel, and let the straight 5 lines AC, BD join them (at the extremities which are) in the same directions (respectively) ; i. 33. 34] PROPOSITIONS 32—34 3*3 I say that AC, BD are also equal and parallel. Let BC be joined. Then, since AB is parallel to CD, 10 and BC has fallen upon them, the alternate angles ABC, BCD are equal to one another. [1. 29] And, since AB is equal to CD, and BC is common, 15 the two sides AB, BC are equal to the two sides DC, CB ; and the angle ABC is equal to the angle BCD ; therefore the base AC is equal to the base BD, and the triangle ABC is equal to the triangle DCB, and the remaining angles will be equal to the remaining angles 20 respectively, namely those which the equal sides subtend ; [1. 4] therefore the angle ACB is equal to the angle CBD. And, since the straight line BC falling on the two straight lines AC, BD has made the alternate angles equal to one another, J S AC is parallel to BD. * [1. 27] And it was also proved equal to it. Therefore etc. Q. e. d. 1. joining. ..{at the extremities which are) In the same directions (respectively). I have for clearness' sake inserted the words irv brackets though they are not in the original Greek, which has "joining., .in the same directions" or "on the same sides," ivl t4 o£ra pip/i tT,fcuy*(i6vi^or (parallel- lined) was analogous to that of ib6trypap.im {straight- lined or rectilineal). 17, 18, +0. DCB and 36. DC, CB. The Greek has in these places "BCD" and %% CD, EC respectively. Cf. note on I. 33, lines 15, 18. After specifying the particular kinds of parallelograms (squares and rhombi) in which the diagonals bisect the angles which they join, as well as the areas, and those (rectangles and rhomboids) in which the diagonals do not bisect the angles, Proclus proceeds {pp. 390 sqq.) to analyse this proposition with reference to the distinction in Aristotle's Anal, Post. (1. 4, 5, 73 a 21 — 74 b 4) between attributes which are only predicable of every individual thing (iitb mrm) in a class and those which are true of it primarily {tovtou wpiorov) and generally (ko$6X.ov), We are apt, says Aristotle, to mistake a proof Kara warm for a proof tovtou irpwrou iraSoXov because it is either impossible to find a higher generality to comprehend all the particulars of which the predicate is true, or to find a name for it. {Part of this passage of Aristotle has been quoted above in the note on 1. 32, pp. 319 — 320.) Now, says Proclus, adapting Aristotle's distinction to theorems, the present proposition exhibits the distinction between theorems which are general and theorems which are not general. According to Proclus, the first part of the proposition stating that the opposite sides and angles of a parallelogram are equal is general because the property is only true of parallelograms ; but the second part which asserts that the diameter bisects the area is not general because it does not include all the figures of which this property is true, e.g. circles and ellipses. Indeed, says Proclus, the first attempts upon problems seem usually to have been of this partial character (fupiKiirtpat), and generality was only attained by degrees. Thus " the ancients, after investigating the fact that the diameter bisects an ellipse, a circle, and a parallelogram respectively, proceeded to investigate what was common to these cases," though "it is difficult to show what is common to an ellipse, a circle and a parallelogram." I doubt whether the supposed distinction between the two parts of the proposition, in point of "generality," can be sustained. Proclus himself admits that it is presupposed that the subject of the proposition is a quadrilateral, because there are other figures (e.g. regular polygons of an even number of sides) besides parallelograms which have their opposite sides and angles equal; therefore the second part of the theorem is, in this respect, no more general than the other, and, if we are entitled to the tacit limitation of the theorem to quadrilaterals in one part, we are equally entitled to it in the other. It would almost appear as though Proclus had drawn the distinction for the mere purpose of alluding to investigations by Greek geometers on the general subject of diameters of all sorts of figures; and it may have been these which brought the subject to the point at which Apollonius could say in the first definitions at the beginning of his Conies that "In any bent line, such as is in one plane, I give the name diameter to any straight line which, being drawn from the bent line, bisects all the straight lines (chords) drawn in the line parallel to any straight line." The term bent line (capuvAi; ypafi/*n') includes, e.g. in Archimedes, not only curves, but any composite line made 3» 35 up of straight lines and curves joined together in any manner. It is of course clear that either diagonal of a parallelogram bisects all lines drawn within the parallelogram parallel to the other diagonal. An-Nairizi gives after 1.31a neat construction for dividing a straight line into any number of equal parts (ed. Curtze, p. 74, ed. Besthorn-Heiberg, pp. 141 — 3) which requires only one measurement repeated, together with the properties of parallel lines including 1. 33, 34. As 1. 33, 34 are assumed, I place the problem here. The particular case taken is the problem of dividing a straight line into tkret equal parts. Let AB be the given straight line. Draw AC, BD at right angles to it on opposite sides. An-Nairizi takes AC, BD of the same length and then bisects AC at E and BD at F. But of course it is even simpler to measure AE, EC along one' perpendicular equal and of any length, and BF, FD along the other also equal and of the same length. Join ED, CF meeting AB in G, H respectively. Then shall AG, GH, HB all be equal. Draw HK parallel to AC, or at right angles to AB. Since now EC, FD are equal and parallel, ED, CFaxe equal and parallel. [1. 33] And HK was drawn parallel to AC. Therefore ECHK is a parallelogram ; whence KH is equal as well as parallel to EC, and therefore to EA. The triangles EAG, KHG have now two angles respectively equal and the sides AE, HK equal Thus the triangles are equal in all respects, and AG is equal to GH. Similarly the triangles KHG, FBH are equal in all respects, and GH is equal to SB. If now we wish to extend the problem to the case where AB is to be divided into » parts, we have only to measure («-i) successive equal lengths along AC and {n— 1) successive lengths, each equal to the others, along BD. Then join the first point arrived at on AC to the last point on BD, the second on AC to the last but one on BD, and so on ; and the joining lines cut AB in points dividing it into n equal parts. Proposition 35. Parallelograms -which are on the same base and in the same parallels are equal to one another. Let A BCD, EBCF be parallelograms on the same base BC and in the same parallels AF, BC ; s I say that ABCD is equal to the parallelogram EBCF, For, since ABCD is a parallelogram, AD is equal to BC. [1. 34] i. 35] PROPOSITIONS 34, 35 3*7 For the same reason also EF is equal to BC, 10 so that AD is also equal to EF; \C. N. 1] and DE is common ; therefore the whole AE is equal to the whole DF. [C. M a] But AB is also equal to DC; [t 34J therefore the two sides EA, AB are equal to the two sides is FD t DC respectively, and the angle FDC is equal to the angle EAB, the exterior to the interior ; [1. 39] therefore the base EB is equal 30 to the base FC, and the triangle EAB will be equal to the triangle FDC. [tfl Let DGE be subtracted from each ; therefore the trapezium ABGD which remains is equal to the trapezium EGCF which remains, [C. N. 3] »s Let the triangle l75C be added to each ; therefore the whole parallelogram A BCD is equal to the whole parallelogram EBCF. [C. N. 2] Therefore etc. Q. E. D. 11. FDC. The text has " DFC." 11. Lei DOE be subtract ed. Euclid speaks of the triangle DGE without any explanation that, in the cue which he takes (where AD, EF have no point in common), BE, CD must meet at a point G between the two parallels. He allow* this to appear from the figure simply. Equality in a new sense. It is important to observe that we are in this proposition introduced for the first time to a new conception of equality between .figures. Hitherto we have had equality in the sense of congruence only, as applied to straight lines, angles, and even triangles (cf. t. 4). Now, without any explicit reference to any change in the meaning of the term, figures are inferred to be equal which are equal in area or in content but need not be of the same form. No definition of equality is anywhere given by Euclid ; we are left to infer its meaning from the few axioms about " equal things." It will be observed that in the above proof the " equality " of two parallelograms on the same base and between the same parallels is inferred by the successive steps (1) of subtracting one and the same area (the triangle DGE) from two areas equal in the sense of congruence (the triangles ABB, DFC), and inferring that the remainders (the trapezia ABGD, EGCF) are "equal"; (2) of adding one and 3i8 BOOK I [i. 3S the same area {the triangle GBC) to each of the latter " equal " trapezia, and inferring the equality of the respective sums (the two given parallelograms). As is well known, Simson (after Clairaut) slightly altered the proof in order to make it applicable to all the three possible cases. The alteration substituted one step of subtracting congruent areas (the triangles AE3, DEC) froth one and the same area (the trapezium ABCF) for the two steps above shown of first subtracting and then adding a certain area. While, in either case, nothing more is explicitly used than the axioms that, if equals be added to equals, the wholes are equal and that, if equals be subtracted from equals, the remainders are equal, there is the further tacit assumption that it b indifferent to what part or from what part of the same or equal areas the same or equal areas are added or subtracted. De Morgan observes that the postulate "an area taken from an area leaves the same area from whatever part it may be taken " is particularly important as the key to equality of non- rectiiineal areas which could not be cut into coincidence geometrically. Legendre introduced the word equivalent to express this wider sense of equality, restricting the term equal to things equal in the sense of congruent ; and this distinction has been found convenient. I do not think it necessary, nor have I the space, to give any account of the recent developments of the theory of equivalence on new lines represented by the researches of W. Bolyai, Duhamel, De Zolt, Stolz, Schur, Veronese, Hilbert and others, and must refer the reader to Ugo Amaldi's article Sulla teoria dell' cquivatenza in Quest ioni riguardanti k matematiche elementari, L (Bologna, 1912), pp. 145 — 198, and to Max Simon, liber die Entwichlung der Elcmcntar-geometrie im XIX. fahrhundert (Leiprig, 1906), pp. 115 — no, with their full references to the literature of the subject. I may however refer to the suggestive distinction of phraseology used by Hilbert (Grundlagen der Geometrie, pp. 39, 40) : (ij "Two polygons are called divisibly-equal (zertegungsg&ich) if they cap be divided into a finite number of triangles which are congruent two and two." (z) "Two polygons are called equal in content (inhaltsgletch) or of equal content if it is possible to add divisibly-equal polygons to them in such a way that the two combined polygons are divisibly-equal." (Amaldi suggests as alternatives for the terms in (1) and (2) the expressions equivalent by sum and equivalent by difference respectively.) From these definitions it follows that "by combining divisibly-equal polygons we again arrive at divisibly-equal polygons; and, if we subtract divisibly-equal polygons from divisibly-equal polygons, the polygons remaining are equal in content." rhe proposition also follows without difficulty that, " if two polygons are divisibly-equal to a third polygon, they are also divisibly-equal to one another ; and, if two polygons are equal in content to a third polygon, they are equal in content to one another." The different cases. As usual, Proclus (pp. 399—400), observing that Euclid has given only the most difficult of the three possible cases, adds the other two with separate proofs. In the case where E in the figure of the proposition falls between A and D, he adds the congruent triangles ABE t DCF respectively to the smaller trapezium EBCD, instead of subtracting them (as Simson does) from the larger trapezium ABCF. 1-35] PROPOSITION 35 3*9 An ancient " Budget of Paradoxes." Proclus observes (p. 396, 12 sqq.) that the present theorem and the similar one relating to triangles are among the so-called paradoxical theorems of mathematics, since the un in strutted might well regard it as impossible that the area of the parallelograms should remain the same while the length of the sides other than the base and the side opposite to it may increase indefinitely. He adds that mathematicians had made a collection of such paradoxes, the so-called treasury of paradoxes (b wapdSofos tottm) — cf. the similar expressions Tojrot ivaXvoinytK (treasury of analysis) and tottos a&rpovonttu/ixvos — in the same way as the Stoics with their illustrations (<* 33> the genesis of unlimited (particulars) within defined limits, so in such theorems the unlimited (particular figures) are confined within defined places or loci (tovdi). And it is this boundary which is the cause of the equality ; for the height of the parallels, which remains the same, while an infinite number of parallelograms are conceived on the same base, is what makes them ail equal to one another." Proposition 36. Parallelograms which are on equal banes and in the same parallels are equal to one another. Let ABCD, EFGH be parallelograms which are on equal bases BC, FG and in the same parallels AH, BG ; I say that the parallelogram ABCD is equal to EFGH. For let BE, CH be joined. Then, since BC is equal to FG while FG is equal to EH, BC is also equal to EH. [C. X. 1] But they are also parallel. And EB, HC join them ; but straight lines joining equal and parallel straight lines (at the extremities which are) in the same directions (respectively) are equal and parallel. [1. 33] Therefore EBCH is a parallelogram. [1. 34] And it is equal to ABCD ; for it has the same base BC with it, and is in the same parallels BC, AHmth it [1. 35] For the same reason also EFGH is equal to the same EBCH; [..35] so that the parallelogram ABCD is also equal to EFGH. [ax 1] Therefore etc. q. e, o. 33* BOOK I [i. 37 Proposition 37. Triangles which are on the same base and in the same parallels are equal to one another. Let ABC, DBC be triangles on the same base BC and in the same parallels AD, BC ; 5 I say that the triangle ABC is equal to the triangle DBC Let AD be produced in both directions to E, F; E a d through B let BE be drawn parallel to CA, [1. 31] 10 and through C let CF be drawn parallel to BD. [1. 31] Then each of the figures EBCA, DBCF is a parallelogram ; and they are equal, ■5 for they are on the same base BC and in the same parallels BC, EF. [1. as] Moreover the triangle ABC is half of the parallelogram EBCA ; for the diameter AB bisects it. [1. 34] And the triangle DBC is half of the parallelogram DBCF; to for the diameter DC bisects it [t 34] [But the halves of equal things are equal to one another.] Therefore the triangle ABC is equal to the triangle DBC. Therefore etc. Q. E. D. Si. Here and in the next proposition t lei berg brackets the words "But the halves of equal things are equal to one another" on the ground that, since the Common Notion which asserted this fact was interpolated at a very early date (before the time of Theon), it is probable that the words here were interpolated at the same time. Cf. note above (p. J34) on the interpolated Common Notion. There is a lacuna in the text of Proclus' notes to 1. 36 and 1. 37. Apparently the end of the former and the beginning of the latter are missing, the mss. and the editio princeps showing no separate note for i. 37 and no lacuna, but going straight on without regard to sense. Proclus had evidently remarked again in the missing passage that, in the case of both parallelograms and triangles between the same parallels, the two sides which stretch from one parallel to the other may increase in length to any extent, while the area remains the same. Thus the perimeter in parallelograms or triangles is of itself no criterion as to their area. Misconception on this subject was rife among non-mathematicians; and Proclus {p. 403, 5 sqq.) tells us (1) of describers of countries (xtapoypafoi) who drew conclusions cegarding the size of cities from their perimeters, and (a) of certain members of communistic I. 37, 38] PROPOSITIONS 37. 38 333 societies in his own time who cheated their fellow members by giving them land of greater perimeter but less area than they took themselves, so that, on the one hand, they got a reputation for greater honesty while, on the other, they took more than their share of produce. Cantor (Gesck. d. Ma/A. i„ p, 172) quotes several remarks of ancient authors which show the prevalence of the same misconception. Thus Thucydides estimates the size of Sicily according to the time required for circumnavigating it. About 130 B.C. Polybius said that there were people who could not understand that camps of the same periphery might have different capacities. Quintilian has a similar remark, and Cantor thinks he may have had in his mind the calculations of Pliny, who compares the size of different parts of the earth by adding their length to their breadth. The comparison however of the areas of different figures of equal contour had not been neglected by mathematicians. Theon of Alexandria, in his commentary on Book 1. of Ptolemy's Syntaxis, has preserved a number of propositions on the subject taken from a treatise by Zenodorus mpl taofiirpw (rxriiidtiiiv {reproduced in Latin on pp. 1190 — tan of Hultsch's edition of Pappus) which was written at some date between, say, 200 b.c. -«.d 90 a.d., and probably not long after the former date, Pappus too has at the beginning of Book v. of bis Collection {pp. 308 sqq.) the same propositions, in which he appears to have followed Zenodorus pretty closely while making some changes in detail. The propositions proved by Zenodorus and Pappus include the following: (1) that, of all polygons of the same number of sides and equal perimeter, the equilateral and equiangular polygon is the greatest in area, {a) that, of regular polygons of equal perimeter, that is the greatest in area which has the most angles, {3} that a circle is greater than any regular polygon of equal contour, (4) that, of all circular segments in which the arcs are equal in length, the semicircle is the greatest. The treatise of Zenodorus was not con- fined to propositions about plane figures, hut gave also the theorem that, of alt solid figures the surfaces of which are equal, the sphere is the greatest in volume. Proposition 38. Triangles which are on equal bases and in the same Parallels are equal to one another. Let ABC, DEF be triangles on equal bases B C, EF and in the same parallels BF, AD ; I say that the triangle ABC is q a_ d h equal to the triangle DEF. For let AD be produced in both directions to G, H\ through B let BG be drawn parallel to CA, (l 31] and through F let FH be drawn parallel to DE. Then each of the figures GBCA, DEFH is a parallelo- gram ; and GBCA is equal to DEFH; 334 BOOK I [i. 38 for they are on equal bases BC, EF and in the same parallels BF, GH. [1. 36] Moreover the triangle ABC is half of the parallelogram GBCA ; for the diameter AB bisects it. [i. 34] And the triangle FED Is half of the parallelogram DEFH\ for the diameter DF bisects it, [i- 34] [But the halves of equal things are equal to one another.] Therefore the triangle ABC is equal to the triangle DEF. Therefore etc. Q. E. D. On this proposition Proclus remarks (pp. 405 — 6) that Euclid seems to him to have given in vi. 1 one proof including all the four theorems from I- 35 to 1. 38, and that most people had failed to notice this. When Euclid, he says, proves that triangles and parallelograms of the same altitude have to one another the same ratio as their bases, he simply proves all these propositions more generally by ( he use of proportion ; for of course to be of the same altitude is equivalent to being in the same parallels. It is true that vi. 1 generalises these propositions, but it must be observed that it does not prove the propositions themselves, as Proclus seems to imply; they ate in fact assumed in order to prove VI. 1. Comparison of areas of triangles of I. 34. The theorem already mentioned as given by Proclus on 1. 34 {pp. 340 — 4) is placed here by Heron, who also enunciates it more clearly (an-Nairtzi, ed. Besthom-Heiberg, pp. 155 — 161, ed. Curtze, pp. 75 — 8). If in two triangles two sides of the one be equal to two sides of the other respectively, and the angle of the one be greater than the angle of the other, namely the angles contained by the equal sides, then, (t) if the sum of the tovo angles contained by the equal sides is equal to two right angles, the two triangles are equal to site another ; (2) if less than two right angles, the triangle which has the greater angle is also itself greater than the other; (3) if greater than two right angles, the triangle which has the less angle is greater than the other triangle. D Let two triangles ABC, DEF have the sides AB, AC respectively equal to DE, DF. (1) First, suppose that the angles at A and D in the triangles ABC, DEF axe together equal to two right angles. Heron's construction is now as follows. Make the angle EDG equal to the angle BAC. Draw /KT parallel to ED meeting DG in H. lwa.EH. L 38] PROPOSITION 38 335 Then, since the angles BAC, EDF are equal to two right angles, the angles EDH, ED Pax*, equal to two right angles. But so are the angles EDH, DHF, Therefore the angles EDF, DHF axe equal. And the alternate angles EDF, DFB are equal. [1. 39] Therefore the angles DHF, DFHaxe equal, and DF'v> equal to DH. [1. 6] Hence the two sides ED, DHaxe equal to the two sides BA, AC\ and the included angles are equal. Therefore the triangles ABC, DEH axe equal in all respects. And the triangles DEF, DEH between the same parallels are equal. ['• 37] Therefore the triangles ABC, DEF axe equal. [Proclus takes the construction of Eucl. 1. 24, i.e., he makes DH equal to DF and then proves that ED, Fffare parallel.] (2) Suppose the angles BAC, EDF together less than two right angles. As before, make the angle EDG equal to the angle BAC, draw FH parallel to ED, and join EH. In this case the angles EDH, EDF are together less than two right angles, while the angles EDH, DHF are equal to two right angles. [1. 39] Hence the angle EDF, and therefore the angle DFH, is less than the angle DHF. Therefore DH is less than DF. [1. 19] Produce DH to G so that DGis equal to Z>^or AC, and join EG. Then the triangle DEG, which is equal to the triangle ABC, is greater than the triangle DEH, and therefore greater than the triangle DEF. (3) Suppose the angles BAC, EDF together greater than two right angles. A J? We make the same construction in this case, and we prove in like manner that the angle DHF is less than the angle DFH, whence DH is greater than DFot AC. Make DG equal to A C, and join EG. It then follows that the triangle DEF is greater than the triangle ABC. [In the second and third cases again Proclus starts from the construction in 1. 34, and proves, in the second case, that the parallel, FH, to ED cuts DG and, in the third case, that it cuts DG produced] 3j6 BOOK I [1.38,39 There is no necessity for Heron to take account of the position of F in relation to the side opposite D. For in the first and third cases F must fall in the position in which Euclid draws it in 1. 24, whatever be the relative lengths of A B, AC. In the second case the figure may be as annexed, but the proof is the same, or rather the case needs no proof at all. Proposition 39. Equal triangles which are on the same base and on the same side are also in the same parallels. Let ABC, DBC be equal triangles which are on the same base BC and on the same side of it ; s [I say that they are also in the same parallels.] And [For] let AD be joined ; I say that AD is parallel to BC. For, if not, let AE be drawn through the point A parallel to the straight line 10 BC, [1. 31] and let EC be joined. Therefore the triangle ABC is equal to the triangle EBC; for it is on the same base BC with it and in the same 15 parallels. [1. 37] But ABC is equal to DBC ; therefore DBC is also equal to EBC, [C. N. 1] the greater to the less: which is impossible. Therefore AE is not parallel to BC. «> Similarly we can prove that neither is any other straight line except AD ; therefore AD is parallel to BC. Therefore etc. Q. E. D. I. 39, 40] PROPOSITIONS 38 -40 337 5. [I say that they are also in the same parallels.] Heiberg has proved \Heruies t xxxviji., [993, p. 50} from a recently discovered papyrus-fragment [Fayiim towns and their papyri, p. go, No. IX.) that these words are an interpolation by some one who did not observe that the words "And let AD be joined " are part of the stlting-eut (IxStnt), but took them as belonging to the construction (KaTaSKtirff) and consequently thought that a Sioptrfioi or "definition {of the thing to be proved) should precede* The interpolator then altered 11 And " into " For " in the next sentence. This theorem is of course the partial converse of t. 37. In 1. 37 we have triangles which are (1) on the same base, {2) in the same parallels, and the theorem proves {3) that the triangles are equal. Here the hypothesis (1) and the conclusion (3) are combined as hypotheses, and the conclusion is the hypothesis (1) of 1. 37, that the triangles are in the same parallels. The additional qualification in this proposition that the triangles must be on the same side of the base is necessary because it is not, as in 1. 37, involved in the other hypotheses. Proclus (p. 407, 4 — 17) remarks that Euclid only converts I. 37 and 1. 38 relative to triangles, and omits the converses of 1. 35, 36 about parallelograms as unnecessary because it is easy to see that the method would be the same, and therefore the reader may properly be left to prove them for himself. The proof is, as Proclus points out (p. 408, 5 — 21), equally easy on the supposition that the assumed parallel AE meets BD or CD produced beyond £>. [Proposition 40. Equal triangles which are on equal bases and on the same side are also in the same parallels. Let ABC, CDE be equal triangles on equal bases BC, CE and on the same side. I say that they are also in the same parallels. For let AD be joined ; I say that AD is parallel to BE. ^ For, if not, let AF be drawn through f\ — pp A parallel to BE [1. 31], and let FE be joined. Therefore the triangle ABC is equal b to the triangle FCE ; for they are on equal bases BC, CE and in the same parallels BE, AF. [1. 38] But the triangle ABC is equal to the triangle DCS ; therefore the triangle DCE is also equal to the triangle FCE, [C. N. 1] the greater to the less : which is impossible. Therefore AF is not parallel to BE. 338 BOOK I [i. 40, 41 Similarly we can prove that neither is any other straight line except AD ; therefore AD is parallel to BE. Therefore etc. Q, E. D.] Heiberg has proved by means of the papyrus-fragment mentioned in the last note that this proposition is an interpolation by some one who thought that there should be a proposition following 1. 39 and related to it in the same way as 1. 38 is related to 1. 37, and 1. 36 to t. 35. Proposition 41. If a parallelogram have the same base milk a triangle and be in the same parallels, the parallelogram is double of the triangle. For let the parallelogram ABCD have the same base BC with the triangle BBC, and let it be in the same parallels BC, AE; I say that the parallelogram ABCD is double of the triangle BEC. For let AC be joined. Then the triangle ABC is equal to the triangle EBC ; for it is on the same base BC with it and in the same parallels BC, AE. ['• 37] But the parallelogram ABCD is double of the triangle ABC; for the diameter AC bisects it ; [1. 34] so that the parallelogram ABCD is also double of the triangle EBC, Therefore etc. Q. E. D. On this proposition Proctus {pp. 414, 15 — 415, 16), "by way of practice" (yiyiriKri'iK Iffica), considers the area of a trapezium (a quadrilateral with only one pair of opposite sides parallel) in comparison with that of the triangles in the same parallels and having the greater and less of the parallel sides of the trapezium for bases respectively, and proves that the trapezium is less than double of the former triangle and more than double of the latter. He next (pp. 415, 22—416, 14) proves the proposition that, If a triangle be formed by joining the middle point of either of the non- parallel sides to the extremities of the opposite side, the area of the trapezium is always double of that of the triangle. 1. 4i, 4a] PROPOSITIONS 40—43 339 Let ABCD be a trapezium in which AD, BC are the parallel sides, and E the middle point of one of the non-parallel sides, say DC. Job EA, EB and produce BE to meet AD produced in F. Then the triangles BEC, FED have two angles equal respectively, and one side CE equal to one side DE ; therefore the triangles are equal in atl respects. Add to each the quadrilateral ABED ; therefore the trapezium ABCD is equal to the triangle ASF, that is, to twice the triangle AEB, since BE is equal to EF. [l. 38] The three properties proved by Prod us may be combined in one enuncia- tion thus : If a triangle be formed by joining the middle point of one side of a trapezium to the extremities of the opposite side, the area of the trapezium is {1) greater than, (1) equal to, or (3) less than, double the area of the triangle according as the side the middle point of which is taken is ( 1 ) the greater of the parallel sides, (a) either of the non-parallel sides, or (3) the lesser of the parallel sides. Proposition 42. To construct, in a given rectilineal angle, a parallelogram equal to a given triangle. Let ABC be the given triangle, and D the given recti- lineal angle ; thus it is required to construct in the rectilineal angle D a parallelogram equal to the triangle ABC. Let BC be bisected at E, and let AE be joined ; on the straight line EC, and at the point E on it, let the angle CEF be constructed equal to the angle D ; [1. 23] through A let AG be drawn parallel to EC, and [1. 31] through C let CG be drawn parallel to EF, Then FECG is a parallelogram. And, since BE is equal to EC, the triangle ABE is also equal to the triangle AEC, for they are on equal bases BE, EC and in the same parallels BC,AG; [..38] therefore the triangle ABC is double of the triangle AEC. 340 BOOK I [t.43,43 But the parallelogram FECG is also double of the triangle A EC, for it has the same base with it and is in the same parallels with it ; [1. 41] therefore the parallelogram FECG is equal to the triangle ABC. And it has the angle CEF equal to the given angle D. Therefore the parallelogram FECG has been constructed equal to the given triangle ABC, in the angle CEF which is equal to D. Q. E. F. Proposition 43. In any parallelogram tke complements of the parallelograms about tke diameter are equal to one another. Let ABCD be a parallelogram, and AC its diameter ; and about AC let EH, FG be parallelograms, and BK, KD S the so-called complements ; I say that the complement BK is equal to the complement KD. For, since ABCD is a parallelogram, and AC its diameter, the triangle ABC is equal to to the triangle A CD. [1. 34] Again, since EH is a parallelo- gram, and AK is its diameter, the triangle AEK is equal to the triangle AHK. 15 For the same reason the triangle KFC is also equal to KGC. Now, since the triangle AEK is equal to the triangle AHK, and KFC to KGC, *o the triangle AEK together with KGC is equal to the triangle AHK together with KFC. [C. M 1] And the whole triangle ABC is also equal to the whole ADC; therefore the complement BK which remains is equal to the a j complement KD which remains. [C. N. 3] Therefore etc. Q. E. D. t. 43. 44] PROPOSITIONS 43— 44 341 1. complement*, iraparXi)pv>umi, the figures put in to fill up (interstices). 4. and About AC..., Euclid's phraseology here and in the next proposition implies that the complements as well at the other parallelograms are " about " the diagonal. Toe words are here rtpl Si iiy/ AT wapaWTjXfr'pawui \ikv fara t& E9, ZH, ri Si Xcy&tt&A TapaT\ypJittaTa ra BK, KA. The expression "the so-called complements" indicates that this technical use of xopaiXij^j^aTo was not new, though it might not be universally known. In the text of Proclus' commentary as we have it, the end of the note on i. 41, the whole of that on 1. 4a, and the beginning of that on 1. 43 are missing. Proclus remarks (p. 418, 15 — 30) that Euclid did not need to give a formal definition of complement because the name was simply suggested by the facts; when once we have the two "parallelograms about the diameter," the complements are necessarily the areas remain- ing over on each side of the diameter, which fill up the complete parallelogram. Thus (p. 417, 1 sqq.) the complements need not be parallelo- grams. They are so if the two "parallelograms about the diameter" are formed by straight lines drawn through one point of the diameter parallel to the sides of the original parallelogram, but not otherwise. If, as in the first of the accompanying figures, the parallelograms have no common point, the complements are five-sided figures as shown. When the parallelograms overlap, as in the second figure, Proclus regards the complements as being the small parallelo- grams FG, EH But, if complements are strictly the areas required to fill up the original parallelo- gram, Proclus is inaccurate in describing FG, EH as the complements. The complements are really (1) the parallelogram FG minus the triangle LM N, and (2) the parallelogram EH minus the triangle KMN, respectively; the possibility that the re- spective differences may be negative merely means the possibility that the sum of the two parallelograms about the diameter may be together greater than the original parallelogram. In all the cases it is easy to show, as Proclus does, that the complements are still equal. Proposition 44. To a given straight line to apply, in a given rectilineal angle, a parallelogram equal to a given triangle. Let AB be the given straight line, C the given triangle and D the given rectilineal angle ; 5 thus it is required to apply to the given straight line AB, in an angle equal to the angle D, a parallelogram equal to the given triangle C. Let the parallelogram BEFG be constructed equal to the triangle C, in the angle EBG which is equal to D [1. 42] ; 10 let it be placed so that BE is in a straight line with AB ; let 34* BOOK I [1.44 FG be drawn through to H, and let AH he drawn through A parallel to either BG or EF. [j. 3*] Let HB be joined. Then, since the straight line HF falls upon the parallels is AH, EF, the angles AHF, HFE are equal to two right angles. [t 39] Therefore the angles BHG, GFE are less than two right angles ; and straight lines produced indefinitely from angles less than *> two right angles meet ; [Post. 5] therefore HB, FE, when produced, wilt meet. Let them be produced and meet at K ; through the point K let KL be drawn parallel to either EA or FH, [1. 31] and let HA, GB be produced to the points L, M. n Then HLKF is a parallelogram, HK is its diameter, and AG, ME are parallelograms, and LB, BF the so-called complements, about HK; therefore LB is equal to BF. [1. 43] But BF is equal to the triangle C ; 3° therefore LB is also equal to C. [C. N. 1] And, since the angle GBE is equal to the angle ABM, [1 'Si while the angle GBE is equal to D, the angle ABM is also equal to the angle D. Therefore the parallelogram LB equal to the given triangle 35 C has been applied to the given straight line AB, in the angle ABM which is equal to D. Q. E. F. 14, since the straight line HP falls.... The verb is in the wrist (triitviy) here ftnd in similar expressions in the following propositions. This proposition wilt always remain one of the most impressive in all geometry when account is taken (i) of the great importance of the result i. 44] PROPOSITION 44 343 obtained, the transformation of a parallelogram of any shape into another with the same angle and of equal area but with one side of any given length, e.g. a unit length, and (a) of the simplicity of the means employed, namely the mere application of the property that the complements of the "parallelograms about the diameter" of a parallelogram are equal. The marvellous ingenuity of the solution is indeed worthy of the " godlike men of old," as Prod us calls the discoverers of the method of " application of areas"; and there would seem to be no reason to doubt that the particular solution, like the whole theory, was Pythagorean, and not a new solution due to Euclid himself. Application of areas. On this proposition Proclus gives (pp. 419, 15 — 420, 23) a valuable note on the method of " application of areas " here introduced, which was one of the most powerful methods on which Greek geometry relied. The note runs as follows : " These things, says Eudemus (ol irrpl rov Ev&wwe), are ancient and are discoveries of the Muse of the Pythagoreans, I mean the application of anas (rapaflokij twi> x M P''' 8, ')i their exceeding (inrtpfjoXij) and their falling-short (£AA«^«). It was from the Pythagoreans that later geometers [i.e. Apollonius] took the names, which they again transferred to the so-called conk lines, designating one of these a parabola (application), another a hyperbola (exceeding) and another an ellipse (falling-short), whereas those godlike men of old saw the things signified by these names in the construction, in a plane, of areas upon a finite straight line. For, when you have a straight line set out and lay the given area exactly alongside the whole of the straight line, then they say that you apply (irupailaXkur) the said area; when however you make the length of the area greater than the straight line itself, it is said to exceed (vrcpfiukkuv), and when you make it less, in which case, after the area has been drawn, there is some part of the straight line extending beyond it, it is said to fall short (MAcia-w). Euclid too, in the sixth book, speaks in this way both of exceeding and falling-short ; but in this place he needed the application simply, as he sought to apply to a given straight line an area equal to a given triangle in order that we might have in our power, not only the construction (crwmurw) of a parallelogram equal to a given triangle, but also the application of it to a finite straight line. For example, given a triangle with an area of 12 feet, and a straight line set out the length of which is 4 feet, we apply to the straight line the area equal 10 the triangle if we take the whole length of 4 feet and find how many feet the breadth must be in order that the parallelogram may be equal to the triangle. In the particular case, if we find a breadth of 3 feet and multiply the length into the breadth, supposing that the angle set out is a right angle, we shall have the area. Such then is the application handed down from early times by the Pythagoreans." Other passages to a similar effect are quoted from Plutarch, (1) "Pytha- goras sacrificed an ox on the strength of his proposition (Siaypa/i/u) as Apollodotus (?-rus) says... whether it was the theorem of the hypotenuse, viz. that the square on it is equal to the squares on the sides containing the right angle, or the problem about the application of an area" (JVon posse suauiter vivi secundum Epieurum, c, n.) (2) "Among the most geometrical theorems, or rather problems, is the following : given two figures; to apply a third equal to the one and similar to the other, on the strength of which discovery they say moreover that Pythagoras sacrificed. This is indeed unquestionably more subtle and more scientific than the theorem which 344 BOOK I [L44 demonstrated that the square on the hypotenuse is equal to the squares on the sides about the right angle " {Symp. vm. 2,4). The story of the sacrifice must (as noted by Bretschneider and Hankel) be given up as inconsistent with Pythagorean ritual, which forbade such sacrifices ; but there is no reason to doubt that the first distinct formulation and introduction into Greek geometry of the method of application of areas was due to the Pythagoreans. The complete exposition of the application of areas, their exceeding and their falling-short, and of the construction of a rectilineal figure equal to one given figure and similar to another, takes us into the sixth Book of Euclid ; but it will be convenient to note here the general features of the theory of application, exceeding and falling-short. The simple application of a parallelogram of given area to a given straight line as one of its sides is what we have in 1. 44 and 45 ; the general form of the problem with regard to exceeding and falling-short may be stated thus: "To apply to a given straight line a rectangle (or, more generally, a parallelogram) equal to a given rectilineal figure and (i) exceeding or (2) falling-short by a square {or, in the more general case, a parallelogram similar to a given parallelogram)." What is meant by saying that the applied parallelogram (1) exceeds or (2) falls short is that, while its base coincides and is coterminous at one end with the straight line, the said base (1) overlaps or (2) falls short of the straight line at the other end, and the portion by which the applied parallelogram exceeds a parallelogram of the same angle and height on the given straight line (exactly) as base is a parallelogram similar to a given parallelogram (or, in particular cases, a square). In the case where the parallelogram is to fall short, a Siofmr/to's is necessary to express the condition of possibility of solution. We shall have occasion to see, when we come to the relative propositions in the second and sixth Books, that the general problem here stated is equivalent to that of solving geometrically a mixed quadratic equation. We shall see that, even by means of 11. 5 and 6, we can solve geometrically the equations ax± x? = P, &—ax — P ; but in vi. 28, 29 Euclid gives the equivalent of the solution of the general equations * _, C — e m We are now in a position to understand the application of the terms parabola (application), hyperbola (exceeding) and ellipse (falling-short) to conic sections. These names were first so applied by A polio ni us as expressing in each case the fundamental property of the curves as stated by him. This fundamental property is the geometrical equivalent of the Cartesian equation referred to any diameter of the conic and the tangent at its extremity as (in general, oblique) axes. If the parameter of the ordinates from the several points of the conic drawn to the given diameter be denoted by p (p being d" 1 accordingly, in the case of the hyperbola and ellipse, equal to -3 , where d is the length of the given diameter and d' that of its conjugate), Apollonius gives the properties of the three conies in the following form. i. 44, 45] PROPOSITIONS 44> 45 345 (1) For the parabola, the square on the ordinate at any point is equal to a rectangle applied to / as base with altitude equal to the corresponding abscissa. That is to say, with the usual notation, (3) For the hyperbola and ellipse, the square on the ordinate is equal to the rectangle applied to p having as its width the abscissa and exceeding (for the hyperbola) or falling-short (for the ellipse) by a figure similar and similarly situated to the rectangle contained by the given diameter and p. That is, in the hyperbola jr =px + -3 pd, or y* -px + ^ *" i and in the ellipse y'=px — -.x*. d The form of these equations will be seen to be exactly the same as that of the general equations above given, and thus Apollonius' nomenclature followed exactly the traditional theory of application, exceeding, a.nA.faliing*short. Proposition 45. To construct, in a given rectilineal angle, a parallelogram equal to a given rectilineal figure. Let ABCD be the given rectilineal figure and E the given rectilineal angle ; 5 thus it is required to construct, in the given angle E, a parallelogram equal to the rectilineal figure ABCD. - - ■ Let DB be joined, and let the parallelogram FH be constructed equal to the triangle ABD, in the angle HKF which is equal to E ; [1. 4a] 10 let the parallelogram GM equal to the triangle DBC be applied to the straight line GH, in the angle GHM which is equal to /:'. [1. 44] Then, since the angle E is equal to each of the angles HKF, GHM, i S the angle HKF is also equal to the angle GHM. [C. N. 1] 34« BOOK I [L45 Let the angle KHG be added to each ; therefore the angles FKH, KHG are equal to the angles KHG, GHM, But the angles FKH, KHG are equal to two right angles ; ft. 3 9 ] *> therefore the angles KHG, GHM are also equal to two right angles. Thus, with a straight line GH, and at the point H on it, two straight lines KH, HM not lying on the same side make the adjacent angles equal to two right angles ; as therefore KH is in a straight line with HM. [i. 14] And, since the straight line HG falls upon the parallels KM, FG, the alternate angles MHG, HGF are equal to one another. [1, 39] Let the angle HGL be added to each ; jo therefore the angles MHG, HGL are equal to the angles HGF, HGL. [C N. 2] But the angies MHG, HGL are equal to two right angles ; % *9] therefore the angles HGF, HGL are also equal to two right angles. [C. N. 1] 35 Therefore FG-is in a straight line with GL. [t. 14] And, since FK is equal and parallel to HG, [1. 34] and HG to ML also, KF is also equal and parallel to ML ; [C. N. 1 ; 1. 30] and the straight lines KM, FL join them (at their extremities); 40 therefore KM, FL are also equal and parallel. [1. 33] Therefore KFLM is a parallelogram. And, since the triangle ABD is equal to the parallelogram FM > and DBC to GM, 45 the whole rectilineal figure ABCD is equal to the whole parallelogram KFLM. Therefore the parallelogram KFLM has been constructed equal to the given rectilineal figure ABCD, in the angle FKM which is equal to the given angle E. q. e. f. ?, 3, £, 45. 48- rectilineal figure, in the Greek "rectilineal" simply, without "figure," fitit-ypaix&o* being here used as a substantive, like the similarly formed rapdXXifi^ypafitttim. Transformation of areas. We can now take stock of how far the propositions 1. 43 — 45 bring us in the matter of transformation 0/ areas, which constitutes so important a part of i. «, 46] PROPOSITIONS 4 5> 46 347 what has been fitly called the geometrical algebra of the Greeks. We have now learnt how to represent any rectilineal area, which can of course be resolved into triangles, by a single parallelogram having one side equal to any given straight line and one angle equal to any given rectilineal angle. Most important of all such parallelograms is the rectangle, which is one of the simplest forms in which an area can be shown. Since a rectangle corresponds to the product of two magnitudes in algebra, we see that application to a given straight line of a rectangle equal to a given area is the geometrical equivalent of algebraical division of the product of two quantities by a third. Further than this, it enables us to add or subtract any rectilineal areas and to represent the sum or difference by one rectangle with one side of any given length, the process being the equivalent of obtaining a common factor. But one step still remains, the finding of a square equal to a given rectangle, i.e. to a given rectilineal figure; and this step is not taken till 11. 14. In general, the transformation of combinations of rectangles and squares into other combinations of rectangles and squares is the subject-matter of Book 11., with the exception of the expression of the sum of two squares as a single square which appears earlier in the other Pythagorean theorem 1. 47. Thus the transformation of rectilineal areas is made complete in one direction, i.e. in the direction of their simplest expression in terms of rectangles and squares, by the end of Book 11. The reverse process of transforming the simpler rectangular area into an equal area which shall be similar to any rectilineal figure requires, of course, the use of proportions, and therefore does not appear till vi. 25. Proclus adds to his note on this proposition the remark (pp. 411, 24 — 413, 6) : "I conceive that it was in consequence of this problem that the ancient geometers were ted to investigate the squaring of the circle as well. For, if a parallelogram can be found equal to any rectilineal figure, it is worth inquiring whether it be not also possibte to prove rectilineal figures equal to circular. And Archimedes actually proved that any circle is equal to the right-angled triangle which has one of its sides about the right angle [the perpendicular] equal to the radius of the circle and its base equal to the pen meter of the circle. But of this elsewhere." Proposition 46. On a given straight line to describe a square. Let AB be the given straight line ; thus it is required to describe a square on the straight line AB. 5 Let AC be drawn at right angles to the straight line AB from the point A on it [j. 11], and let AD be made equal through the point D let DE be drawn 10 parallel to AB, and through the point B let BE be drawn parallel to AD. [1. 31] 348 BOOK I [1.46 Therefore ADEB is a parallelogram ; therefore AB is equal to DE, and AD to BE. [1. 34] But AB is equal to AD ; •5 therefore the four straight lines BA, AD, DE, EB are equal to one another ; therefore the parallelogram ADEB is equilateral. I say next that it is also right-angled. For, since the straight line AD falls upon the parallels *>AB, DE, the angles BAD, ADE are equal to two right angles. [1. a 9 ] But the angle BAD is right ; therefore the angle ADE is also right. And in parallelogrammic areas the opposite sides and as angles are equal to one another ; [1. 34] therefore each of the opposite angles ABE, BED is also right Therefore ADEB is right-angled. And it was also proved equilateral. 30 Therefore it is a square ; and it is described on the straight line AB. Q. E. F. i (p. 4 13, iSsqq.) notes the difference between the ward constntet I by Euclid to the construction of a triangle (and! he might have added, I, 3, 30. Proclus {ffveripatdiu) ap of an angle) and the words describe en {dvtypiipt .r iri) used of drawing » square on a given straight line as one side. The triangle (or angk) is, so to say, pieced together, while the describing of a square on a given straight line is the making of a figure " from " one side, and corresponds to the multiplication of the number representing the side by itself. Proclus (pp. 424 — s) proves that, if squares are described on equal straight lines, the squares are equal; and, conversely^ that, if two squares are equal, the straight lines are equal on which they are described. The first proposition is immediately obvious if we divide tbe squares into two triangles by drawing a diagonal in each. The converse is proved as follows. Place the two equal squares AF, CG so that AB, BC are in a straight line. Then, since the angles are right, FB, BG will also be in a straight line. Join AF, FC, CG, GA. Now, since the squares are equal, the triangles ABF, CBG are equal. Add to each the triangle FBC ; therefore the triangles AFC, GFC are equal, and hence they must be in the same parallels. 1. 46, 47] PROPOSITIONS 46, 4J 349 Therefore AG, CFaie parallel. Also, since each of the alternate angles AFG, FGC is half a right angle, AF, CG are parallel. Hence AFCG is a parallelogram ; and AF, CG are equal. Thus the triangles ABF, CBG have two angles and one side respectively equal; therefore AB is equal to BC, and BF to BG. Proposition 47. In right-angled triangles the square on the side subtending the right angle is equal to the squares on the sides containing the right angle. Let ABC be a right-angled triangle having the angle %BAC right; I say that the square on BC is equal to the squares on BA, AC. For let there be described on BC the square BDEC, 10 and on BA, AC the squares GB,HC\ [1.46] through A let AL be drawn parallel to either BD or CE, and let AD, FC be joined. is Then, since each of the angles BAC, BAG is right, it follows that with a straight line BA, and at the point A on it, the two straight lines 20 AC, AG not lying on the same side make the adjacent angles equal to two right angles; therefore CA is in a straight line with AG. For the same reason BA is also in a straight line with AH. And, since the angle DBC is equal to the angle FBA : for each is right : let the angle ABC be added to each ; jo therefore the whole angle DBA is equal to the whole angle FBC. [C N. *] [.. .4] 15 35° BOOK I [i. 47 And, since DB is equal to BC, and FB to BA, the two sides AB, BD are equal to the two sides FB, BC respectively , 35 and the angle ABD is equal to the angle FBC ; therefore the base AD is equal to the base FC, and the triangle ABD is equal to the triangle FBC. [i. 4] Now the parallelogram BL is double of the triangle ABD, for they have the same base BD and are in the same parallels 4° BD, AL, [1. 4 i] And the square GB is double of the triangle FBC, for they again have the same base FB and are in the same parallels FB, GC. [1. 41] [But the doubles of equals are equal to one another,] 45 Therefore the parallelogram BL is also equal to the square GB. Similarly, if AE, BK be joined, the parallelogram CL can also be proved equal to the square HC; jo therefore the whole square BDEC is equal to the two squares GB, HC. [C. N. a] And the square BDEC is described on BC, and the squares GB, HC on BA, AC. Therefore the square on the side BC is equal to the 55 squares on the sides BA, AC. Therefore etc. Q. E. D. 1. the square on, rb farb...TtTp&y' =(!«■ 8*+ 6" =io' l6 J + 12* = 20 1 . It will be seen that 4 1 + 3* = 5* can be derived from each of these by multiplying, or dividing out, by one and the same factor. We may therefore admit that the Egyptians knew that 3* + 4*= 5'. But there seems to be no evidence that they knew that the triangle {3, 4, 5) is right-angled ; indeed, according to the latest authority {T. Eric Peet, The Rhind Mathematical Papyrus, 1923)1 nothing in Egyptian mathematics suggests that the Egyptians were acquainted with this or any special cases of the Pythagorean theorem. How then did Pythagoras discover the general theorem ? Observing that 3, 4, 5 was a right-angled triangle, while 3 + 4 1 = $', he was probably led to consider whether a similar relation was true of the sides of right-angled triangles other than the particular one. The simplest case (geometrically) to investigate was that of the isosceles right-angled triangle ; and the truth of the theorem in this particular case would easily appear from the mere construction of a figure. Cantor {i„ p. 185) and Allman {Greek Geometry from Thales to Euclid, p, 29) illustrate by a figure in which the squares are drawn outwards, as in 1. 47, and divided by diagonals into equal triangles ; but I think that the truth was morejikely to be^first observed from a figure of the kind suggested by Biirk (Das ApastambaSulba-Sutra in Zeilschrift der deuts ken morgentand. Gesellschaft, lv., 1901, p. 557) to explain how the Indians arrived at the same thing. The two figures are as shown above. When the geometrical 1.47] PROPOSITION 47 353 consideration of the figure had shown that the isosceles right-angled triangle had the property in question, the investigation of the same fact from the arithmetical point of view would ultimately lead to the other momentous discovery of the irrationality of the length of the diagonal of a square expressed in terms of its side. The irrational will come up for discussion later ; and our next question is : Assuming that Pythagoras had observed the geometrical truth of the theorem in the case of the two particular triangles, and doubtless of Other rational right-angled triangles, how did he establish it generally ? There is no positive evidence on this point. Two possible lines are however marked out. (1) Tannery says (La Gtomitrie greegue, p. 105) that the geometry of Pythagoras was sufficiently advanced to make it possible for him to prove the theorem by similar triangles. He does not say in what particular manner similar triangles would be used, but their use must apparently have involved the use of proportions, and, in order that the proof should be conclusive, of the theory of proportions in its complete form applicable to incommensurable as well as commensurable magnitudes. Now Eudoxus was the first to make the theory of proportion independent of the hypothesis of commensurability ; and as, before Eudoxus' time, this had not been done, any proof of the general theorem by means of proportions given by Pythagoras must at least have been inconclusive. But this does not constitute any objection to the supposition that the truth of the general theorem may have been discovered in such a manner; on the contrary, the supposition that Pythagoras proved it by means of an imperfect theory of proportions would better than anything else account for the fact that Euclid had to devise an entirely new proof, as Proclus says he did in 1. 47. This proof had to be independent of the theory of proportion even in its rigorous form, because the plan of the Elements postponed that theory to Books v. and vi. , while the Pythagorean theorem was required as early as Book U. On the other hand, if the Pythagorean proof had been based on the doctrine of Books 1. and 11. only, it would scarcely have been necessary for Euclid to supply a new proof. The possible proofs by means of proportion would seem to be practically limited to two. (a) One method is to prove, from the similarity of the triangles ABC, DBA, that the rectangle CB, BD is equal to the square on BA, and, from the similarity of the triangles ABC, DAC, that the rectangle BC, CD is equal to the square on CA ; whence the result follows by addition. It will be observed that this proof is in substance identical with that of Euclid, the only difference being that the equality of the two smaller squares to the respective rectangles is inferred by the method of Book vi. instead of from the relation between the areas of parallelograms and triangles on the same base and between the same parallels established in Book 1. It occurred to me whether, if Pythagoras' proof had come, even in substance, so near to Euclid's, Proclus would have emphasised so much as he does the originality of Euclid's, or would have gone so far as to say that he marvelled more at that proof than at the original discovery of the theorem. But on the whole I see no difficulty ; for there can be little doubt that the proof by proportion is what suggested to Euclid the method of 1. 47, and the transformation of 3S4 BOOK I L-.47 the method of proportions into one based on Book i. only, effected by a construction and proof so extraordinarily ingenious, is a veritable tour de font which compels admiration, notwithstanding the ignorant strictures of Schopenhauer, who wanted something as obvious as the second figure in the case of the isosceles right-angled triangle (p. 353), and accordingly (Siimmtlithr Wtrkt, 111. § 39 and 1. § 15) calls Euclid's proof "a mouse-trap proof" and "a proof walking on stilts, nay, a mean, underhand, proof" ("Des Eukleides stelzbeiniger, ja, hi nterlis tiger Beweis "). {b) The other possible method is this. As it would be seen that the triangles into which the original triangle is divided by the perpendicular from the right angle on the hypotenuse are similar to one another and to the whole triangle, while in these three triangles the two sides about the right angle in the original triangle, and the hypotenuse of the original triangle, are corresponding sides, and that the sum of the two former similar triangles is identically equal to the simitar triangle on the hypotenuse, it might be inferred that the same would also be true of squares described on the corresponding three sides respectively, because squares as well as similar triangles are to one another in the duplicate ratio of corresponding sides. But the same thing is equally true of any similar rectilineal figures, so that this proof would practically establish the extended theorem of Eucl. vi. 31, which theorem, however, Proclus appears to regard as being entirely Euclid's discovery. On the whole, the most probable supposition seems to me to be that Pythagoras used the first method (a) of proof by means of the theory of proportion as he knew it, i.e. in the defective form which was in use up to the date of Eudoxus. (1) I have pointed out the difficulty in the way of the supposition that Pythagoras' proof depended upon the principles of Eucl. Books 1. and 11. only. ,,„ I , » , 5 ' A \ \ \ \ \ Were it not for this difficulty, the conjecture of Bretschneider (p. 81), followed by Hankel (p. 98), would be the most tempting hypothesis. According to this suggestion, we are to suppose a figure like that of Eucl. it. 4 in which a, b are the sides of the two inner squares respectively, and a + b is the side of the complete square. Then, if the two complements, which are equal, are divided by their two diagonals into four equal triangles of sides a, b, t, we can place these triangles round another square of the same size as the whole square, in the manner shown in the second figure, so that the sides a, b of sutmsnx triangles make up one of the sides of the square and are arranged in cyclic order. It readily follows that the remainder of the square when the four triangles are deducted is, in the one case, a square whose side is e, and in the other the sum of two squares whose sides are a, b respectively. Therefore the.square on t is equal 47] PROPOSITION 47 355 to the sum of the squares on a, b. All that can be said against this con- jectural proof is that it has no specifically Greek colouring but rather recalls the Indian method. Thus Bhaskara (bom 1 1 14 a.d. ; see Cantor, 1,, p. 656) simply draws four right-angled triangles equal to the original one in- wards, one on each side of the squaie on the hypotenuse, and says " see ! ", without even adding that inspection shows that c> = ^ + {a-i,y-- a> + o*. Though, for the reason given, there is difficulty in supposing that Pythagoras used a general proof of this kind, which applies of course to right- angled triangles with sides incommensurable as well as commensurable, there is no objection, I think, to supposing that the truth of the proposition in the case of the first rational right-angled triangles discovered, e.g. 3, 4, 5, was proved by a method of this sort. Where the sides are commensurable in this way, the squares can be divided up into small (unit) squares, which would much facilitate the comparison between them. That this subdivision was in fact resorted to in adding and subtracting squares is made probable by Aristotle's allusion to odd numbers as gnomons placed round unity to form successive squares in Physics lit. 4 ; this must mean that the squares were represented by dots arranged in the form of a square and a gnomon formed of dots put round, or that (if the given square was drawn in the usual way) the gnomon was divided up into unit squares. Zeuthen has shown (" Thiorime de Pythagore" Origine de la Giomctrie scUntifique in Comptes rendus du //'" Congres international de Philosophic, Geneve, 1904), how easily the proposition could be proved by a method of this kind for the triangle 3, 4, 5. To admit of the two smaller squares being shown side by side, take a square on a line containing 7 units of length (4 + 3), and divide it up into 49 small squares. It would be obvious that the whole square could be exhibited as containing four rectangles of sides 4, 3 cyclically arranged round the figure with one unit sqi^re in the middle. (This same figure is given by Cantor, t„ p. 680. to illustrate the method given in. the Chinese " Ch6u-pel ",) It would be seen that (i) the. whole square (7*) is made up of two squares 3* and 4*, and two rectangles 3, 4; (ii) the same square is made up of the square EFGH and the halves of four of the same rect- angles 3t 4) whence the square EFGH, being equal to the sum of the squares 3* and 4', must contain 25 unit squares and its side, or the diagonal of one of the rectangles, must contain 5 units of length. Or the result might equally be seen by observing that (i) the square EFGH on the diagonal of one of the rectangles is made up of the halves of four rectangles and the unit square in the middle, while (ii) the squares 3* and 4* placed at adjacent comers of the large square make up two rectangles 3, 4 with the unit square in the middle. The procedure would be equally easy for any rational right-angled triangle, and would be a natural method of trying to prove the property when it had ^ -\ 1 7* \ / \ \ v \ r* ?' A '/ V -V / -c 356 BOOK I [i. 47 once been empirically observed that triangles like 3, 4, 5 did in fact contain a right angle. Zeuthen has, in the same paper, shown in a most ingenious way how the property of the triangle 3, 4, 5 could be verified by a sort of combination of the second possible method by similar triangles, (t) on p. 354 above, with subdivision of rectangles into simitar smalt rectangles. 1 give the method on account of its interest, although it is no doubt too advanced to have been used by those who first proved the property of the particular triangle. Let ABC be a triangle right-angled at A, and such that the lengths of the sides A B, A C are 4 and 3 units respectively. Draw the perpendicular AD, divide up AB, AC into unit lengths, complete the rectangle on B C as base and with AD as altitude, and subdivide this rectangle into small rectangles by drawing parallels to BC, AD through the points of division of AB, AC. Now, since the diagonals of the small rectangles are all equal, each being of unit length, it follows by similar triangles that the small rectangles are all equal. And the rectangle with AB for diagonal contains ifi of the small rectangles, while the rectangle with diagonal A C contains 9 of them. But the sum of the triangles ABD, ADC is equal to the triangle ABC. Hence the rectangle with BC as diagonal contains 9+16 or 15 of the small rectangles ; and therefore BC- 5. Rational right-angled triangles from the arithmetical stand- point. Pythagoras investigated the arithmetical problem of finding rational numbers which could be made the sides of right-angled triangles, or of finding square numbers which are the sum of two squares ; and herein we find the beginning of the indeterminate analysis which reached so high a stage of development in Diophantus. Fortunately Prod us has preserved Pythagoras' method of solution in the following passage (pp. 418, 7 — 419, 8). "Certain methods for the discovery of triangles of this kind are handed down, one of which they refer to Plato, and another to Pythagoras. [The latter] starts from odd numbers. For it makes the odd number the smaller of the sides about the right angle; then it takes the square of it, subtracts unity, and makes half the difference the greater of the sides about the right angle; lastly it adds unjty to this and so forms the remaining side, the hypotenuse. For example, taking 3, squaring it, and subtracting unity from the 9, the method takes naif of the 8, namely 4 ; then, adding unity to it again, it makes 5, and a right- angled triangle has been found with one side 3, another 4 and another 5. But the method of Plato argues from even numbers. For it takes the given even number and makes it one of the sides about the right angle ; then, bisecting this number and squaring the half, it add si unity to the square to form the hypotenuse, and subtracts unity from the square to form the other side about the right angle. For example, taking 4, the method squares half of this, or 2, and so makes 4 ; then, subtracting unity; it produces 3, and adding unity it produces 5. Thus it has formed the same triangle as that which was obtained by the other method." 47] PROPOSITION 47 357 The formula of Pythagoras amounts, if fn be an odd number, to the sides of the right-angled triangle being m, — , . Cantor 2 2 (ii, pp. 185—6), taking up an idea of Roth (Geschuhte der abendldndisehen Phitosophie, 11. 517), gives the following as a possible explanation of the way in which Pythagoras arrived at his formula. It £ - a* + 1?, it follows that d> = t'-b* = {e + b)(e-b). Numbers can be found satisfying the first equation if (i) c + b and c — b are either both even or both odd, and if further (2) e + b and t — b are such numbers as, when multiplied together, produce a square number. The first condition is necessary because, in order that e and b may both be whole numbers, the sum and difference of c + b and c — b must both be even. The second condition is satisfied if e + b and e—b are what were called similar numbers (o/i« + pfcag y = ( "ffw y But I think a serious, and even fatal, objection to the conjecture of Cantor and Roth is the very fact that the method enables both the Pythagorean and the Platonic series of triangles to be deduced with equal ease. If this had been the case with the method used by Pythagoras, it would not, I think, hate been left to Plato to discover the second series of such triangles. It seems to me therefore that Pythagoras must have used some method which would produce his rule only; and further it would be some less recondite method, suggested by direct obstrvation rather than by argument from general principles. One solution satisfying these conditions is that of Bretschneider (p. 83), who suggests the following simple method Pythagoras was certainly aware that the successive odd numbers are gnomons, or the differences between successive square numbers. It was then a simple matter to write down in three rows (a) the natural numbers, (/>) their squares, (V) the successive odd numbers constituting the differences between the successive squares in (b), thus: r 2 3 4 5 6 7 8 9 10 n 12 13 14 1 4 9 16 35 36 49 64 81 100 121 144 169 196 1357 9 11 13 15 17 19 21 *3 25 27 Pythagoras had then only to pick out the numbers in the third row which are squares, and his rule would be obtained by finding the formula connecting the square in the third line with the two adjacent squares in the second line. But even this would require some little argument ; and I think a still better suggestion, because making pure observation play a greater part, is that of P. Treutlein {Zeiisckrift fiir Mathematik and Physik, xxviil, 1883, Hist.-litt. Abtheilung, pp. 209 sqq.). We have the best evidence (e.g. in Theon of Smyrna) of the practice of representing square numbers and other figured numbers, e.g. oblong, triangular, hexagonal, by dots or signs arranged in the shape of the particular figure. (Cf. Aristotle, Mttaph. 1092 b 12). Thus, says Treutlein, it would be easily seen that any square number can be turned into the next higher square by putting a single row of dots round two adjacent sides, in the form of a gnomon (see figures on next page). If a is the side of a particular square, the gnomon round it is shown by simple inspection to contain 2a + 1 dots or units. Now, in order that za + 1 may itself be a square, let us suppose a» + 1 = n\ whence a = J (« a - 1 ), and a + 1 = § (n* + 1). '•47] PROPOSITION 47 359 In order that a and a + t may be integral, n must be odd, and we have at once the Pythagorean formula I think Treutlein's hypothesis is shown to be the conect one by the passuge in Aristotle's Physics already quoted, where the reference is undoubtedly to the Pythagoreans, and odd numbers are clearly identified with gnomons " placed round i." But the ancient commentaries on the passage make the matter clearer still. Philoponus says: "As a proof... the Pythagoreans refer to what 13 3 M3 happens with the addition of numbers ; for when the odd numbers are successively added to a square number they keep it square and equilateral.... Odd numbers are accordingly called gnomons because, when added to what are already squares, they preserve the square form,... Alexander has excellently said in explanation that the phrase ' when gnomons are placed round ' means making a figure with the odd numbers (rip/ xara tow s-tptrroin ap^funn U 3«° BOOK I [i. 47 whence a-n'-i, a + 2 = » a + i, and we have the Platonic formula (2«) , +(n , -i)'= (y+i)*. I think this is, in substance, the right explanation, but, in form, not quite correct The Greeks would not, I think, have treated the double row as a gnomon. Their com- parison would have been between (i) a certain ....... square plus a single-row gnomon and (2) the same ■ * square minus a single-row gnomon. As the III**"" application of Eucl. It 4 to the case where the ...tit. segments of the side of the square are a, 1 enables the Pythagorean formula to be obtained as ....... Treutlein obtains it, so I think that Eucl. II. 8 * * - ) * * * * confirms the idea that the Platonic formula was ~. obtained by comparing a square plus a gnomon with the same square minus a gnomon. For n. 8 proves that 4ab + (a-b)'~(a + ty, whence, substituting 1 for b, we have 4a + (a- I )' = (a+ i)«, and we have only to put a — n* to obtain Plato's formula. The "theorem of Pythagoras" in India. This question has been discussed anew in the last few years as the result of the publication of two important papers by Albert Biirk on Das Apastamba- Sulba-Sutra in the Zeitsehrift der deutsehen morgenldndischtn Gesellschaft (lv., 190 r, pp. 543— 591, and lvi., 190a, pp. 337 — 391), The first of the two papers contains the introduction and the text, the second the translation with notes. A selection of the most important parts of the material was made and issued by G. Thibaut in the Journal of the Asiatic Society of Bengal, xliv., 1875, Part 1. (reprinted also at Calcutta, 1875, as Tke Sulvasutras, by G. Thibaut). Thibaut in this work gave a most valuable comparison of extracts from the three Sulvasutras by Baudhayana, Apastamba and Katyayana respectively, with a running commentary and an estimate of the date and originality of the geometry of the Indians. Biirk has however done good service by making the Apasta.mba-S.-S. accessible in its entirety and investigating the whole subject afresh. With the natural enthusiasm of an editor for the work he is editing, he roundly maintains, not only that the Pythagorean theorem was known and proved in all its generality by the Indians long before the date of Pythagoras (about 580 — 500 n.c.), but that they had also discovered the irrational; and further that, so far from Indian geometry being indebted to the Greek, the much-travelled Pythagoras probably obtained his theory from India (loe. eit. LV., p. 575 note). Three impor- tant notices and criticisms of Biirk's work have followed, by H. G. Zeuthen (" Thioreme de lythagore," Origine de la Giomitrie scientifique, 1904, already quoted), by Moritz Cantor ( Uber die dJteste indische Mathematik in the Arehiv der Mathematik und Physik, vin., 1905, pp. 63 — 72) and by Heinrich Vogt {Haben die alien Inder den Pythagoreisfhm Lehrsatx und das Irrational* gekanntt in the Bibliotheca Mathematiea, vn„ 1906, pp. 6 — 23. See also Cantor's Gesehkhtt der Mathematik, i s pp. 635—645. i. 47] PROPOSITION 47 361 The general effect of the criticisms is, I think, to show the necessity for the greatest caution, to say the least, in accepting Biirk's conclusions. I proceed to give a short summary of the portions of the contents of the Apastamba-S.-S. which are important in the present connexion. It may be premised that the general object of the book is to show how to construct altars of certain shapes, and to vary the dimensions of altars without altering the form. It is a collection of rules for carrying out certain constructions. There are no proofs, the nearest approach to a proof being in the rule for obtaining the area of an isosceles trapezium, which is done by drawing a perpendicular from one extremity of the smaller of the two parallel sides to the greater, and then taking away the triangle so cut off and placing it, the other side up, adjacent to the other equal side of the trapezium, thereby transforming the trapezium info a rectangle. It should also be observed that Apastamba does not speak of right-angled triangles, but of two adjacent sides and the diagonal of a rectangle. For brevity, I shall use the expression " rational rectangle " to denote a rectangle the two sides and the diagonal of which can be expressed in terms of rational numbers. The references in brackets are to the chapters and numbers of Apastamba's work. (1) Constructions of right angles by means of cords of the following relative lengths respectively : 3, 4, S ('■ 3, v. , n, 16, 30 (v-3) '5, *«» *5 (v. 3) 5. ". »3 (v. 4) >S. 36. 39 (1. a, v. : !, 4) 8, IS. *7 (v-S) I*. 35. 37 (v-S) (2) A general enunciation of the Pythagorean theorem thus: "The diagonal of a rectangle produces [i.e. the square on the diagonal is equal to] the sum of what the longer and shorter sides separately produce (i.e. the squares on the two sides]." (t. 4} (3) The application of the Pythagorean theorem to a square instead of a rectangle [i.e. to an Uoseeles right-angled triangle] : "The diagonal of a square produces an area double [of the original square]." (1. 5) {4) An approximation to the value of s /a ; the diagonal of a square is (1 + - + ) times the side (1. 6) 3 3-4 3-4-34/ v ' (5) Application of this approximate value to the construction of a square with side of any length. (n. 1) (6) The construction of a ^3, by means of the Pythagorean theorem, as the diagonal of a rectangle with sides a and a J2. (11. 3) (7) Remarks equivalent to the following : (a) a */f is the side of £ (a ^3)*, or a J$ = $a ^3. (11. 3) {*} A square on length of 1 unit gives 1 unit square (lit. 4) „ „ % units gives 4 unit squares (in. 6) » » 3 .. 9 .. (<"• *) i§ „ *i ., ("I- 8) D C H Q K e F A B 362 BOOK I [i. 47 A square, on length of 2 J units gives 6} unit squares (it I. 8) „ „ i unit gives | unit square (in. io) 1 » h » ('■'■ "°) (c) Generally, the square on any length contains as many rows (of small, unit, squares) as the length contains units. (m. 7) (8) Constructions, by means of the Pythagorean theorem, of (a) the sum of two squares as one square, (11. 4) (b) the diffcraia of two squares as one square. (11- 5) {9) A transformation of a rectangle into a square. (tl. 7) [This is not directly done as by Euclid in 11. 14, but the rectangle is first transformed into a gnomon, i.e. into the difference between two squares, which difference is then trans- formed into one square by the preceding rule. If A BCD be the given rectangle of which BC is the longer side, cut off the square ABEF, bisect the rectangle DE left over by HG parallel to FE, move the upper half DG and place it on AF as base in the position AK. Then the rectangle ABCD is equal to the gnomon which is the difference between the square LB and the square LF. In other words, Apastamba transforms the rectangle ab into the difference between the squares f } and f ) .] (io) An attempt at a transformation of a square (a*) into a rectangle which shall have one side of given length (b). (in. t) [This shows no sign of such a procedure as that of Eucl. i. 44, and indeed does no more than say that we must subtract ab from a" and then adapt the remainder c? - ah so that it may "fit on " to the rectangle ab. The problem is therefore only reduced to another of the same kind, and presumably it was only solved arithmttimlly in the case where a, b are given numerically. The Indian was therefore far from the general, geometrical, solution.] (11) Increase of a given square into a larger square. (111.9) [This amounts to saying that you must add two rectangles (a, b) and another square {b r ) in order to transform a square a' into a square (a + />)-. The formula is therefore that of Eucl. n. 4, a 1 + iab + b' = (a+ b)*A The first important question in relation to the above is that of date. Biirk assigns to the Apastamba- &ulba-$atra a date at least as early as the 5th or 4th century b.c. He observes however (what is likely enough) that the matter of it must have been much older than the book itself. Further, as regards one of the constructions for right angles, that by means of cords of lengths 15, 36, 39, he show? that it was known at the time of the TAittirtya- Samhita and the SatapatAa-Br&Amatfa, still older works belonging to the 8th century b.c. at latest. It may be that (as Biirk maintains) the discovery that triangles with sides (a, b, c) in rational numbers such that «* + b* = f are right-angled was nowhere made so early as in India. We find however in two ancient Chinese treatises (1) a statement that the diagonal of the rectangle (3i 4) is 5 ar) d (2) a rule for finding the hypotenuse of a " right triangle " from the sides, while tradition connects both works with the name of Chou Kung I. 47] PROPOSITION 47 363 who died 1105 b.c (D. E. Smith, History of Mathematics, 1, pp. 30—33, 11. p. 288). As regards the various " rational rectangles " used by Apastamba, it is to be observed that two of the seven, viz. 8, 15, 17 and 12, 35, 37, do not belong to the Pythagorean series, the others consist of two which belong to it, viz. 3, 4, S ar| d 5 1 '*i '3> a "d multiples of these. It is true, as remarked by Zeuthen (op, cit. p. 841), that the rules of 11. 7 and in. 9, numbered (9) and (ti) above respectively, would furnish the means of finding any number of "rational rectangles." But it would not appear that the Indians had been able to formulate any general rule ; otherwise their list of such rectangles would hardly have been so meagre. Apastamba mentions seven only, really reducible to four {though one other, 7, 24, 25, appears in the Baudhayana- S\-S., supposed to be older than Apastamba). These are all that Apastamba knew of, for he adds (v. 6): "So many recognisable (erkennbare) constructions are there," implying that be knew of no other "rational rectangles" that could be employed. But the words also imply that the theorem of the square on the diagonal is also true of other rectangles not of the " recognisable " kind, i.e. rectangles in which the sides and the diagonal are not in the ratio of integers; this is indeed implied by the constructions for ^2, ^3 etc. up to ^/6 (cf. 11. 2, vi it. s). This is alt that can be said. The theorem is, it is true, enunciated as a general proposition, but there is no sign of anything like a general proof; there is nothing to show thai the assumption of its universal truth was founded on anything better than an imperfect induction from a certain number of cases, discovered empirically, of triangles with sides in the ratio of whole numbers in which the property {1) that the square on the longest side is equal to the sum of the squares on the other two sides was found to be always accompanied by the property {2) that the latteT two sides include a right angle. It remains to consider Biirk's claim that the Indians had discovered the irrational. This is based upon the approximate value of J 2 given by Apastamba in his rule 1. 6 numbered (4) above. There is nothing to show how this was arrived at, but Thibaut's suggestion certainly seems the best and most natural. The Indians may have observed that 17'= 289 is nearly double of i2* =144. If so, the next question which would naturally occur to them would be, by how much the side 17 must be diminished in order that the square on it may be 288 exactly. If, in accordance with the Indian fashion, a gnomon with unit area were to be subtracted from a square with 17 as side, this would approximately be secured by giving the gnomon the breadth V ' T , for 1 * 1 7 x T ' T = 1. The side of the smaller square thus arrived at would be 17-3^=12 + 4+1— -^, whence, dividing out by 1 2, we have I r r ^iti+-t , approximately. 3 3 ■ 4 3 ■ 4 ■ 34 But it is a far cry from this calculation of an approximate value to the discovery of the irrational. First, we ask, is there any sign that this value was known to be inexact? It comes directly after the statement (1. 6) that the square on the diagonal of a square is double of that square, and the rule is quite boldly stated without any qualification : " lengthen the unit by one-third and the tatter by one-quarter of itself less one-thirty-fourth of this part." Further, the approximate value is actually used for the purpose of constructing a square when the side is given (n. 1). So familiar was the formula that it was apparently made the basis of a sub-division of measures of length. 364 BOOK I [i.47 Thibaut observes (Journal of the Asiatic Socitty of Bengal, xLlx., p. 141) that, according to Baudhayana, the unit of length was divided into 1 2 fingerereadths, and that one of two divisions of the fingefbrcadtA was into 34 sesame-corns, and he adds that he has no doubt that this division, which he has not elsewhere met, owes its origin to the formula for Jt. The result of using this sub- division would be that, in a square with side equal to 1 1 fingerbrtadtks, the diagonal would be 1 7 fingerbrtadtks less 1 sesame-corn. Is it conceivable that a sub-division of a measure of length would be based on an evaluation known to be inexact ? No doubt the first discoverer would be aware that the area of & gnomon with breadth W ' T and outer side 1 7 is not exactly equal to 1 but less than it by the square of ^ r or by ttWi an ^ therefore that, in taking that gnomon as the proper area to be subtracted from 1 7", be was leaving out of account the small fraction , JL . ; as, however, the object of the whole proceeding was purely practical, he would, without hesitation, ignore this as being of no practical importance, and, thereafter, the formula would be handed down and taken as a matter of course without arousing suspicion as to its accuracy. This supposition is confirmed by reference to the sort of rules which the Indians allowed themselves to regard as accurate. Thus Apastamba himself gives a construction for a circle equal in area to a given square, which is equivalent to taking ir = 3 09, and yet observes that it gives the required circle " exactly " (ill, *), while bis construction of a square equal to a circle, which he equally calls "exact," makes the side of the square equal to yfths of the diameter of the circle (111. 3), and is equivalent to taking t m 3-004. But, even if some who used the approximation for J 2 were conscious that it was not quite accurate (of which there is no evidence), there is ar immeasurable difference between arrival at this consciousness and the discovery of the irrational. As Vogt says, three stages had to be passed through before the trrationalit)' of the diagonal of a square was discovered in any real sense, (t) All values found by direct measurement or calculations based thereon have to be recognised as being inaccurate. Next {2) must supervene the conviction that it is impossible to arrive at an accurate arithmetical expression of the value. And lastly (3) the impossibility must be proved. Now there is no real evidence that the Indians, at the date in question, had even reached the first stage, still less the second or third. The net results then of Burk's papers and of the criticisms to which they have given rise appear to be these. (1) It must be admitted that Indian geometry had reached the stage at which we find it in Apastamba quite independently of Greek influence. But (2) the old Indian geometry was purely empirical and practical, far removed from abstractions such as the irrational. The Indians had indeed, by -trial in particular cases, persuaded themselves of the truth of the Pythagorean theoren- and enunciated it in all its generality ; but they had not established it by scientific proof. Alternative proofs. I. The well-known proof of 1. 47 obtained by putting two squares side by side, with their bases continuous, and cutting off right-angled triangles which can then be put on again in different positions, is attributed by an-Nairizi to Thabit b. Qurra (816 — 901 ,\.rj.). His actual construction proceeds thus. Let ABC be the given triangle right-angled at A. Construct on AB the square AD ; produce AC to Fso that BFn&.y be equal to AC. '■47] PROPOSITION 47 365 Construct on EF the square EG, and produce DH to A" so that DK may be equal to AC. It is then proved that, in the triangles BAC, CFG, KHG, BDK, the sides BA, CF, KH, BD are all equal, and the sides AC, FG, HG, DKare all equal. The angles included by the equal sides are all right angles ; hence the four triangles are equal in all respects. [l. 4] Hence BC, CG, GK, KB are all equal. Further the angles DBK, ABC are equal; hence, if we add to each the angle DBC, the angle KBC is equal to the angle ABD and is therefore a right angle. In the same way the angle CGK is right ; therefore BCGK is a square, i.e. the square on BC. Now the sum of the quadrilateral GCLH and the triangle LDB together with two of the equal triangles make the squares on AB, AC, and together with the other two make the square on BC. Therefore etc. II. Another proof is easily arrived at by taking the particular case of Pappus' more general proposition given below in which the given triangle is right-angled and the parallelograms on the sides containing the right angles are squares. If the figure is drawn, it will he seen that, with no more than one additional line inserted, it contains Thsbit's figure, so that T habit's proof may have been practically derived from that of Pappus. III. The most interesting of the remaining proofs seems to be that shown in the accompanying figure. It is given by J. W. Miiller, Systema- tistht Zusammensttllung der wiehtigstm bisher bekannttn Btweist des Pythag. Likrsatzis (Numbers, 1819), and in the second edition (Mainz, 182 1) of Ign. Hoffmann, Der Pythag. Lehr- tah mil 31 thdls bekannttn theih ruuett Btweisen [3 more in second edition]. It appears to come from one of the scientific papers of Lion- ardo da Vinci ^452— 1519). The triangle HKL is constructed on the base KH with the side KL equal to BC and the side LH equal to AB. Then the triangle HLK is equal in all respects to the triangle ABC, and to the triangle EBF. Now D£, BG, which bisect the angles ABE, CBF respectively, are in a straight line. Join BL. It is easily proved that the four quadrilaterals ADGC, EDGF, ABLK, HLBC are all equal. 366 BOOK I [i. 47 Hence the hexagons ADEFGC, ABCHLK are equal. Subtracting from the former the two triangles ABC, EBF, and from the Utter the two equal triangles ABC, HLK, we prove that the square CK is equal to the sum of the squares AE, CF. Pappus' extension of I. 47. In this elegant extension the triangle may be any triangle (not necessarily right-angled), and any parallelograms take the place of squares on two of the sides. Pappus (iv. p. 177) enunciates the theorem as follows: If ABC be a triangle, and any parallelograms whatever ABED, BCFG be described on AB, BC, and if DE, FG be produced to H, and HB be joined, the parallelograms ABED, BCFG are equal to the parallelogram contained by AC, HB in an angle which is equal to the sum of the angles BAC, DHB. Produce MB to K; through A, C draw AL, CM parallel to HK, and join LM. Then, since ALHB is a parallelo- gram, AL, HB are equal and parallel. Similarly MC, HB are equal and parallel. Therefore AL, MC are equal and parallel; whence LM, A C are also equal and parallel, and ALMC is a parallelogram. Further, the angle LAC al this parallelogram is equal to the sum of the angles BAC, DHB, since the angle DHB is equal to the angle LAB. Now, since the parallelogram DABE is equal to the parallelogram LABH (for they are on the same base AB and in the same parallels AB, DM), and likewise LABH is equal to LAKN (for they are on the same base LA and in the same parallels LA, HK), the parallelogram DABE is equal to the parallelogram LAKH. For the same reason, the parallelogram BGFC is equal to the parallelogram NKCM. Therefore the sum of the parallelograms DABE, BGFC is equal to the parallelogram LA CM, that is, to the parallelogram which is contained by AC, HB in an angle LAC which is equal to the sum of the angles BAC, BHD. "And this is far more general than what is proved in the Elements about squares in the case of right-angled (triangles)." Heron's proof that AL, BK, CF in Euclid's figure meet in a point. The final words of Proclus' note on 1. 47 (p. 429, 9 — 15) are historically interesting. He says : "The demonstration by the writer of the Elements being clear, I consider that it is unnecessary to add anything further, and that we may be satisfied with what has been written, since, in fact those who have added anything more, like Pappus and Heron, were obliged to draw upon what is proved in the sixth Book, for no really useful object." These words cannot 1.47] PROPOSITION 47 3«7 of course refer to the extension of 1, 47 given by Pappus ; but the key to them, so far as Heron is concerned, is to be found in the commentary of an-NairizI (pp. 175— 185, ed. Besthorn-Heiberg ; pp. 78 — 84, ed. Curtze) on 1. 47, wherein he gives Heron's proof that the lines AL, FC, BK in Euclid's figure meet in a point. Heron proved this by means of three lemmas which would most naturally be proved from the principle of similitude as laid down in Book vi., but which Heron, as a tour it force, proved on the principles of Book 1. only. The first lemma is to the following effect If, in a triangle ABC, DE be drawn parallel to the base BC, and if AF be drawn from the vertex A to the middle point F of BC, then AF wilt also bisect DE. This is proved by drawing HK through A parallel _, Q' to DE or BC, and HDL, KEM through D, E re \ "";"""" ? D spectively parallel to AGE, and lastly joining DE, EF. \ ;' ..•'' Then the triangles ABE, AFC are equal (being on equal bases), and the triangles DBF, EFC are also equal (being on equal bases and between the same parallels). Therefore, by subtraction, the triangles ADF, AEF are equal, and hence the parallelograms AL, AM are equal. These parallelograms are between the same parallels JLM, HK '; therefore LF, FM ait equal, whence DG, GE are also equal. The second lemma is an extension of this to the case where DE meets BA, CA produced beyond A. The third lemma proves the converse of Euclid 1. 43, that, If a paral- lelogram AB it cat into four others ADGE, DF, FGCB, CE, so that DF, CE are equal, the common vertex G will be on the diagonal AB. Heron produces AG till it meets CF in H. Then, if we join HB, we have to prove that AHB is one straight line. The proof is as follows. Since the areas DE, EC are equal, the triangles DGF, ECG are equal. If we add to each the triangle GCF, the triangles ECF, DCF are equal ; therefore MD, CF are parallel. Now it follows from 1, 34, 20 and 26 that the triangles AKE, GKD are equal in all respects; therefore EK is equal to KD. Hence, by the second lemma, CH is equal to HE. Therefore, in the triangles FHB, CHG, the two sides BF, FH are equal to the two sides G C, CH, and the angle BFHh equal to the angle GCH; hence the triangles are equal in all respects, and the angle B HE is equal to the angle GHC. Adding to each the angle GHF, we find that the angles BHF, FHG are equal to the angles CHG, GHF, and therefore to two right angles. Therefore AHB is a straight line. 368 BOOK I [■■ 47. 48 Heron now proceeds to prove the proposition that, in the accompanying figure, if AKL perpendicular to BC meet EC in M, and if BM, MG be joined, BM, MG are in one straight line. Parallelograms are completed as shown in the figure, and the diagonals OA, FH of the parallelogram FH arc drawn. Then the triangles FAH, BAC are clearly equal in all respects ; therefore the angle UFA is equal to the angle ABC, and therefore to the angle CAK (since AK is perpendicular to BC). But, the diagonals of the rectangle FH cutting one another in Y, FY is equal to YA, and the angle HFA is equal to the angle OAF. Therefore the angles OAF, CAK are equal, and accordingly OA, AK are in a straight line. Hence OM is the diagonal of SQ; therefore AS is equal to AQ, and, if we add AM to each, FM is equal to MH. But, since EC is the diagonal of the parallelogram FN, FM \s equal to MN. Therefore MH is equal to MN; and, by the third lemma, BM, MG are in a straight line. Proposition 48. If in a triangle the square on one of the sides be equal to the squares on the remaining two sides of the triangle, the angle contained by the remaining two sides of the triangle is right. For in the triangle ABC let the square on one side BC be equal to the squares on the sides BA, A C ; I say that the angle BA C is right. For let AD be drawn from the point A at right angles to the straight line AC, let AD be made equal to BA, and let DC be joined. Since DA is equal to AB, the square on DA is also equal to the square on AB. Let the square on A C be added to each ; I. 48] PROPOSITIONS 47, 48 369 therefore the squares on DA, AC are equal to the squares on BA, AC. But the square on DC is equal to the squares on DA, AC, for the angle DAC is right ; [i. 47] and the square on BC is equal to the squares on BA, AC, for this is the hypothesis ; therefore the square on DC is equal to the square on BC, so that the side DC is also equal to BC And, since DA is equal to AB, and AC is common, the two sides DA, AC are equal to the two sides BA, AC; and the base DC is equal to the base BC ; therefore the angle DAC is equal to the angle BAC. [1, 8] But the angle DA C is right ; therefore the angle BAC is also right. Therefore etc, q, e. d P roc! us' note (p. 430) on this proposition, though it does not mention Heron's name, gives an alternative proof, which is the same as that definitely attributed by an-NairizT to Heron, the only difference being that Proclus demonstrates two cases in full, while Heron dismisses the second with a " similarly." The alternative proof is another instance of the use of 1. 7 as a means of answering objections. If, says Proclus, it be not admitted that the perpendicular AD may be drawn on the opposite side of A C from B, we may draw it on the same side as AB, in which case it is impossible that it should not coincide with AB. Proclus takes two cases, first supposing that the perpendicular falls, as AD, within the angle CAB, and secondly that it falls, as AE, outside that angle. In either case the absurdity results that, on the same straight line AC and on the same side of it, AD, DC must be re- spectively equal to AB, BC, which contradicts 1. 7. Much to the same effect is the note of De Morgan that there is here " an appearance of avoiding indirect demonstration by drawing the triangles on different sides of the base and appealing to I. 8, because drawing them on the same side would make the appeal to 1. 7 (on which, however, 1. 8 is founded)." BOOK II. DEFINITIONS. i . Any rectangular parallelogram is said to be contained by the two straight lines containing the right angle. 2. And in any parallelogrammic area let any one whatever of the parallelograms about its diameter with the two comple- ments be called a gnomon. Definition i. II aV mpak\Ti\Bypa.fLfiQv 6p$ayutvtov Trfpicx«rdat Xfytrai vrro Svo ruv ryy 6p9^y ymvlav TTcptcxovo-cuy ttf}u&v> As the full expression in Greek for "the angle BAC" is "the angle contained by the (straight lines) BA, AC," ij ujto tuv BA, Ar irtpit^pftirj] ywrta., so the full expression for " the rectangle contained by BA, AC" is rd vro riav BA, Ar wipLttfptyov 6pSoyuvu>ir. In this case too BA, Ar is commonly abbreviated by the Greek geometers into BAI\ Thus in Archi- medes and Apollonius to vwh BAT or to vh-o tup BAT means the rectangle BA, AC, just as tJ vtto BAr means the angle BAC; the gender of the article shows which is meant in each case. In the early Books Euclid uses the full expression to vwo t£v BA, Ar; but the shorter form to vtto w BAr is found from Book x. onwards. Cf. xn. u, where to (jfixjpaTa) titl t) a = ab + a*, 4. {a + bf - a* + b* + tab. GEOMETRICAL ALGEBRA 373 . (a + b ,\ a (a + 4\* or (a + £) (a - £) + 0* = a*, 6. (za + *)* + a* = (,i + 4)*, or <«+#{/»-«} +#-# 7. {a + b)' + a 1 = a (a + 6) a + P, 8. 4(a + b)a + b* = t(a4i) + e\\ or( + {a-0)*=j(a« + jS'), 10. (2a + b) 1 + ** = x la' + (a + *)«}, or (a + £)* + (0 - a)* = 3 (a' + 0*). f he form of these identities may of course be varied according to the different symbols which we may use to denote particular portions of the lines given in Euclid's figures. They are, for the most part, simple identities, but there is no reason to suppose that these were the only applications of the geometrical algebra that Euclid and his predecessors had been able to make. We may infer the very contrary from the fact that Apollonius in his Comes frequently states without proof much more complicated propositions of the kind. It is important however to bear in mind that the whole procedure of Book it. is geometrical; rectangles and squares are shown in the figures, and the equality of certain combinations to other combinations is proved by those figures. We gather that this was the classical or standard method of proving such propositions, and that the algebraical method of proving them, with no figure except a tine with points marked thereon, was a later introduction. Accordingly Eutocius 1 method of proving certain lemmas assumed by Apollonius {ConUs, 11. 23 and in. 29) probably represents more nearly than Pappus' proof of the same the point of view from which Apollonius regarded them. It would appear that Heron was the first to adopt the algebraical method of demonstrating the propositions of Book 11., beginning from the second, without figures, as consequences of the first proposition corresponding to a(b + c + d) = ab + ae + ad, According to an-Nairiii (ed. Curtze, p. 89), Heron explains that it is not possible to prove 11. 1 without drawing a number of lines (i.e. without actually drawing the rectangles), but that the following propositions up to 11. 10 inclusive can be proved by merely drawing one line. He distinguishes two varieties of the method, one by dissolutio, the other by composition by which he seems to mean splitting-up of rectangles and squares, and combination of them into others. But in his proofs he sometimes combines the two varieties. When he comes to 11. 11, he says that it is not possible to do without a figure because the proposition is a problem, which accordingly requires an operation and therefore the drawing of a figure. The algebraical method has been preferred to Euclid's by some English editors ; but it should not find favour with those who wish to preserve the J74 BOOK II essential features of Greek geometry as presented by its greatest exponents, or to appreciate their point of view. It may not be out of place to add a word with reference to the geometrical equivalent of the algebraical operations. The addition and subtraction of quantities represented in the geometrical algebra by lines is of course effected by producing the line to the required extent or cutting off a portion of it. The equivalent of multiplication is the construction of the rectangle of which the given lines are adjacent sides. The equivalent of the division of one quantity represented by a line by another quantity represented by a line is simply the statement of a ratio between lines on the principles of Books v. and vi. The division of a product of two quantities by a third is represented in the geometrical algebra by the finding of a rectangle with one side of a given length and equal to a given rectangle or square. This is the problem of application of areas solved tn I. 44, 45. The addition and subtraction of products is, in the geometrical algebra, the addition and subtraction of rectangles or squares ; the sum or difference can be transformed into a single rectangle by means of the application of areas to any line of given length, corresponding to the algebraical process of finding a common measure. Lastly, the extraction of the square root is, in the geometrical algebra, the finding of a square equal to a given rectangle, which is done in ll. 14 with the help of 1. 47. BOOK II. PROPOSITIONS. Proposition i. If there be two straight lines, and one of them be cut into any number of segments whatever, the rectangle contained by the two straight lines is equal to the rectangles contained by the uncut straight line and each of the segments. s Let A, BC be two straight lines, and let BC be cut at random at the points D, E ; I say that the rectangle contained by A, BC is equal to the rectangle contained by A, BD, that contained by A, DE and 10 that contained by A, EC. For let BF be drawn from B at right angles to BC; [i. it] let BG be made equal to A, [t. 3] through G let GH be drawn is parallel to BC, [u 31] and through D, E, C let DK, EL, CH be drawn parallel to BG. Then BH is equal to BK, DL, EH. ao Now BH is the rectangle A, BC, for it is contained by GB, BC, and BG is equal to A ; BK is the rectangle A, BD, for it is contained by GB, BD, and BG is equal to A ; and DL is the rectangle A, DE, far DK, that is BG [1. 34], ^5 is equal to A. Similarly also EH is the rectangle A, EC. Therefore the rectangle A, BC is equal to the rectangle A, BD, the rectangle A, DE and the rectangle A, EC. Therefore etc. Q. E. D. 376 BOOK II [n. i, 3 so. the rectangle A, BC. From this point onward I shall Iran slate thus in cases where Euclid leaves out the word contained {irtpitxwtrw}- Though the word "rectangle" is also omitted in the Greek (the neuter article l>eing sufficient to show that the rectangle is meant), it cannot he dispensed with in English. De Morgan advises the use of the expres- sion "the rectangle uiuitr two lines." This does not seem to me a very good expression, and, if used in a translation from the Greek, it might suggest that irwi in ri iri meant uxdrr, which it does not. This proposition, the geometrical equivalent of the algebraical formula a(b + e + d+ ...)=-ab + M+-ad+ ..., can, of course, easily be extended so as to correspond to the more general algebraical proposition that the product of an expression consisting of any number of terms added together and another expression also consisting of any number of terms added together is equal to the sum of all the products obtained by multiplying each term of one expression by all the terms of the other expression, one after another. The geometrical proof of the more general proposition would be effected by means of a figure showing all the rectangles corresponding to the partial products, in the same way as they are shown in the simpler case of II. i ; the difference would be that a series or parallels to DC would have to be drawn as well as the series of parallels to bjk Proposition 2. If a straight line be cut at random, the rectangle contained by the whole and both of the segments is equal to the square on the whole. For let the straight line AB be cut at random at the point C; I say that the rectangle contained by AB, BC together with the rectangle contained by BA, AC is equal to the square on AB. For let the square ADEB be described on AB [1. 46], and let CF be drawn through C parallel to either AD or BE. [1. 31] Then AE is equal to AF, CE. Now AE is the square on AB ; AF is the rectangle contained by BA, AC, for it is contained by DA, AC, and AD is equal to AB ; and CE is the rectangle AB, BC, for BE is equal to AB. Therefore the rectangle BA, AC together with the rect- angle AB, BC is equal to the square on AB. Therefore etc. Q. E, D. A C B OF I! It, a] PROPOSITIONS i. 2 377 The fact asserted in the enunciation of this proposition has already been used in the proof of 1. 47 ; but there was no occasion in that proof to observe that the two rectangles BL, CL making up the square on BC are the rectangles contained by BC and the two parts, respectively, into which it is divided by the perpendicular from A on BC. It is this fact which it is necessary to state in this proposition, in accordance with the plan of Book IT. The second and third propositions are of course particular cases of the first They were no doubt separately enunciated by Euclid in order that they might be immediately available for use hereafter, instead of having to be deduced for the particular occasion from 11. 1. For, if they had not been thus separately stated, it would scarcely have been practicable to quote them later without explaining at the same time that they are included in 11. 1 as particular cases. And, though the propositions are not used by Euclid in the later propositions of Book 11., they are used afterwards in MIL 10 and IX. 15 respectively ; and they are of extreme importance for geometry generally, being constantly used by Pappus, for example, who frequently quotes the third proposition by the Book and number. Attention has been called to the fact that 11. t is never used by Euclid ; and this may seem no less remarkable than the fact that 11. 2, 3 are not again used in Book 11. But it is important, I think, to observe that the proof) of all the first ten propositions of Book 11. are practically independent of each other, though the results are really so interwoven that they can often be deduced from each other in a variety of ways. What then was Euclid's intention, first in inserting some propositions not immediately required, and secondly in making the proofs of the first ten practically independent of each other? Surely the object was to show the power of the method of geometrical algebra as much as to arrive at results. From the point of view of illustrating the method, there can be no doubt that Euclid's procedure is far more instructive than the semi-algebraical substitutes which seem to rind a good deal of favour; practically it means that, instead of relying on our memory of a few standard formulae, we can use the machinery given us by Euclid's method to prove immediately ab initio any of the propositions taken at random. Let us contrast with Euclid's plan the semi-algebraical alternative. One editor, for example, thinks that, as II. 1 is not used by Euclid afterwards, it seems more logical to deduce from it those of the subsequent propositions which can be readily so deduced. Putting this idea into practice, he proves 11. 2 and 3 by quoting 11. 1, then proves 11. 4 by means of 11. 1 and 3, 11. 5 and 6 by means of 11. 1, 3 and 4, and so on. The result is ultimately to deduce the whole of the first ten propositions from 11. 1, which Euclid does not use at all ; and this is to give an importance to 11. 1 which is altogether dispro- portionate and, by starting with such a narrow foundation, to make the whole structure of Book 11. top-heavy. Editors have of course been much influenced by a desire to make the proofs of the propositions of Book 11. easier, as they think, for schoolboys. But, even from this point of view, is it an improvement to deduce 11. 2 and 3 from 11. 1 as corollaries P I doubt it. For, in the first place, Euclid's figures visua/ist the results and so make it easier to grasp their meaning ; the truth of the propositions is made clear even to the eye. Then, in the matter of brevity, to which such an exaggerated importance is attached, Euclid's proof positively has the advantage. Counting a capital letter or a collocation of such as one word, I find, e.g., that Mr H. M. Taylor's proof of 11. 2 contains 378 BOOK II [ii. *, 3 1 20 words, of which 8 represent the construction. Euclid's as above trans- lated has 126 words, of which 22 are descriptive of the construction; therefore the actual proof by Euclid has 8 words fewer than Mr Taylor's, and the extra words due to the construction in Euclid are much more than atoned for by the advantage of picturing the result in the figure. The advantages then which Euclid's method may claim are, I think, these: in the case of II. 2, 3 it produces the result more easily and clearly than does the alternative proof by means of 11. 1, and, in its general application, it is more powerful in that it makes us independent of any recollection of results. Proposition 3. If a straight line be cut at random, the rectangle contained by the whole and one of the segments is equal to the rectangle contained by the segments and the square on the aforesaid segment. For let the straight line AB be cut at random at C; I say that the rectangle contained by AB, BC is equal to the rectangle contained by AC, CB together with the square on BC. For let the square CDEB be de- scribed on CB ; [1. 46] let ED be drawn through to F, and through A let AF be drawn parallel to either CD or BE. [1. 31] Then AE is equal to AD, CE. Now AE is the rectangle contained by AB, BC, for it is contained by AB, BE, and BE is equal to BC ; AD is the rectangle AC, CB, for DC is equal to CB ; and DB is the square on CB. Therefore the rectangle contained by AB, BC is equal to the rectangle contained by AC, CB together with the square on BC. Therefore etc. Q. E. D. If we leave out of account the contents of Book 11. itself and merely look to the applicability of propositions to general use, this proposition and the preceding are, as already indicated, of great importance, and particularly so to the semi-algebraical method just described, which seems to have found its first exponents in Heron and Pappus. Thus the proposition that the difference of the squares on two straight Urns is equal to the rectangle contained by the sum "■ 3. 4] PROPOSITIONS a— 4 379 and the differtna of the straight lines, which is generally given as equivalent to ii. 5, 6, can be proved by means of ii. i, z, 3, as shown by Lardner. For suppose the given straight lines are ft C P A B, BC, the latter being measured along BA. , Then, by n. 2, the square on AB is equal to. the sum of the rectangles AB, BC and AB, AC. By 11. 3, the rectangle AB, BC is equal to the sum of the square on BC and the rectangle AC, CB. Therefore the square on AB is equal to the square BC together with the sum of the rectangles AC, AB and AC, CB. But, by 11. 1, the sum of the latter rectangles is equal to the rectangle contained by AC and the sum of A3, BC, i.e. the rectangle contained by the sum and difference of AB, BC. Hence the square en AB is equal to the square on BC and the rectangle contained by the sum and difference of AB, BC\ that is, the difference of the squares on AB, BCis equal to the rectangle contained by the sum and difference of AB, BC. Proposition 4. If a straight tine be cut at random, the square on the whole is equal to the squares on the segments and twice the rectangle contained by the segments. For let the straight line AB be cut at random at C ; s I say that the square on AB is equal to the squares on A C, CB and twice the rectangle contained by AC, CB. For let the square ADEB be de- scribed on AB, [1. 46] 10 let BD be joined ; through C let CF be drawn parallel to either AD or EB, and through G let HK be drawn parallel to either AB or DE. [1. 31] is Then, since CF is parallel to AD, and BD has fallen on them, the exterior angle CGB is equal to the interior and opposite angle ADB. [t 19] But the angle ADB is equal to the angle ABD, io since the side BA is also equal to AD ; [f. 5] therefore the angle CGB is also equal to the angle GBC, so that the side BC is also equal to the side CG. [1. 6] A C 1 i H K Q C > ( E 3«o BOOK II [n. 4 But CB is equal to GK, and CG to KB ; [i. 34] therefore GK is also equal to KB ; *s therefore CGKB is equilateral. I say next that it is also right-angled. For, since CG is parallel to BK, the angles KBC, GCB are equal to two right angles. [l. 2 9 ] But the angle KBC is right ; 3° therefore the angle BCG is also right, so that the opposite angles CGK, GKB are also right. b- 34] Therefore CGKB is right-angled ; and it was also proved equilateral ; therefore it is a square ; 35 and it is described on CB. For the same reason HF is also a square ; and it is described on HG, that is AC. [1. 34] Therefore the squares HF, AX" are the squares on AC, CB. 4° Now, since A G is equal to GE, and A G is the rectangle A C, CB, for GC is equal to CB, therefore GE is also equal to the rectangle AC, CB. Therefore AG, GE are, equal to twice the rectangle AC, CB. +s But the squares HF, CK are also the squares on AC, CB; therefore the four areas HF, CK, AG, GE are equal to the squares on AC, CB and twice the rectangle contained by AC, CB. But HF, CK, AG, GE are the whole A DEB, 50 which is the square on AB. Therefore the square on AB is equal to the squares on AC, CB and twice the rectangle contained by AC, CB. Therefore etc. Q. E. D- 1. twice the rectangle contained by the segments. By a carious idiom ihjs is in Greek "the rectangle Mi contained by the segments." Similarly "twice the rectangle contained by AC, CB" is expressed as "the rectangle twin contained by AC, CB" (rilli vri>&* Ar, TB ittpaxbfAtirw Apfhy4vtQi'), 35, 38. described, jo, 43. the squares (before "on"). These words are not in the Greek, which limply says that the squares "are on " {tlalr iri) their respective sides. 46. areas. It is necessary to supply some substantive (the Greek leaves it to be under- stood); and I prefer "areas " to " figures." it. 4] PROPOSITION 4 381 The editions of the Greek text which preceded that of E. F. August (Berlin, 1826 — 9) give a second proof of this proposition introduced by the usual word dXAun or " otherwise thus." Heiberg follows August in omitting this proof, which is attributed to Theon, and which is indeed not worth reproducing, since it only differs from the genuine proof in that portion of it which proves that CGKB is a square. The proof that CGKB is equilateral is rather longer than Euclid's, and the only interesting point to notice is that, whereas Euclid still, as in 1. 46, seems to regard it as necessary to prove that all the angles of CGKB are right angles before he concludes that it is right- angled, Theon says simply "And it also has the angle CBK right ; therefore CK is a square." The shorter form indicates a legitimate abbreviation of the genuine proof; because there can be no need to repeat exactly that part of the proof of 1. 46 which shows that all the angles of the figure there constructed are right when one is. There is also In the Greek text a Porism which is undoubtedly interpolated : " From this it is manifest that in square areas the parallelograms about the diameter are squares." Heiberg doubted its genuineness when preparing his edition, and conjectured that it too may have been added by Theon ; but the matter is placed beyond doubt by a papyrus-fragment referred to already (see Heiberg, Paralipomena zu Euklid, in Hermes xxxvm., 1903, p. 48) in which the Porism was evidently wanting. It is the only Porism in Book 11., but does not correspond to Proclus* remark (p. 304, 2) that "the Porism found in the second book belongs to a problem." Heiberg regards these words as referring to the Porism to iv. 15, the correct reading having probably been not itmifnf but 8', i.e. TwapTiii. The semi -algebraical proof of tins proposition is very easy, and is of course old enough, being found in Clavius and in most later editions. It proceeds thus. By 11, 1, the square on AB is equal to the sum of the rectangles AB, AC and AB, CB. But, by 11. 3, the rectangle AB, AC is equal to the sum of the square on AC and the rectangle AC, CB ; while, by 11, 3, the rectangle AB, CB is equal to the sum of the square on BC and the rectangle AC, CB. Therefore the square on AB is equal to the sum of the squares on AC, CB and twice the rectangle AC, CB. The figure of the proposition also helps to visualise, in the orthodox manner, the proof of the theorem deduced above from 11. 1 — 3, viz. that the difference of the squares on two given straight lines is equal to the rectangle contained by the sum and the difference of the lines. For, if the lines be AB, BC respectively, the shorter of the lines being measured along BA, the figure shows that the square AE is equal to the sum of the square CK and the rectangles AF,FK, that is, the square on A B is equal to the sum of the square on BC and the rectangles AB, AC and AC, BC. But the rectangles AB, AC and BC, AC ait, by 11. 1, together equal to the rectangle contained by ^Cand the sum of AB, BC, i.e to the rectangle contained by the sum and difference of AB, BC. Whence the result follows as before. 3»a BOOK II ["• 4. 5 The proposition 11. 4 can also be extended to the case where a straight line is divided into any number of segments ; for the figure will show in like manner that the square on the whole line is equal to the sum of the squares on all the parts together with twice the rectangles contained by every pair of the parts. Proposition 5. If a straight line be cut into equal and unequal segments, the rectangle contained by the unequal segments of the whole together with the square on the straight line between the points of section is equal to the square on the half For let a straight line AB be cut into equal segments at C and into unequal segments at D ; I say that the rectangle contained by AD, DB together with the square on CD is equal to the square on CB. For let the square CEFB be described on CB, [1. 46] and let BE be joined ; through D let DG be drawn parallel to either CE or BF, through H again let KM be drawn parallel to either AB or EF t and again through A let AK be drawn parallel to either CL or BM. [1.31] Then, since the complement CH is equal to the comple- ment HF % [1. 43] let DM be added to each ; therefore the whole CM is equal to the whole DF. But CM is equal to AL, since AC is also equal to CB ; • [1. 3 6 ] therefore AL is also equal to DF. Let CH be added to each ; therefore the whole AH is equal to the gnomon NOP. ii. 5] PROPOSITIONS 4, S 3°*3 But AH is the rectangle AD, DB, for DH is equal to DB, therefore the gnomon NOP is also equal to the rectangle AD, DB. Let LG, which is equal to the square on CD, be added to each ; therefore the gnomon NOP and LG are equal to the rectangle contained by AD, DB and the square on CD. But the gnomon NOP and LG are the whole square CEFB, which is described on CB ; therefore the rectangle contained by AD, DB together with the square on CD is equal to the square on CB. Therefore etc. q. e. d. 3. between the points of section, literally "between the icct/mr," ihc word being the same (rofii) is that used of ■ conic irriim. It will be observed that the gnomon is indicated in the figure by three separate letters and a dotted carve. This is no doubt a clearer way of showing what exactly the gnomon is than the method usual in our text -books. In this particular case the figure of the ttss. has hiv M's in it, the gnomon being MN2. I have corrected the lettering to avoid confusion. It is easily seen that this proposition and [he next give exactly the theorem already alluded to under the last propositions, namely that the difference of the squares on two straight lines is equal te the rectangle contained by their sum and difference. The two given lines are, in 11. 5, the lines CB and CD, and their sum and difference are respectively equal to AD and DB. To show that 11. 6 gives the same theorem we have only to make CD the greater line and CB the less, i.e. to draw CD' equal to CB, measure . cob CB along it equal to CD, and then ' ' produce B C to A', making A'C equal a| gj p' O * to BC\ whence it is immediately clear that A' D' on the second line is equal to AD on the first, while DB is also equal to DB, so that the rectangles AD, DB and A'D", DB are equal, while the difference of the squares on CB, CD is equal to the difference of the squares on CD, CB. Perhaps the most important fact about 11. 5, 6 is however their bearing on the Geometrical solution of a quadratic equation. Suppose, in the figure of 11. 5, that AB = a, DB = x ; then «*-*■= the rectangle AH = the gnomon NOP. Thus, if the area of the gnomon is given (=*', say), and if a is given (= AB), the problem of solving the equation is, in the language of geometry, To a given straight lint (a) to apply a rectangle which shall Si equal to a given square (c5*) and shall fall snort by a square figure, Le. to construct the rectangle AH at the gnomon NOP, Now we are told by Proclus (on 1. 44) that "these propositions are ancient 3»4 BOOK II t»-5 a teu-q r^-> d e 6 / O // H and the discoveries of the Muse of the Pythagoreans, the application of areas, their exceeding and their falling-short" We can therefore hardly avoid crediting the Pythagoreans with the geometrical solution, based upon ti. 5, 6, of the problems corresponding to the quadratic equations which are directly obtainable from them. It is certain that the Pythagoreans solved the problem in n. 1 1, which corresponds to the quadratic equation a (a — *) = ar 1 , and Simson has suggested the following easy solution of the equation now in question, ax~x*=P, on exactly similar lines. Draw CO perpendicular to AB and equal to b; produce OC to A 7 so that ON= CB (or \a); and with O as centre and radius ON describe a circle cutting CB in D. Then DB (or x) is found, and therefore the required rectangle AH. For the rectangle AD, DB together with the square on CD is equal to the square on CB, [n. S ] i.e. to the square on OD, i.e. to the squares on OC, CD; [l. 47] whence the rectangle AD, DB is equal to the square on OC, or ax - x* = b*. It is of course a necessary condition of the possibility of a real solution that P must not be greater that (Ja)'. This condition itself can easily be obtained from Euclid's proposition ; for, since the sum of the rectangle AD, DB and the square on CD is equal to the square on CB, which is constant, it follows that, as CD diminishes, i.e. as D moves nearer to C, the rectangle AD, DB increases and, when D actually coincides with C, so that CD vanishes, the rectangle AD, DB becomes the rectangle AC, CB, i.e. the square on CB, and is a maximum. It wilt be seen also that the geometrical solution of the quadratic equation derived from Euclid does not differ from our practice of solving a quadratic by completing the square on the side containing the terms in x* and x. But, while in this case there are two geometrically real solutions (because the circle described with ON as radius will not only cut CB in D but will also cut AC in another point E), Euclid's fig a re corresponds to one only of the two solutions. Not that there is any doubt that Euclid was aware that the method of solving the quadratic gives two solutions ; he could not fail to see that x = BE satisfies the equation as well as x = BD. If however he hud actually given us the solution of the equation, he would probably have omitted to specify the solution * = BE because the rectangle found by means of it, which would be a rectangle on the base AE (equal to BD) and with altitude EB (equal to AD), is really an equal rectangle to that corresponding to the other solution x = BD ; there is therefore no real object in distinguishing two solutions. This is easily understood when we regard the equation as a statement of the problem of finding two magnitudes when their sum (a) and product (b*) are given, i.e. as equivalent to the simultaneous equations x+y = a, xy = b*. ii. s, 6] PROPOSITIONS 5, 6 385 These symmetrical equations have really only one solution, as the two apparent solutions are simply the result of interchanging the values of x and y. This form of the problem was known to Euclid, as appears from the Data, Prop. 85, which states that, If two straight lines contain a parallelogram given in magnitude in a given angle, and if the sum of them be given, then shall each of them be given. This proposition then enables us to solve the problem of finding a rectangle the area and perimeter of which are both given ; and it also enables us to infer that, of all rectangles of given perimeter, the square has the greatest area, while, the more unequal the sides are, the less is the area. If in the figure of 11. 5 we suppose that AD=a, BD=b, we find that CB = (« + *)/ 2 and CD = (a—b)li, and we may state the result of the proposition in the following algebraical form ffl-ffi-* This way of stating it (which could hardly have escaped the Pythagoreans) gives a ready means of obtaining the two rales, respectively attributed to the Pythagoreans and Plato, for finding integral square numbers which are the sum of two other integral square numbers. We have only to make ab a perfect square in the above formula. The simplest way in which this can be done is to put a = n', b=i, whence we have and in order that the first two squares may be integral a 1 , and therefore n, must be odd Hence the Pythagorean rule. Suppose next that a = in*, b-%, and we have («*+.)'-(«*-i)' = 4A\ whence Plato's rale starting from an even number in. Proposition 6. If a straight line be bisected and a straight line be added to ii in a straight line, the rectangle contained by the whole with the added straight line and the added straight line together with the square on the half is equal to the square on the straight line made up of the half and the added straight line. For let a straight line AB be bisected at the point C, and let a straight line BD be added to it in a straight line ; I say that the rectangle contained by AD, DB together with the square on CB is equal to the square on CD. For let the square CEFD be described on CD, [1. 46] and let DE be joined ; through the point B let BG be drawn parallel to either EC or DF, A C B D .' H M K L N / j f~~ r 3 S6 BOOK II [11.6 through the point H let KM be drawn parallel to either A 3 or EF, and further through A let y4A* be drawn parallel to either CL or DM. [i. 31] Then, since ^4C is equal toC#, j4Z is also equal to CH. [1. 36] But CH is equal to HF Ql 43] Therefore AL is also equal toJKE Let CM be added to each ; therefore the whole AM is equal to the gnomon NOP. But /f4f is the rectangle AD, DB, for DM is equal to Z?/? ; therefore the gnomon NOP is also equal to the rectangle AD, DB. Let LG, which is equal to the square on BC, be added to each ; therefore the rectangle contained by AD, DB together with the square on CB is equal to the gnomon NOP and LG. But the gnomon NOP and LG are the whole square CEFD, which is described on CD ; therefore the rectangle contained by AD, DB together with the square on CB is equal to the square on CD. Therefore etc. Q. E, D. In this case the rectangle AD, DB is "a rectangle applied to a given straight line (AB) but exceeding by a square (the side of which is equal to BD) " ; and the problem suggested by 11. 6 is to rind a rectangle of this description equal to a given area, which we will, for convenience, suppose to be a square ; Le., in the language of geometry, to apply to a given straight line a rectangle which shall be equal to a given square and shall exceed by a square figure. We suppose that in Euclid's figure AB = a, BD=x; then, if the given square be F, the problem is to solve geometrically the equation ax + jc* = #•. The solution of a problem theoretically equivalent to the solution of a quadratic equation of this kind is presupposed in the fragment of Hippocrates' Quadrature of lunes preserved in a quotation by Simplicius (Comment, in ii. 6] PROPOSITION 6 387 Aristot. Phys. pp. 61 — 68, ed. Diels) from Eudemus' History of Geometry, In this fragment Hippocrates (5th cent. b,c.) assumes the following construction. AB being the diameter and O the centre of a semicircle, and C being the middle point of OB and CD at right angles to AB, a straight line of length such that its square is \\ times the square on the radius (i.e. of length aj$, where a is the radius) is to be so placed, as EF, between CD and the circumference AD Jiat it "verges towards B," that is, EF when produced passes through B. Now the right-angled triangles BFC, BAE are similar, so that BF:BC=BA .BE, and therefore the rectangle BE, BF= rect BA, BC = sq. on BO. In other words, EF ( = o ,/f) being given in length, BF ( = x, say) has to be found such that (7t « + *)#=**; or the quadratic equation ,/£ ax + ar* = a" has to be solved. A straight line of length ajl would easily be constructed, for, in the figure, CD*=AC. CB = \a\ or CD=\aJs, and aj\ is the diagonal of a square of which CD is t,he side. There is no doubt that Hippocrates could have solved the equation by the geometrical construction given below, but he may have contemplated, on this occasion, the merely mcehanital process of placing the straight line of the length required between CD and the circumference AD and moving it until E, F, B were in a straight line. Zeuthen {Die Lehre von den Kegelschnittm im Altertum, pp. 370, 27 r) thinks this probable because, curiously enough, the fragment speaks immediately afterwards of "joining B to F." To solve the equation we have to find the rectangle AH, or the gnomon NOP, which is equal in area to £* and has one of the sides containing the inner right angle equal to CB or \a. Thus we know (Ja)* and £*, and we have to find, by I. 47, a square equal to the sum of two given squares. To do this Simson draws BQ at right angles to AB and equal to b, joins CQ and, with centre C and radius CQ, describes a circle cutting AB produced in D. Thus BD, or x, is found. Now the rectangle AD, DB together with the square on CB is equal to the square on CD, i.e. to the square on CQ, i.e. to the squares on CB, BQ. 3 88 BOOK II [11. 6, 7 Therefore the rectangle AD, DB is equal to the square on BQ, that is, jx + x* — fi. From Euclid's point of view there would only be one solution in this case. This proposition enables us also to solve the equation x* — ax-& in a similar manner. We have only to suppose that AB = a, and AD (instead of BD) = x ; then .» x*—ax = the gnomon. To find the gnomon we have its area (P) and the area, CB 1 or (|o)*, by which the gnomon differs from CD 1 . Thus we can find D (and therefore AD or x) by the same construction as that just given. Converse propositions to 11. 5, 6 are given by Pappus (vii. pp. 948—950) among his lemmas to the Conits of Apollonius to the effect that, (1) if D be a point dividing AB unequally, and C another point on AB such that the rectangle AD, DB together with the square on CD is equal to the square on AC, then ^Cis equal to CB; (a) if D be a point on AB produced, and C a point on AB such that the rectangle AD, DB together with the square on CB is equal to the square on CD, then AC is equal to CB. Proposition 7. If a straight line be cut at random, the square on the whole and that on one of the segments both together are equal to twice the rectangle contained by the whole and the said segment and the square on the remaining segment. For let a straight line AB be cut at random at the point C; I say that the squares on AB, BC are equal to twice the rectangle contained by AB, BC and the square on CA. For let the square ADEB be described on AB, [1. 46] and let the figure be drawn. Then, since AG is equal to GE, [i. 43] let CF be added to each ; therefore the whole AF is equal to the whole CE, Therefore AF, CE are double of AF. But AF, CF are the gnomon KLM and the square CF ; therefore the gnomon KLM and the square CF are double of AF. ii. 7. 8] PROPOSITIONS 6—8 389 But twice the rectangle AB, BC is also double of AF ; for BF is equal to BC ; therefore the gnomon KLM and the square CF are equal to twice the rectangle AB, BC. Let DG, which is the square on AC, be added to each ; therefore the gnomon KLM and the squares BG, GD are equal to twice the rectangle contained by AB, BC and the square on AC. But the gnomon KLM and the squares BG, GD are the whole ADEB and CF, which are squares described on AB, BC ; therefore the squares on AB, BC are equal to twice the rectangle contained by AB, BC together with the square on AC. Therefore etc. q. £. D. An interesting variation of the form of this proposition may be obtained by regarding AB, BC as two given straight lines of which AS is the greater, and AC as the difference between the two straight lines. Thus the proposition shows that the squares on two straight lines are together equal to twice the rectangle contained by them and the square on their difference. That is, the square en the different of two straight lines is equal to the sum of the squares on the straight lines diminished by twite the rectangle contained by them. In other words, just as 11. 4 is the geometrical equivalent of the identity (a + b) , ^d , + b t + 2ab, so 11. 7 proves that (a -t) , = a* + P-tab. The addition and subtraction of these formulae give the algebraical equivalent of the propositions it. 9, 10 and 11. 8 respectively ; and we have accordingly a suggestion of alternative methods of proving those propositions. Proposition 8. If a straight line be cut at random, four times the rectangle contained by the whole and one of the segments together with the square on the remaining segment is equal to the square described on the whole and the aforesaid segment as on one straight line. For let a straight line AB be cut at random at the point C; 1 say that four times the rectangle contained by AB, BC together with the square on AC is equal to the square described on AB, BC as on one straight line. 39° BOOK II [n. 8 A O B D M Sp N 1 K ■ 8 / I H L For let [the straight line] .5Z> be produced in a straight line [with AB\, and let BD be made equal to CB ; let the square A FFD be described on AD, and let the figure be drawn double. Then, since CB is equal to BD, while CB is equal to GK, and BD to AW, therefore <7^T is also equal to KN. For the same reason QR is also equal to RP. And, since BC is equal to BD, and 6l#" to KN, therefore CK is also equal to KD, and GA to RN [i. 36] But CK is equal to AjV, for they are complements of the parallelogram CP ; [1. 43] therefore KD is also equal to GR \ therefore the four areas DK, CK, GR, RN are equal to one another. Therefore the four are quadruple of CK. Again, since CB is equal to BD, while BD is equal to BK, that is CG, and CB is equal to GK, that is GQ, therefore CG is also equal to GQ. And, since CG is equal to GQ, and QR to RP, AG'xs also equal to MQ, and £>Z. to RF. [1. 36] But J/0 is equal to QL, for they are complements of the parallelogram ML ; [1. 43) therefore AG is also equal to RF; therefore the four areas AG, MQ, QL, RF are equal to one another. Therefore the four are quadruple of AG. But the four areas CK, KD, GR, RN were proved to be quadruple of CK; therefore the eight areas, which contain the gnomon STU, are quadruple of AK, Now, since AK is the rectangle AB, BD, for BK is equal to BD, ii. 8] PROPOSITION 8 391 therefore four times the rectangle AB, BD is quadruple of AK. But the gnomon STU was also proved to be quadruple otAK; therefore four times the rectangle AB, BD is equal to the gnomon STU. Let OH, which is equal to the square on AC, be added to each ; therefore four times the rectangle AB, BD together with the square on AC is equal to the gnomon STU and OH. But the gnomon STU and OH are the whole square AEFD, which is described on AD • therefore four times the rectangle AB, BD together with the square on AC is equal to the square on AD But BD is equal to BC; therefore four times the rectangle contained by AB, BC together with the square on AC is equal to the square on AD, that is to the square described on AB and BC as on one straight line. Therefore etc. This proposition is quoted by Pappus {p. 418, ed. Hultsch) and is used also by Euclid himself in the Data, Prop. 86. Further, it is of decided use in proving the fundamental property of a parabola. Two alternative proofs are worth giving. The first is that suggested hy the consideration mentioned in the last note, though the proof is old enough, being given by Clavius and others. It is of the semi-algebraical type. Produce AB to D (in the figure of the pro- position), so that BD is equal to BC. By 11. 4, the square on AD is equal to the squares on AB, BD and twice the rectangle AB, BD, i.e. to the squares on AB, BC and twice the rectangle AB, BC. By 11. 7, the squares on AB, BC are equal to twice the rectangle AB, BC together with the square on AC Therefore the square on AD is equal to four times the rectangle AB, BC together with the ' " square on AC. The second proof is after the manner of Euclid but with a difference. Produce BA to D so that AD is equal to BC On BD construct the square BEFD. M 39* BOOK TI [ii. 8, 9 Take BG, Elf, FK each equal to BC or AD, and draw ALP, HNM parallel to BE and GML, KPW parallel to BD. Then it can be shown that each of the rectangles BL, AK, FN, EM is equal to the rectangle AB, BC, and that PM is equal to the square on AC. Therefore the square on BD is equal to four times the rectangle AB, BC together with the square on AC. Proposition 9. If a straight line be cut into equal and unequal segments, the squares on the unequal segments of the whole are double of the square on the half and of the square on the straight line between the points of section. For let a straight line AB be cut into equal segments at C, and into unequal segments at D\ I say that the squares on AD, DB are double of the squares on AC, CD. For let CE be drawn from C at right angles to AB, and let it be made equal to either AC at CB ; let EA, EB be joined, let DF be drawn through D parallel to EC, and FG through F parallel to AB, and let AF be joined. Then, since AC is equal to CE, the angle EAC is also equal to the angle A EC. And, since the angle at C is right, the remaining angles EAC, AEC are equal to one right angle. ['• 3*] And they are equal ; therefore each of the angles CEA, CAE is half a right angle. For the same reason each of the angles CEB, EBC is also half a right angle ; therefore the whole angle AEB is right And, since the angle GEF is half a right angle. 1 : f n t / *■ \ A < > 1 ) B ii. 9] PROPOSITIONS 8, 9 393 and the angle EGF is right, for it is equal to the interior and opposite angle ECB, [l ag] the remaining angle EFG is half a right angle ; [1. 3 2 ] therefore the angle GEF is equal to the angle EFG, so that the side EG is also equal to GF. [1. 6] Again, since the angle at B is half a right angle, and the angle FDB is right, for it is again equal to the interior and opposite angle ECB, [i- *9] the remaining angle BFD is half a right angle ; [1. 3*] therefore the angle at B is equal to the angle DFB, so that the side FD is also equal to the side DB. [1. 6] Now, since AC is equal to CE, the square on AC is also equal to the square on CE; therefore the squares on AC, CE are double of the square on AC. But the square on EA is equal to the squares on AC, CE t for the angle A CE is right ; [1. 47] therefore the square on EA is double of the square on A C. Again, since EG is equal to GF, the square on EG is also equal to the square on GF; therefore the squares on EG, GF are double of the square on GF. But the square on EF is equal to the squares on EG, GF; therefore the square on EF is double of the square on GF. But GF is equal to CD ; [j. 34) therefore the square on EF is double of the square on CD. But the square on EA is also double of the square on AC; therefore the squares on AE, EF&re double of the squares on AC, CD. And the square on AF is equal to the squares on AE, EF, for the angle AEF is right ; [1. 47] therefore the square on AF is double of the squares on AC, CD. But the squares on AD, DF are equal to the square on AF, for the angle at D is right ; [1. 47] therefore the squares on AD, DF are double of the squares on AC, CD. 394 BOOK II [n. g And DF is equal to DB ; therefore the squares on AD, DB are double of the squares on AC, CD. Therefore etc. Q. E. D. It is noteworthy that, while the first eight propositions of Book it. are proved independently of the Pythagorean theorem i. 47, all the remaining propositions beginning with the 9th are proved by means of it. Also the 9th and 10th propositions mark a new departure in another respect ; the method of demonstration by showing in the figures the various rectangles and squares to which the theorems relate is here abandoned. The 9th and 10th propositions are related to one another in the same way as the 5th and 6th ; they really prove the same result which can, as in the earlier case, be comprised in a single enunciation thus : The sum of the squares on the sum and difference of two given straight lines is equal to twice the sum of the squares on the lines. The semi-algebraical proof of Prop, 9 is that suggested by the remark on the algebraical formulae given at the end of the note on 11. 7. It applies with a very slight modification to both u. 9 and 11. 10. We will put in brackets the variations belonging to 11. 10. The first of the annexed lines is the figure ^ COB for 11. 9 and the second for ti. 10. ' ' By 11. 4, the square on AD is equal to a C a D the squares on AC, CD and twice the ¥ > rectangle AC, CD. By 11. 7, the squares on CB, CD {CD, C£) are equal to twice the rectangle CB, CD together with the square on BD. By addition of these equals crosswise, the squares on AD, DB together with twice the rectangle CB, CD are equal to the squares on AC, CD, CB, CD together with twice the rectangle AC, CD. But AC, CB are equal, and therefore the rectangles AC, CD and CB, CD are equal. Taking away the equals, we see that the squares on AD t DB are equal to the squares on AC, CD, CB, CD, i.e. to twice the squares on AC, CD. To show also that the method of geometrical algebra illustrated by 11. 1 — 8 is still effective for the purpose of proving 11. 9, 10, we will now prove 11. 9 in that manner. Draw squares on AD, DB respectively as shown in the figure. Measure DH along DE equal to CD, and HL along HE also equal to CD. Draw HK, LNO parallel to EF, and CNM parallel to DE. Measure NP along NO equal to CD, F " Q M E and draw PQ parallel to DB. 1 1 C P B K H L P N 6 ii. g, io] PROPOSITIONS 9, 10 395 Now, since AD, CD are respectively equal to DE, DH, HE is equal to AC or CB ; and, since HL is equal to CD, LE is equal to DB. Similarly, since each of the segments EM, MQ is equal to CD, EQ is equal to EL or BD. Therefore OQ is equal to the square on DB. We have to prove that the squares on AD, DB are equal to twice the squares on AC, CD. Now the square on AD includes KM (the square on AC) and CH, HN (that is, twice the square on CD). Therefore we have to prove that what is left over of the square on AD together with the square on DB is equal to the square on AC. The parts left over are the rectangles CK and NE, which are equal to KJV, PM respectively. But the latter with the square on DB are equal to the rectangles KN, BMand the square OQ, i.e. to the square KM, or the square on AC. Hence the required result follows. Proposition 10. If a straight line be bisected, and a straight line be added to it in a straight line, the square on the whole with the added straight line and the square $n the added straight line both together are double of the square on the half and of the square described on the straight line made up of the half and the added straight line as on one straight line. For let a straight line AB be bisected at C, and let a straight line BD be added to it in a straight line ; I say that the squares on AD, DB are double of the squares on AC, CD. For let CE be drawn from the point C at right angles to AB [1. 11], and let it be made equal to either AC ox CB [1. 3] ; let EA, EB be joined ; through E let EF be drawn parallel to AD, and through D let FD be drawn parallel to CE. [1. 31] Then, since a straight line EF falls on the parallel straight lines EC, FD, 39$ BOOK II [n. io the angles CEF, EFD are equal to two right angles ; [i. a$] therefore the angles FEB, EFD are less than two right angles. But straight lines produced from angles less than two right angles meet ; [i. Post 5] therefore EB, FD, if produced in the direction B, D, will meet. Let them be produced and meet at G, and let AG be joined. Then, since A C is equal to CE, the angle EA C is also equal to the angle AEC ; [1. 5] and the angle at C is right ; therefore each of the angles EAC, AEC is half a right angle. [1. 32] For the same reason each of the angles CEB, EBC is also half a right angle ; therefore the angle AEB is right And, since the angle EBC is half a right angle, the angle DBG is also half a right angle. [1. 15] Rut the angle BDG is also right, for it is equal to the angle DCE, they being alternate; [1. 19] therefore the remaining angle DGB is half a right angle ; ['• 3"] therefore the angle DGB is equal to the angle DBG, so that the side BD is also equal to the side GD, [1. 6] Again, since the angle EGF is half a right angle, and the angle at F is right, for it is equal to the opposite angle, the angle at C, [1. 34] the remaining angle FEG is half a right angle ; [1. 3*] therefore the angle EGF is equal to the angle FEG, so that the side GF is also equal to the side EF. [1. 6] Now, since the square on EC is equal to the square on CA, the squares on EC, CA are double of the square on CA. But the square on EA is equal to the squares on EC, CA ; b- «] therefore the square on EA is double of the square on A C. [a k 1] ii. io] PROPOSITION to 397 Again, since FG is equal to EF, the square on FG is also equal to the square on FE ; therefore the squares on GF, FE are double of the square on EF But the square on EG is equal to the squares on GF, FE; [>• 47] therefore the square on EG is double of the square on EF. And EF is equal to CD ; [i- 34] therefore the square on EG is double of the square on CD. But the square on EA was also proved double of the square on AC; therefore the squares on AE, EG are double of the squares on AC, CD. And the square on AG is equal to the squares on AE, EG ; [i. 47] therefore the square on AG is double of the squares on AC, CD. But the squares on AD, DG are equal to the square on AG ; [•■47] therefore the squares on AD, DG are double of the squares on AC, CD. And DG is equal to DB ; therefore the squares on AD, DB are double of the squares on AC, CD. Therefore etc. Q. E. D. The alternative proof of this proposition by means of the principles exhibited in n. i — 8 follows the lines of that which I have given for the preceding proposition. It is at once obvious from the figure that the square on AD includes within it twice the square on AC together with once the square on CD. What is left over is the sum of the rectangles AH, KE. These, which are equivalent to BH, GK, make up the square on CD less the square on BD. Adding therefore the square BG to each side, we have the required result. Another alternative proof of the theorem which includes both n. 9 and 10 is worth giving. The theorem states that the sum of the squares on the sum and difference of two given straight lines is equal to twite the sum of the squares on the lines. A < ; b d H K 1 F E H L K 398 BOOK II [11,10 Let AD, DB be the two given straight lines (of which AD is the greater), placed so as to be in one straight line. Make AC equal to DB and com- pJete the figure as shown, each of the segments CG and DH being equal to AC or DB. AC B Now, AD, DB being the given straight lines, AB is their sum and CD is equal to their difference. Also AD is equal to BC. And AE is the square on AB, GK is equal to the square on CD, AK or Fffia the square on AD, and BL the square on CB, while each of the small squares AG, BH, EK, FL is equal to the square on ACazDB. We have to prove that twice the squares on AD, DB are equal to the squares on AB, CD. Now twice the square on AD is the sum of the squares on AD, CB, which is equal to the sum of the squares BL, FH , and the figure shows these to be equal to twice the inner square GK and once the remainder of the large square AE excluding the two squares AG, KE, which latter squares are equal to twice the square on AC 01 DB. Therefore twice the squares on AD, DB are equal to twice the inner square GK together with once the remainder of the large square AE, that is, to the sum of the squares AE, GK, which are the squares on AB, CD. " Side" and " diagonal " numbers giving successive approxi- mations to J2. Zeuthen pointed out {Dii Lthre von den Kegehcknitten im Alia- turn, rS86, pp. 37, 38) that 11. 9, 10 have great interest in connexion with a problem of indeterminate : g g B analysis which received much attention from the ancient Greeks. If we take the straight line AB divided at C and D as in 11. 9, and if we put CD = x, DB=y, the result obtained by Euclid, namely : AEP + DB* ^lAC + i CD*, or AD t ~iAC = iCD'-DB', becomes the formula (tx+y)* — a(x*yf = tx*-jf. If therefore x, y be numbers which satisfy one of the two equations 2x* —y* — ± 1, the formula gives us two higher numbers, x+y and 2x +y, which satisfy the other of the two equations. Euclid's propositions thus give a general proof of the very formula used for the formation of the succession of what were called " side " and " diagonal numbers." As is well known, Theon of Smyrna (pp. 43, 44, ed. Hiller) describes this system of numbers. The unit, being the beginning of all things, must be potentially both a side and a diameter. Consequently we begin with two units, the one being the first side and the other the first diameter, and (a) from the sum of them, (£) from the sum of twice the first unit and once the second, we form two new numbers 1.1 + 1 = *, 1.1 + 1 = 3. ii. ro] PROPOSITION 10 399 Of these new numbers the first is a side- and the second a diagonal-^ umber, or (as we may say) flj=2, d,= 3. In the same way as these numbers were formed from a I = 1, d, = 1, successive pairs of numbers are formed from o,, d % , and so on, according to the formula o«+i = o* + Jn rf»+i = **.. + + rf «-0* '= + t&J - a^n-A m 'ike manner, ana so on, while d,* - 30,' ■ - i. Thus the theorem is established. Euclid's propositions enable us to establish the theorem geometrically; and this fact might well be thought to confirm the conjecture that the investigation of the indeterminate equation 2^-y'-±i in the manner explained by Theon was no new thing but began at a period long before Euclid's time. No one familiar with the truth of the proposition stated by Theon could have failed to observe that, as the corresponding side- and diagonal-numbers were successively formed, the value of rf„*/ a »" would approach more and more nearly to 3, and consequently that the successive fractions dja n would give nearer and nearer approximations to the value of /, mill II il It is fairly clear that in the famous passage of Plato's Republk (546 c) about the "geometrical number" some such system of approximations is hinted at Plato there contrasts the "rational diameter of five' (/Jijti) StafMrpot tt}<; vtftwdSfK) with the " irrational " (diameter). This was certainly taken from the Pythagorean theory of numbers (cf. the expression immediately preceding, 546 B, C ntn xptxnjyopa no! /Sirra irpw aXkijka dviifnprav, with the phrase wavta yv Q • Then, by the theorem of Eucl. 11. 10, the squares on AD, DC are double of the squares on AB, BD, But the square on DC (i.e. BE) is double of the square on AB; therefore, by subtraction, the square on AD Is double of the square on BD. And the square on DF, the diagonal corresponding to the side BD, is double the square of BD. Therefore the square on DP is equal to the square on AD, so that j&Pis equal to AD. That is, while the side BD is, with our notation, a + d, the corresponding diagonal, being equal to AD, is 1a + d. In the above reference by Proclus to 11. 10 dw' imivav "by him" must apparently mean w BinXxiSov, " by Euclid," although Euclid's name has not been mentioned in the chapter; the phrase would be equivalent to saying " in the second Book of the famous Elements." But, when Proclus says "this is firwed in the second Book of the Elements," he does not imply that it had not been proved before ; on the contrary, it is clear that the theorem had been proved by the Pythagoreans, and we have therefore here a confirmation of the inference from the part played by the gnomon and by 1. 47 in Book 11. that the whole of the substance of that Book was Pythagorean. For further detailed explanation of the passages of Proclus reference should be made to Hultsch's note in KroU's Vol. 11. pp. 393 — 400, and to the separate article, also by Hultsch, in the Bibliotheca Mathematica I,, 1900, pp. 8 — 12. P. Bergh has an ingenious suggestion (see Zeitschrift fur Math. u. Physik II. IOJ PROPOSITION 10 401 xxxi. Hist-titL Abt. p. 135, and Cantor, Geschithte der Mathematik, i„ p. 437) as to the way in which the formation of the successive side- and diagonal-numbers may have been discovered, namely by observation from a very simple geometrical figure. Let ABC be an isosceles triangle, right-angled at A, with sides o._i, «„-,, d n _, respectively. If now the two sides AS, AC about the right angle be lengthened by adding Fy»it &t> ttj} tA iinTax&fr. 35, 36- which can be described, expressed by the future passive participle, iyvypa^ij. Heiberg {Mathematisches zu AristoteUs, p. »o) quotes as bearing on this proposition Aristotle's remark (De ant ma 11. 2, 413 a 19: cf. Metaph. 996 b zi) that " squaring " (Terpavuii'Krftoe) is better defined as the " finding of the mean (proportional) " than as " the making of an equilateral rectangle equal to a riven oblong," because the former definition states the cause, the latter the :ondusion only. This, Heiberg thinks, implies that in the text-books which were in Aristotle's hands the problem of 11. 14 was solved by means of proportions. As a matter of fact, the actual construction is the same in 11. 14 as in vi. 13 ; and the change made by Euclid must have been confined to substituting in the proof of the correctness of the construction an argument based on the principles of Books 1. and 11. instead of Book vi. As n. t2, 13 are supplementary to 1. 47, so 11. 14 completes the theory of transformation of areas so far as it can be carried without the use of proportions. As we have seen, the propositions 1. 42, 44, 45 enable us to construct a parallelogram having a given side and angle, and equal to any given rectilineal figure. The parallelogram can also be transformed into an equal triangle with the same given side and angle by making the other side about the angle twice the length. Thus we can, as a particular case, construct a rectangle on a given base (or a right-angled triangle with one of the sides about the right angle of given length) equal to a given square. Further, 1. 47 enables us to make a square equal to the sum of any number of squares or to the difference between any two squares. The problem still remaining unsolved is to transform any rectangle (as representing an area equal to that of any rectilineal figure) into a square of equal area. The solution of this problem, given in 11. 14, is of course the equivalent of the extraction of the square root, or of the solution of the pure quadratic equation x* = ai. Simson pointed out that, in the construction given by Euclid in this case, it was not necessary to put in the words " Let BE be greater," since the construction is not affected by the question whether BE or ED is the greater. This is true, but after all the words do little harm, and perhaps Euclid may have regarded it as conducive to clearness to have the points B, G, E, F in the same relative positions as the corresponding points A, C, D, B in the figure of 11. 5 which he quotes in the proof. EXCURSUS I. PYTHAGORAS AND THE PYTHAGOREANS. The problem of determining how much of the Pythagorean discoveries in mathematics can be attributed to Pythagoras himself is not only difficult ; it may be said to be insoluble. Tradition on the subject is very meagre and uncertain, and further doubt is thrown upon it by the well-known tendency of the later Pythagoreans to ascribe everything to the Master himself {airot iifta, Ipse dixit). Pythagoras himself left no written exposition of his doctrines, nor did any of his immediate successors, not even Hippasus, about whom the different stories ran ( t ) that he was expelled from the school because he pub- lished doctrines of Pythagoras, and (2) that he was drowned at sea for revealing the construction of the dodecahedron in the sphere and claiming it as his own, or (as others have it) for making known the discovery of the irrational or in- commensurable. Nor is the absence of any written record of Pythagorean doctrines down to the time of Philolaus to be put down to a pledge of secrecy binding the school ; at all events this did not apply to their mathematics or their physics; and it may be that the supposed secrecy was invented to account for the absence of documents. The fact seems to be that oral com- munication was the tradition of the school, while their doctrines would in the main be too abstruse to be understood by the generality of people outside. Even Aristotle felt the difficulty ; he evidently knew nothing for certain about any ethical or physical doctrines going back to Pythagoras himself; when he speaks of the Pythagorean system, he always refers it to " the Pythagoreans." sometimes even to " the so-called Pythagoreans." Since my note on Eucl. I. 47 was originally written the part of Pythagoras in the Pythagorean mathematical discoveries has been further discussed and every scrap of evidence closely, and even meticulously, examined in two long articles by Heinrich Vogt, "Die Geometrie des Pythagoras " {BibUotheca Maihemafica ix„ 1908/9, pp. 15 — 54) and "Die Entstehungsgeschichte des Irrationalen nach Plato und anderen Quelien des 4. Jahrhunderts " (Siblie- theax Mathematita x„ 1910, pp. 97 — 155). These papers would not indeed have enabled me to modify greatly what I have written regarding the supposed discoveries of Pythagoras and the early Pythagoreans, because I have through- out been careful to give the traditions on the subject for what they are worth and no more, and not to build too much upon them. It is right however to give, in a separate note, a few details of Vogt's arguments. G. Junge had, in his paper " Wann haben die Griechen das Irrationale entdeckt?" mentioned above (p. 351), tried to prove that Pythagoras himself could not have discovered the irrational ; and the object of Vogt s papers is to go further on the same lines and to show (1) that it was only the later Pythagoreans who (before 410 B.C.) recognised the incommensurability of the diagonal with the side of a square, (2) that the theory of the irrational was first discovered by Theodorus, to whom Plato refers ( Theaetttvs 141 d), and (3) that Pythagoras himself could not have been the discoverer of any one of 413 PYTHAGORAS AND THE PYTHAGOREANS the things specifically attributed to him, namely (a) the theorem of Eucl. 1. 47, (d) the construction of the five regular solids in the sense in which they are respectively constructed in Eucl. xm, (c) the application of an area in its widest sense, equivalent to the solution of a quadratic equation in its most general form. Vogt's main argument as regards (a) the theorem of 1. 47 is based on a new translation which he gives of the well-known passage of Proclus' note on the proposition (p. 426, 6 — -9), Ttijk fLtv Itrroptiv ra dp%aia fiavkopivw oxovovTa? to Qiwpij/ta toSto tit Tlvtiayopar avairtjairwruii' larlv tvpilv xal /$ov8vrt)v \ty6vra»> avrir iwl rj} cupttrtt Vogt translates this as follows : " Unter den en, welche das Altertum erforschen wollen, kann man einige tin den, welche denen Gehor geben, die dieses Theorem auf Pythagoras zuruckfiihren und ihn als Stier- opferer bei dieser Gelegenheit bezeichnen," " Among those who have a taste for research into antiquity, we can find some who give ear to those who refer this theorem to Pythagoras and describe him as sacrificing an ox on the strength of the discovery." According to this version the words t£v... ftovkopJvwv and the words dwnvfpTrovTviv.,,Kal.,,\iy6i>Tprjpa.TO<; aX^fo t'rj , p.fiQ'>vw/«>, is indeterminate. But, if these were so, all lines (including commensurable lines) would be "without a ratio" to one another, whereas the title of Democritus' work clearly implies that SXoyoi -ypau/uu are a class or classes of lines distinguished from other lines. The fact is that Democritus was too good a mathematician to have anything to do with "indivisible lines." This is confirmed by a scholium to Aristode's lie eaclo (p. 469 b 14, Brandis) which implicitly denies to Democritus any theory of indivisible lines : "of those who have maintained the existence of indivisibles, some, as for example Leucippus and Democritus, believe in indivisible bodies, others, like Xenocrates, in indivisible lines." Moreover Simpiicius tells us that, according to Democritus himself, even the atoms were, in a mathematical sense, divisible further and in fact ad infinitum. Coming now to {b) the construction of the cosmic figures, 9 rmr KoajUK&v iT)(r)tta-ra>v uvrrnuTii (Prod us, p. 65, 20), I agree with Vogt to the following extent. It is unlikely that Pythagoras or even the early Pythagoreans " con- structed" the five regular solids in the sense of a complete theoretical con- struction such as we find, say, in Eucl. xm. ; and it is possible that Theaetetus was the first to give these constructions, whether ty paft in Suidas' notice, jrpurov Si to «it< KaXov/uva artpta. typrnjit, means " constructed " or " wrote upon." But cruoracri! in the above phrase of Froclus may well mean something less than the theoretical constructions and proofs of Eucl. xm. ; it may mean, as Vogt says, simply the " putting together " of the figures in the same way as Plato puts them together in the JYmaeus, i.e. by bringing a certain number of 4M PYTHAGORAS AND THE PYTHAGOREANS angles of equilateral triangles and of regular pentagons together at one point. There is no reason why the early Pythagoreans should r.ot have. " constructed " the five regular solids in this sense ; in fact the supposition that they did so agrees well with what we know of their having put angles of certain regular figures together round a point (in connexion with the theorem of Eucl. I. 32) and shown that only three kinds of such angles would fill up the space in one plane round the point. But I do not agree in the apparent refusal of Vogt to credit the Pythagoreans with the knowledge of the theoretical construction of the regular pentagon as we find it in Eucl. iv. 10, 1 1. I do not know of any reason for rejecting the evidence of the Scholia iv, Nos. 2 and 4 which say categorically that " this Book " {Book iv.) and " the whole of the theorems " in it {including therefore Props. 10, n) are discoveries of the Pythagoreans. And the division of a straight line in extreme and mean ratio, on which the construction of the regular pentagon depends, comes in Eucl. Book n. (Prop, n), while we have sufficient grounds for regarding the whole of the substance of Book a. as Pythagorean. I will permit myself one more criticism on Vogt's first paper. I think he bases too much on the fact that it was left for Oenopides (in the period from, say, 470 to 450 B.C.) to discover two elementary constructions (with ruler and compasses only), namely that of a perpendicular to a straight line from an external point (Eucl. 1. 1 2), and that of an angle equal to a given rectilineal angle (Eucl. 1. 23). Vogt infers that geometry must have been in a very rudimentary condition at the time. I do not think this follows ; the explana- tion would seem to be rather that, the restriction of the instruments used in constructions to the ruler and compasses not having been definitely estab- lished before the time when Oenopides wrote, it had not previously occurred to anyone to substitute new constructions based on that principle for others previously in vogue. In the case of the perpendicular, for example, the con- struction would no doubt, in earlier days, have been made by means of a set square. EXCURSUS II. POPULAR NAMES FOR EUCLIDEAN PROPOSITIONS. Although some of these time-honoured names are famijiar to most educated people, it seems to be impossible to trace them to their original sources, or to say who applied them for the first time respectively. It may be that they were handed down by oral tradition for long periods in each case before they found their way into written documents. We begin with I. S . i. This proposition is in this country universally known as the Pons Asinorum, "Asses' Bridge." Even in this case opinion is not unanimous as to the exact implication of the term. Perhaps the more general view is that taken in the Stanford Dictionary of Anglicistd Words and Phrases (by C. A. M. Fennell) where the description is : " Name of the fifth proposition of the first Book of Euclid, suggested by the figure and the difficulties which poor geometricians find in mastering it." This is certainly the equivalent of what I gathered, in my early days at school, from a former Fellow of St John's, the Reverend Anthony Bower, who was a high Wrangler in 1846 and a friend of Todhunter's. The " ass " on this interpretation is a synonym for " fool." But there is another view (as I have learnt lately) which is more complimentary to the ass. It is that, the figure of the proposition being like that of a trestle- bridge, with a ramp at each end which is the more practicable the flatter the figure is drawn, the bridge is such that, while a horse could not surmount the ramp, an ass could ; in other words, the term is meant to refer to the s u re footed - ness of the ass rather than to any want of intelligence on his part. (I may perhaps mention that Sir George Green hill is a strong supporter of this view.) An epigram of 1780 is the earliest reference to the term in Murray's English Dictionary : "If this be rightly called the bridge of asses, He's not the fool that sticks but he that passes." The writer's own view is not too clear. He seems to imply that, while the inventor of the name msant that only the fool finds the bridge difficult to pass, the more proper view would be that, since the ass can get over, and " ass " is synonymous with " fool," therefore it must be the fool who can get over ; in other words, he seems to object to the phrase as being a contradic- tion in terms. But we have also to take account of the fact that the French apply the term to t. 47. Now in Euclid's figure for 1. 47 there is no suggestion of a bridge, and the reference can only be to the nature of the theorem, its diffi- culty or otherwise. It is curious that the French dictionaries give two different explanations of Pont aux ana, Littre makes it "ce que person ne ne doit ni ne peut ignorer ; ce qui est si facile que tout le monde doit y reussir." Now no intelligent person could have applied the name to Eucl. 1. 47 for this reason, namely that it was so easy that even a fool could not help knowing it. Larousse is better informed ; there we find "Pont aux ants, certaine difficulte, certainc question qui n'arrete que les ignorant s, et qui sert de criterium 416 POPULAR NAMES FOR EUCLIDEAN PROPOSITIONS pour juger l'intelligence de quelqu'un, et partieulierement d'un ecolier. C'est ainsi que, dans Its classes de mathematiques, on ne manque jamais dt dire que le carre - de I'hypotenuse est le pant aux dries de la geometric La plupatt des dictionnaires entendent par ce mot une chose si -simple, si facile, que personne ne doit l'ignorer : c'est une erreur e"vidente." Larousse is clearly right. But it will be observed that, so far as it goes, Larousse's interpretation rather supports the first of the two alternative explanations of the meaning of "Asses' Bridge" as applied to I. 5, namely that it is difficult for the fool (= "ass") to master. In the Stanford Die tiff nary it is added that " in logic the term was in the 16 c. applied to the conversion of propositions by the aid of a difficult diagram for finding middle terms"; and if the mathematicians borrowed the term from logic, this again would be rather in favour of the first explanation of its use for 1. 5. If it is permitted deufiere in loco, I would add for the benefit of future generations (in the hope that they will still be able to appreciate the joke or, in the alternative, will be tempted to discuss learnedly what could possibly have been meant) a very topical allusion in a recent Punch (14 Oct. 1935) : "When they film Euclid, as is suggested, we shall no doubt see a very thrilling rescue over the burning Pons Asinorum." — And yet it is safe to prophesy that the " Asses' Bridge " will outlive the " film " I 2. Elefuga. This name for Eucl. 1. 5 is mentioned by Roger Bacon (about 1250), wno also gives an explanation of it (Opus Tertium, c. vi). He observes that in his day people in general, finding no utility in any science such as geometry, for example, recoiled from the idea of studying it unless they were boys farced to it by the rod, so that they would hardly learn so much as three or four pro- positions. Hence it is, he says, that the fifth proposition is called " Elefuga, id est, fuga miserorum ; elegia enim Graece dicitur, Latine miseria ; et elegi sunt miseri." That is, according to Roger Bacon, Elefuga is "flight of the miserable." This explanation no doubt accounts for the verses about Dul- carnon in Chaucer's Troilus and Criseyde, hi, 11. 933-5 : " Dulcarnon called is * fltminge of wrecches'; It secmeth hard, for wrecches wol not lere For venay slouthe or othere wilful tecches"; since "fleminge of wrecches," "banishment of the miserable," is a translation of " fuga miserorum." Only Dulcarnon is there wrongly taken to be the same proposition as Elefuga, i.e. 1. 5, whereas, as we shall see, Dulcarnon was really the name for the Pythagorean theorem 1. 47. Ety mo logically, Roger Bacon's explanation leaves something to be de- sired. The word would really seem to be an attempt to compound the two Greek words iXtos, pity (or the object of pity), and vyrj, flight (cf. note ad loc. in Skeat's edition of Chaucer). Notwithstanding the confusion of tongues, the object seems to be a play upon the two words Elementa and (Aangled [S7 Afupfa, indivisible 41, 168 ;.>■: 0/ division (Aristotle) 165, 170, 17 j : method of division (exhaustion) 185: Euclid's TtplStaipttrtiiHtR.i), 18,87,00 duurdrtif, almost = dimensions 157, 158 ouMTaTi* extended., i$' tv one may, irl Stio two ways, £rl rpta, three ways (of lines, surfaces and solids respectively) r jB, 170 auL(?«a ypa/Af^, perpendicular iftf-9, 371: ''plane' and "solid" 172 ftaAfYtrnJt 1o jca^v^Xot, curved (of lines) 1 50 KaTOOKtw/j, construction, or machinery, one of divisions of a proposition 129; sometimes unnecessary 130 KaraTQv.il xnvbvot, Sectio canonis, attributed to Euclid 17 4 so INDEX OF GREEK WORDS AND FORMS icebrtw, "let it be made" 169 KttitpLfiiti), bent (of lines) 159, 176 Ktrrpar, centre 183, 184, 199 : rr *v roO xirrpou = radius 199 «-we«t*;i (yiarta), hem-like (angle) [77, 17S, 181 rXor, to inflect or dejlett, nexXduOat, kck\h*t- pJrt), itXimr. 1 1 8, 1 jo, 159, 176, 178 ftXXrcf, inclination, 1J0 Koi\trfiiivioy t hollow-angled figure (Zenodoras) 17, 188 jcwai human, Common Notions (=axioms) 131-1 : called also tA jcoo'd, rural $6£m (Aristotle) no, 111 xowi) rpoOKth&w, A&Qpfyrfkt 176 topu^rj, vertex : caraffopi*^^ vertical (angles) 178 Mpiitot, ring (Heron) 1(13 \fyma, lemma ("something assumed, Xs>t- fiari)i«i«r) 133-4 XwrrVr : X«r^ 4 A A Xo<*- jj r£ BII frig iarlv 145 prfpr, parts ( = direction) 190, 308, 313 : (=siae) 171 ufam, length, 158-9 /swrontfo, lune-ltkt (of angle) 3 6, 201 : tA ita-nt-lit (rrrt/w), lune 1B7 /tun-it, "mixed (of lines 01 curves) ifji, 161; (of surfaces) 170 (jorii rpotrXn^oDira SAn>>, definition of a /tain/ ■55 tiarbrrptxpos f\i% "single- turn spiral" 113- 3»., 164-3: in Pappus = cylindrical helii rrvrrett, inclinations, a class or problems 150-1: mfccr, to verge 118, 150 IwrrptinJ-fr, temper-tike (of angle) 178 «>iwi*t)t, "of the same form" ijo 5>UHOT t c1 similar M (ofnumbers)357: (of angles) = equal (Thales, Aristotle) 151 d»i«o*upi$i, uniform (of lines 01 curves) 40, 161-1 «^rra (Ywla), acute (angle) 181 (Jfiryiimu, acute-angled 187 ffnp B«* eVxfoi (or roi^rat) Q.E.D. (or F.) 57 ipfaywnot, right-angled : as used of quadn- latcrats = rectangular 188-9 6pot, Spurpot, definition 143: original mean- ing of Spot 143: m boundary, limit 181 rtytf, visual ray 166 rimj utTa\mfiBaw6iMtnu, " taken together in any manner" 181 TapapoXI) T&y xuptvr, application of areas 36, 343-5; contrasted with bwtppoM) (exceed- ing) and AXtt^ir (faliing-short) 343 : rapa- fi*M contrasted with iriVrMii (construction) 343 : application of terms to conies by Apollonius 344-5 MpdJofot rim, 4, "the Treasury of Para- doxes" 319 rofoK\irra, "fall beside" or "awry" 161 rapaw\^ptaftn t complement, q. V. ripas, extremity 165, 181 1 Wpai airf%hiin* fflasuionius' definition of figure) 183 Ttp&xpiibii (of angle), Ttpaxiueror (of Teet- angle), contained 370 : ri Sit iifxixbturov. cwicethe rectangle contained 380: (of figure) contained ot bounded i8j, 183,1841186, 187 rtpiipipaa, circumference 184 Trcpi&ypawun, contained by a circumference of a circle or by arcs of circles 181, 184 lrXirot, breadth 158-9 n\t ori{w (*pt>fi\Tiiia), "(problem) in excess" 119 ittu, a mathematical instrument 370 ToX&r\eupor t many-Bided figure 187 roptattoffat, to "find" or M furnish" 115 rhpurpn, porism q.v. Tpop\Ttpa. ¥ problem q.v. TpvrryvbyxvM, leading', (of conversion) = oom- filete J56-7 : ipmrfotfatrnr (Smijnj'-n.) leading theorem) contrasted with converse 157 rp6t, in geometry, various meanings 177 Tp&raiTLi, enunciation 149-30 Tporetvta, "propound" 118 rp&rof, prime, two senses of, 146 tt&tu, case 134 fart*, rational 137: farll tii^trpot rBt np> tdicn ("rational diameter of 5") 399 trytitlw, point 155-6 sriSuif, a mathematical instrument 371 rrtyni, point 156 uToixtTor , dement 1 14-6 rrptryy&\. Thibit b. Qurra 88 Abu 'Uthmin ad-Dimasbqi is, 77 Aba '1 Waft al-Burjinl 77, 85, 86 Aba Yusuf Va'tjftb b. Ishiq b- as-$abb*h al- Kindl 86 Aba Yusuf Ya'qfib b. Muh. ar-Rail 86 Adjacent {l$i(fjt). meaning 1S1 Aeuaeas (or Aigeias) of Hierapolis 18, 311 Aganis 17-81 191 Ahmad b. al-Husain al-Ahwiil al-Kalib 89 Ahmad b. 'Umar a]-KarabisT 85 al-Ahwirl 89 Aigeias (?Aenaeas) of Hierapolis )8, 311 Alexander A ph rod is ten sis 7 n., 19 Algebra, geometrical, 371-* : classical method was that of Eucl. 11. (cf. Apollonius) 373 : preferable to semi- algebraical method 377- 8: semi-algebraical method due to Heron 373, and favoured by Pappus 373: geome- trical equivalents of algebraical operations 374 : algebraical equivalents of propositions in Book 11. 371-3 All b. Ahmad Abu ') Qasim al-Antlkl 66 Allman, G. J- 135 «., 318, 351 Alternate (angles) 308 Alternative proofs, interpolated, 58, 59 Amaldi 175, 179-80, 193, 101, 313, 318 Ambiguous case 306-7 Amphinomus 1131 118, 130 n. Amyclas of Heraclea 1 17 Analysis (and synthesis) 18 ; alternative proofs of xtll. i-j by, 137: definitions of, interpolated, 138: described by Pappus 138-9: modem studies of Greek analysis 1 39 : theoretical and problematical analysis 1 38 : Treasury of analysis (rAw amXv4- lifvm) 8, 10, 11, 138: method of analysis and precautions necessary to 139-40: analysis and synthesis of problems 140-1 ; two parts of analysis (a) transformation, (b) resolution , and two parts of synthesis, (a) construction, (b) demonstration 141 : example from Pappus 141-1 : analysis should also reveal hoptaptn (conditions of possibility) 141 Analytical method 36: supposed discovery of, by Plato 134, 137 Anaximander 370 Anchor- ring 163 And run 116 Angle. Curvilineal and rectilineal, Euclid's definition of, 1 76" sq. : definition criticised by Syrianus 176: Aristotle's notion of angle as icMait 1 76 : Apollonius' view of, as amir action 176, 1 77: Plutarch and Carpus on, 177 : to which category does it belong? quantum, Plutarch, Carpus, "A- ganis 177, Euclid 1781 quote, Aristotle and £udemus 177-8: relation, Euclid 178 : Syrianus 1 compromise 1 78 : treatise on the Angle by Eudemus 34, 38, 177-8: classifi- cation of angles (Geminus) 178-9: eurvt- lineal and "mixed" angles 16, 176-9, korn-iikt [KfoaroctSfy) 177, 178, 181, 165, lune-tike (pi)roe<&?t) i6 t 176-9, scraptr-Wu (fuoTpMtSijsj 178: angle ofn segment 1531 angle of a semicircle 181, 153: definitions of angle classified 1 79 : recent Italian views 179-81 : angle as cluster of straight lines or rays iSo-i, defined by Veronese 180: as part of a plane ("angular sector") 179- 80: flat angle (Veronese etc.) 180-1, 169: three kinds of angles, which is prior (Aristotle)? 181-1: adjacent angles 181: alternate 308 : similar ( = equal) 178, 181, 151 : vertical 178: exterior and interior (to a figure) 163, 180: exterior when re- entrant 163: interior and opposite 180: construction by Apollonius of angle equal to angle 196 : angle in a semicircle, theorem of, 317-19: trisect ion of angle, by conchoid of Nicomedes »6$-6, by qundratrix of Hippias 166, by spiral of Archimedes 167 a]- Antaki 86 Antiphon 7*.. 35 422 ENGLISH INDEX ^Anthtsathus" (or "Abthiniathus") 203 Apa&tamba-Sulba- Sutra 352 : evidence in, as to early discovery of EdcL I. 47 and use of gnomon 360-4: Burk's claim that Indians had discovered the irrational 563— 4; approximation to *Ji and Thibaut's explanation 361, 363-4: inaccurate values of w in, 364 Apollodorus 4t Legist icus" 37, 319, 351 Apollonius: disparaged by Pappus in com- parison with Euclid 3 : supposed by some Arabians to be author of the Elements 5 : a "carpenter** 5 : on elementary geometry 44: on the tine 159: on the turtle 176: general definition of diameter 325 : tried to prove axioms 43, 61, 313-3: his "general treatise" 43: constructions by, for bisection of straight line 268, for a perpendicular 170, for an angle equal 1.0 an angle 296: on para I Id -axiom (?) 41-3 : adaptation to conies of theory of application of areas 344-5: geometrical algebra in , 373: Plant Loci 14, 1591330. Planet* form 151 ; com- parison of dodecahedron and icosahedron 6; on the eoc Alias $4, 41, 16* -. on unordered irrationals 43 , 115: 138, 188,211, 313,246, *49> *S9p 37*, 373 Application of areas 36, 343-5 ■ contrasted with exceeding and failing-short 343 : complete method equivalent to geometric solution of mixed quadratic equation 344-5, 383-5, 386-8: adaptation to conies (Apol- lonius) 344-5 : application contrasted with construction (Proclus) 343 "Aqaton" SS Arabian editors and commentators 75-90 Arabic numerals in scholia to Book X-, nth c, , 71 Archibald, R. C 9*1., 10 Archimedes 20, 21, m6, 141: "postulates" in, j 2 Op 1 33 : famous ■ ' lemma " (assumption) known as Postulate of Archimedes 234: "PorismV'in, n a., 13: spiral of, 36, 367 : on straight line 166: on plane 171-2 : 225, U9* 370 Archytas to Areskong, M. E. 113 Arethas, Bishop of Caesarea 48; owned Bodleian MS. (B) 47-8: had famous Plato MS, of Patmos (Cod. Clarkianus) written 48 ArgyTus, Isaak 74 Aristae us 138: on conies 3: Solid Loci 16, 329,: comparison of Ave {regular solid] figures 6 Aristotelian Problems 166, 182, 187 Aristotle: on nature of elements 116: on first principles 117 sqq. : on definitions 117, 1 19-20, [43-41 [46-50: on distinction be- tween hypotheses and definitions 1 19, 120, between hypotheses and postulates 118, 119, between hypotheses and axioms 120: od axioms 119-11 : axioms indemon- strable 121: on definition by negation 156-7 ■ On points 1 55-6, 165: on tines, definitions of 158-9, classification of 159- 60 : quotes Plato's definition of straight line 166: on definitions of surface 170: on the angle 1 76-8 ■ on priority as between right and acute angles 18 [-3: onjSgure and definition of 183-3: definitions of "squaring" 149-50, 410: on parallels 190- »p 308-9: on gn#mon M $$i t 355, 359: on attributes k*t& varr6t and -wpOrrav kw6\ov 319, 310, 325: on the objection 135: on reduction 135: on redwtio ad aosurdutn 136; on the infinite 233-4 1 supposed pos- tulate or axiom about divergent Lines taken by Proclus from, 45, 107 1 gives pre-Eucli- dean proof of l. 5 352-3: on theorem of angle in a semicircle 149: on sum of angles of triangle 31 9-21 : on sum of exterior angles of polygon 322 : 38,45, n?, 150**-, iBl, 184, 185, 187, r88, 195, 102, 303, 331, 232, 33T t 326, 259, 262-3, 283, 4II Arithmetical calculations in scholia to Bk, x. tun al-Arjani, Ibn Rahawaihi 86 Ashkal at-ta'sis 5 n. Asbraf Sbamsaddln as-Samarqandl, Muh. b, 5"*» 89 Astaroff, Ivan 1(3 Asymptotic (non-secant) lines 40, i6i t 203 Athelhard of Bath 78, 93-6 Athena eus 30, 351 Atbenaeus of Cyiicus 117 August, £. F. 103 Austin, W. 103, in Autolycus, On the moving sphere 17 Avicenna 77, 89 Axioms, distinguished from postulates by Amtotte 1 1 8-9, by Proclus (Geminus and "others") 40, 121-3: Proclus on diffi- culties in distinctions 133-4: distinguished from hypotheses, by Aristotle 120-1, by Proclus l*f~4J indemonstrable 111: at- tempt by Apollonius to prove 233-3: = "common {things) " or "common opinions' 3 in Aristotle 120, 33i : common to all sciences 119, 120: called "common notions" in Euclid 121, 221: which are genuine? 33 1 sqq* : Proclus recognises five 332, Heron three 222 : interpolated axioms 234, 233: Pappus* additions to axioms a 5, *33* 224, *3*: axioms of congruence, (1) Euclid's Common Notion 4, 234-7, (2) modern "systems (Fasch, Veronese and Hilbert) 228-31: "axiom" with Stoics = every simple declaratory statement 41, 121 Bacon, Roger 94, 416 Balbus, de mensuris 91 Barbarin 219 Barlaam, arithmetical commentary on Eucl. II. 74 Barrow, 103, roj, 110, 111 Base, meaning 248-9 Basel, editio prineeps of Eucl. 100- 1 Basil ides of Tyre 5, 6 Baudhayana Sulba-Sutra 360 Uayfius (Ba'i'f, Lasare) [00 ENGLISH INDEX 4*3 Becker, J. K. 174 Beet 176 Beha-ad-din 417 Beltrami, E. no Benjamin of Lesbos 113 Bergh, P. 400-1 Bernard, Edward, 103 Besthom and Heiberg, edition of al-Hajjaj's translation and an-Nairlzl's commentary ai, V}n., 79(1. Bhaskira 35 j, 41S Billingsley, Sir Henry 109-10, 418 al-Blrunl 90 Bjiirnbo, Axel Anthon ifn., 03 Boccaccio $6 Bodleian us. (B) 47, 48 Boeeich 331, 371 Boethius 92, 95, 1S4 Bologna MS. (b) 49 Bolyai, I 119 Bolyai, W. 174-5, a'9> 318 Boliano 167 Boncompagni 93 «., 104 n. Bono! a, K. ioj, 119 Borelli, Giovanni Alfonso 10$, 194 Boundary (tyer) 183, 183 Br&kenhjelm, P. R. Ill Breitkopf, I oh. Gottlfeb Immanuel 97 Bretschneider 136(1., 137, 19s, 304, 344, 354. JjS Briconnet. Francois 100 " Bride, Theorem of the," = M.v.t\. 1. 47,417-8 "Bride's Chair," name for I. 47, 417-8 Briggs, Henry 101 Brit- Mm. palimpsest, 7th— 8th c, 30 Bryson 8 m. Burk, A. 35a, 360-4 BUrkten 179 Buteo (Borrel), Johannes 104 Caba&ilas, Nicolaus and Theodorus 74 Caiani, Angelo 101 Camcrarius, Joachim tot Cameier, J. G. 103, 193 Camorano, Rodrigo til Campanus, Johannes, 3, 78, 94-96, 104, too, no, 407 Candalla, Franciscus Flussatea {Francois de Foil, Comte de Candale) 3, 104, no Cantor, Moriti 171, 304, 318, 310, 333, 3SL 3S5, 357-8, 360, 101 Carduchi, L. 111 Carpus, on Astronomy 34, 43: 45, 117, ia8, '77 Cast, technical term 134: cases interpolated J8, 59 Casm 4 b., 1711. Cassiodorus, Magnus Anrelius 91 Caraldi, Pietro Antonio loh Catoplrka, attributed to Euclid, probably Tbeon's 1 7 : Catoptrics of Heron 1 1 , 153 "Cause" : consideration of, omitted by com- mentators 19, 45 : definition should state cause (Aristotle) 149: cause— middle term (Aristotle) 149: question whether geometry should investigate cause (Ge minus), 45, 130*. Censorinus 91 Centre, tirrpor 184-5 Ceria Aristetcliea 35 Chasles on Purisms of Euclid 10, 11, 14, 15 Chaucer: Dttttarium in 416-7, 418 Chinese, knowledge of triangle 3, 4, 5, 3£i " "Chou-pei™ 35J Chou Kung 361 Chrysippus 330 Cicero 91, 351 Circle: definition of, 183-5: = round, vrtttrr -fiXar (Plato) 184: = n^t^inww (Aristotle) 1B4: a plant figure 183-4: centre of, 184-5: pole of, 183: bisected by diameter (T hales) 185, (Saccheri) 185-6: intersections with straight line 137-81 171-4, with another circle 138-40, 341-3, 193-4 Circumference, irrpvpiptui 184 Cissoid i6t, 164, 176, 330 Clairuut 318 Claymundus, Joan. 101 Clavius (Christoph Klau ?) 103, 105, 194, 13»> 38 ■• 39'- 407 Cleonides, Introduction to Harmony 17 Cochliae or roth lion (cylindrical helix) r6l Code* Leideniis 399, t : 11, 17 »., 79 n. Coets, Hendrik 109 Commandinu* 4, 101, 103, 104-5, IO *> Ito > in, 407: scholia included in translation of Elements 73: edited (with Dee) De divisionibus 8, 0, no Commentators on Eucl, criticised by Proclus ■a. ' 6 . « , Common Notions : s= axioms 01 , 1 10- 1,111-1: which ire genuine ? iiisq,: meaning and appropriation of term 11 1 : called " axioms " by Proclus sir CompIemsnt,Trapa.*\HfiiojiAi meaning of, 341 : "about diameter" 341: not necessarily parallelograms 341 : use for application of areas J41-3 Composite, 370 Desargues 193 Describe wt {Arayp&fair dr6) contrasted with (instruct 34 8 De 2olt 31 S Diagonal [Uay vriw) 185 "Diagonal" numbers: see " Side- " and "diagonal^ " numbers Diameter (3uv**rp«) T of circle or parallelo- gram 185 : as applied to figures generally 335: "rational" and " irrational" diameter of 5 (Plato) 399, taken from Pythagoreans 309-400,413 Diels, II. 413 Dimensions (cf. Stafrfatit) 157, 158: Aris- totle's view of, 158-^9 Dinostraths 117, 466 Diodes 164 Diodorus 303 Diogenes Laertius 37, 305, 317, 351 Dionysius, friend of Heron, 4t Diophantus So Diorismtts [6topiCfi&i) — [a) "definition" or ** specification," a formal division of a proposition 119: [b) condition of possibility 118, determines how far solution possible and in how many ways 130-1, 343 : dio- rismi said to have been discovered by Leon Ef6l revealed by analysis 143: in- troduced by Sit dJf 393: first instances in Elements 134, 193 Dippc 108 Direction, as primary notion, discussed 179: direction -theory of parallels 191-1 Distance, iidtrnjaa : = radius 1 99 : in Aristotle has usual general sense and m dim ension 199 Division (method of), Plato's 134 Divisions {o/Jigures) by Euclid 8, 9 : trans- lated by Muhammad al-Bugdadi 8: found (by Woepcke) in Arabic 9, and (by Dee) in Latin translation 8, 9: 110 Dodecahedron in sphere 411 Dodgson, C- L 194, 754, 161, 313 Dou, Jan Fieterszoon 108 Duhamel [39, 318 Duliarnon*, name for Eucl. J. 47, 4 1 6, 418 Egyptians, knowledge of 3* + 4"= 5^ 35a Elbe, Thyra 1 1 3 Elefuga, name for End. 1. 5, 416-7 Elemental pre- Euclidean Elements, by Hip pocratcs of Chios, Leon 1 [6, Theudius 1 17 : contributions to, by Eudoxus 1, 37, Theae- tetus I, 37, Hermotimus of Colophon 117: Euclid's Elements, ultimate aims of 2 , 113-6: commentators on £9-45, Proclui to, 39-45 and passim. Heron 10-S4, an- Niirizi 31-34, Porphyry 34, Pappus 34- 37, Simplicius 38, Aenacas (Aigeias) *8 : mss. of 46-51; Theon*s changes in text 5*~5^ : nieans of comparing Theonine with ante-Theonine text 5J-53: interpolations before Theon's time 5H-63 : scholia 64-74: external sources throwing light on text. Heron, Taurus, Sextus Empiricus, Proclus, lamblichus 61-3: Arabic translations (1) byal-tfajjaj 75, 76, 79, Bo, 83-4, (3) by Ishaq and Thabit b, Qurra 75-80/83-4, (3) NasTraddln at-TusJ 77-80, 84: Hebrew translation by Moses b. Tibbon or Jakob b. Machir 76: Arabian versions compared with Greek text 79-^3, with one another 83, 84 : translation by Boethius 93 : old translation of 10th c*, 91: translation by Athelhard 93-6, Gherard of Cremona 93-4, ENGLISH WDEX 4*5 Campanus 94-6, 07-1 io etc., Zamberti 98-100, Command inn 5 104-5 * introduc- tion into England, toth c, oj : translation by Billi Lesley 109-10: Greek texts, ediiio princcpt too- 1, Gregory's 101-3, Peyrard's 103, August's 103, Heibetg's*af«>B : trans- lations and editions generally 97-1 13: on the nature of tkmtnts (Proclus) 1 14-6, (Menaechmus} 1 1 4. (Aristotle) 116: Proclus on advantages of Euclid's Elements 115: immediate recognition of, 1 16: first princi- ples of, definitions, postulates, common notions (axioms) 117-24: technical terms In connexion with, 135-41 : no definitions of such technical terms 1 50 ; sections of Book 1. 308 EHnitam 05 Engel and Slack el 319, 31 r Enriques, F. 1 r 3, 1 j;, 1 75, r 93, 19 j, 30 1 , 31 3 Enunciation (wpbraats), one of formal di- visions of a proposition 199-30 Epicureans, objection to 1. 10 41, 387; Savile on, 387 Equality, in sense different from that of congruence (-"equivalent," Legendre) 317-8: two senses of equal(i) "divisibly- equal" (Hilbert) or "equivalent by sum" (Amaldi),(i) "equal in content " (Hilbert) ot "equivalent by difference" (Amaldi) 338: modern definition of, 228 Eratosthenes I : contemporary with Archi- medes i, a Errard, Jean, de Bar-Ie-Duc 108 Erycinus 37, 390, 349; Euclid: account of, in Proclus' summary 1 ; date 1-1: allusions to in Archimedes 1: (according to Proclus) a Platonist 3 : taught at Alexandria 3 : Pappus on personality of, 3: story of fin Stobaeus) 3 1 not "of Megaia" 3. 4: supposed to have been bom at Geia 4 : Arabian traditions about, 4, S : "of Tyre" 4-6: "of Tus" 4, 5 «. 1 Arabian derivation of name ("key of geometry ") 6 : Elements, ultimate aim of, a, 115-6: other works, Conks 16, Pttu- dtsria 7, Data 8, 133, 141, 385, 391, On divisions (of figures) 8, 9, Porisms 10-15, Surface-loci 15, 16, Phaenomena 16, 17, Optics 1 7 , Elements of Music or Seeiio Canonis 17: on "three- and four-line locus 3 ' 3: Arabian list of works 17, 18: bibliography 91-113 Eudemus 29: On the Angle 34, 38, 177-8 : History of Geometry 34, 35-8, 578, 395. 304- 317. 3". 387, 411 Eudoxus 1, 37, 74, u(i: discoverer of theory of proportion as expounded generally in Bits, v., vi. 137, 351: on the golden section 137 : founder of method of ex- haustion 334 : inventor of a certain curve, the hippopede, horse* fetter 163 : possibly wrote Sphaerica 17 Euler, Leonhard 401 Eutocias j{, 3J, 39, 141, 161, [64, 159, 317, 3'9. 33o. 373 Exterior and interior (of angles) 163, 280 Extremity, ripea 181, 183 Falk, H. 113 aE-Faradi 8 »,, 90 Figure, as viewed by Plato 181, by Aristotle 182-3, by Euclid 183: according to Posi- donius is confining boundary only 41, 183 : figures bounded by two lines classified 187: angle-less {dywvtov) figure 187 Figures, printing of, 97 Ftkrist 4 »., 5*1., 17, »c, 34, 35, 17 : list of Euclid'a works in 17, 18 Finaeus, Orontius (Oronce Fine) 101, 104 Flauti, Vincenzo 107 Florence MS. Laurent, xxvtii. 3, (F) 47 Flussates, see Cauda] U Forcadel, Pierre 108 Fourier (73-4 Franeisci tunica, " Franciscan's cowl," name for Eucl. 1- 4;, 41S Frankland, W. B. 173, 199 Frischauf r 74 Garti 17 ft. Gauss 171, 193, 194, soi, 119, 311 Geminus: name 38-9: tide of work (#**o- iiXia) quoted from by Proclus 39: ele- ments of astronomy 38: coram, on Posi- donius 39 : Proclus obligations to, 39—42 : on postulates and axioms 121-3 - on theo- rems and problems 1 »8 : two classifications of lines (or curves) 160-1: on homoeo- meric (uniform) lines 162: on "mixed" lines (curves) and surfaces 162: classifica- tion of surfaces 170, of angles 178-9: on parallels 191 : on Postulate 4, 300 : on stages of proof of theorem of 1. 33, 317-M: »7-S, 37, 44, 4S, 74. "33 «■. 303, 365, 330 Geometrical algebra 372-4 1 Euclid's method in Book It, evidently the classical method 373 : preferable to semi-algebraical method „ 377-8 Georgius Fa chyme res 417 Gherard of Cremona, translator of Elements 9 3-4 \ of an- N ai rizi 's com mentary 11 , 1 7 n * , 94: of tract De dtvisionidus 9, ion, Giordano, Vitalc 106, 176 Given, Se*5tvii«T p different senses* 131-3 Gnomon: Literally M that enabling (something) to be known " 64, 370: successive senses of, (1) upright marker of sundial 181,185,171- 1, introduced into Greece by Anjuimander 370, ti) carpenter's square for drawing right angles 371, (3] figure placed round square to make larger square 351, 371, Indian use of gnomon in this sense 361, {4) use extended by Euclid to parallelograms 37 1 * {5) hy Heron and Thean to any figure* 371-1: Euclid's method of denoting in figure 383: arithmetical u& e of, 358-60, 37' "Gnomon- wist h ' (tar A yydipora.), old name for perpendicular (ndfftTot) 50, 161, 171 4i6 ENGLISH INDEX Gotland, A. 133, 134 "Golden section " = section in extreme and mean ratio 1 37 : connexion with theory of irrationals 137 "Goose's foot" [pes anseris), name for Eud. in. 7, 99, 4 1 H Gow, James 1 35 *r. Gracilis, Stephanus 101-1 Grandi, Guido 107 GreenhiU, Sir George 415, 418 Gregory, David 101-3 Gregory of St Vincent 401, 404 Gromaiict 91 n. t 95 Grynaens 100-1 al-Haitham 88, 89 al-Hajjaj h. Yiisuf b. Matar, translator of the Elements 17, 75, 76, 79, 80, 83, 84 Halifax, William ro6, no Halliwell 05 n. Hankel, H. 139, 14L 131. 134. 344. 3S4 Harmonica of Ptolemy, Comm. on, 17 Harmony, Introduction to, not by Euclid r 7 HirOn sr-Rashld 73 al- Hasan h, 'Uliaidallih b. Sulaiman b, Wahb 87 Hauff, J. K. F. to8 ''Heavy and Light," tract on, iS Heiberg, J. L. passim Helix, cylindrical 161, 1G1, 319, 330 Helmholfct 116, 117 Henrici and Treutlein 313, 4O4 Henrion, Denis roB Herigtme, Pierre 108 Herlm, Christian 100 Hermotimus of Colophon r, 1 17 Herodotus 37 »., 370 "Heromidea" i s 8 Heron of Alexandria, meehantevi, date of 10-1 : Heron and Vitruvius so-i ; com- mentary on Euclid's Elements 10-4 : direct proof of I. 73, 301 : comparison of areas of triangles in 1. 14, 334-5 : addi- tion to 1-47, 366-8 : apparently originated semi-algebraical method of proving theo- rems of Book 11, 373, 378: 137 »,, rjg, 163, 168, 1 Jo, rjr-i, rj6, 183, 184, 185, 188, 1 89, 111, 113, 143, 113, 183, 38;, 109, 351. 369, 371, 403, 407, 408 Heron, Proclus instructor 19 " Herandes" 156 Hieronymus of Rhodes 303 Hilbert 157, 193, lot, 118-31, *49. 313. 318 Hipparchus 4 n. , 10, 30 n., 74 n. Hippasus 4IX Hippias of Elis 41, 165-6, 413 Hippocrates of Chios 8 it., 19, 35, 38, 116, ■35. '3*»-. 386-7> 4 '3 Hippopede (fn-ev W£»}), a certain carve used by Eudoxus 161-3, r ?6 Hoffmann, Heinrich 107 Hoffmann, John Jos. Ign. ro8, 365 Holtimann, Wilhelm (Xylander) 107 Bsmoesmerie (uniform) lines 40, 161, 161 Hornlike (angle), masntiHn 177, 178, 181, 165 Horsley, Samuel 106 Hoilel, J- 119 Hudson, John 101 Huttsch, F. 17 »., 74, 319, 400 HuruLin I). labia al-'Ibidt 75 Hypotheses, in Plato 11s: in Aristotle 118- 10 : confused by Proclus with definition* 111-1 : geometer's hypotheses not raise (Aristotle) 119 Hypothetical construction 109 Hypsicles 5: author of Book xiv, 3, 6 lamblichus 63, 83, 417 Ibn al-'Amid 86 Ibn al-Haitham 88, 89 Ibn al-Lubudl 90 Ibn Rahawaihi al-Arjanl 86 Ibn Slna (Avicenna) 7;, 89 "Iflaton" 88 Incomposite (of tines) 160- l , (of surfaces) 170 Indivisible lines (Jto^ku yoau^iat), theory of, rebutted 168 Infinite, Aristotle on the, 131-4 1 infinite division not assumed, but proved, by geo- meters 16S Infinity, parallels meeting at, 191-3 Ingrami, G. 1 75, 193, ins, 101, 117-8 Interior and exterior (of angles) 163, 180: interior and opposite angle 180 Interpolations in the Elements before Theon's time 58-63: by Theon 46, 55-6: 1. 40 interpolated 338 Irrational: discovered with reference to ^1 35l.4H.4H~3 ■ claim of India to priority of discovery 363-4 : " irrational diameter of 5" {Pythagoreans and Plato) 399-400, 413: approximation to Ji by means of "side- and "diagonal- numbers 399- 401 , to J 1 and «/3 in sexagesimal fractions 7+fl-: Indian approximation to -J 1 36 r, 363-4: unordered irrationals (Apollonius) 41, t is : irrational ratio (dpprrror Xayot) 137 Isaacus Monachus for Argyrus) 73-4, 407 Ishiq b. I.lunain b. Ishaq al-'lbadl, Abu Ya qub, translation ol Elements by, 75-80, 83-4 Isma'il b. Bulbul 88 Isoperimetric (or isometric) figures : Pappus and Zenodorus on, 16, 17, 333 isosceles {laoaitt\iri) 1871 of numbers (= even) 188: isosceles right-angled triangle 351 Jakob b. Machir 76 Jan, C. 17 al-Jauhari, al-' Abbas b. Said 85 al-Jayyanl 90 Joannes Pediasimus 71-3 Junge, G-, on attribution of theorem of I. 47 and discovery of irrationals to Pythagoras 35h 41'. 4'3 Kastner, A, G. 78, 97, tot al-Karublsi 85 ENGLISH INDEX 4*7 Kitviyana iiulbaSutra 360 Keill, John lot,, 1 £0-1 1 Kepler 193 al-Khasin, AbQ Jafar 77, 85 Killing, W, 194, 319, a? £-6, 135, 341, 172 al-Kindi j jj„ 86 Flamroth, M. 75-84 Klau (?), Chrisloph = Clavius 105 Kluge), G. S. «» Knesa, Jakob III Knoche jia, 3j«., 73 Kroll, W. 399-400 al-Kuhl 88 Lambert, J. H. 111-3 Lordlier, Dionysius III, 146, J 50, 198, . «•* Laseaxis, Constant inns 3 Leading theorems (as distinct from converse) 157 : leading variety of conversion 156-7 Let lit, John no Leftvre, Jacques 100 Legendre, Adrien Marie ui, 169, 113-9 Leibnii 145, 169, 176, 194 Leiden its. 399, 1 of al-Hsjjij and tut- Nairlzl 11 Lemma 114; meaning 133-4: lemmas inter- polated 59-60, especially from Pappus 67 Leodamas of Thasos 36, 134 Leon 116 Leonardo of Pisa 9 n., 10 Lcucippus 413 Linderup, II. C, 113 Line: Platonic definition 158: objection of Aristotle 158: "magnitude extended one wav" {Aristotle, " Heromides") 158: "divisible or continuous one way" (Aris- totle) 158-9: "flux of point" 159: Apol- lonlus on, if.ii: classification of lines, Plato and Aristotle 159-60, Heron 139-60, Geminus, first classification 160-1, second 161 : straight (tWita), curved (MWfAdt), circular (repi^^t), spiral -shaped MAixo- w*^f), beat (*ixafi)i/nj) h broken (f»ff\a- fitinf), round {rrpayyGXoi) 159, composite (ffOrfartt), incomposite [asrvrQeTo*}, "form- ing a figure" (rxij^ftVorwovffa), determinate (fjipiefiJrTf), indeterminate {iopurrtn) 160 : *' asymptotic" or nan -secant [daOfnrrwTM}, secant (ffv^Mrrwror) 161 : simple, "mixed" 161— 1 : homoeomtrii (uniform) 161-1 : P roc I us on lines without extremities 165 : loci on lints 319, 330 Linear, loci 330: problems 330 Liomardo da Vinci, proof of 1, 47 365-6 Lippert 88 ». Lobachewsky, K. L 174-5, nj, 119 Locus-theorems (rwiira fftvpittsara) and ioci Sritroi) ; locus defined by Proclus 319 : oci likened by Chrysippus to Platonic idea* 330- 1 : locus-theorems and loci ( 1 ) on lines (a) plane loci (straight lines and circles) (b) solid loci (conies), (1) on sur- faces 319 : corresponding distinction be- tween plant and solid problems, to which Pappus adds linear problems 330: further distinction in Pappus between ( 1) t5. 408 Napoleon 103 Nasi rid din at-Tusl 4, s "■• 77> 8 4> 8 9> 108-10 Naslf b. Yumn (Yaman) al-Qass 76, 77, 87 Neide, J. G. C- 103 Nicomachm 91, 417 Nicomedes 43, 160— 1, 165-6 Nipsus, Marcos Junius 305 Nominal and real definitions: see Definitions "Nuptial Number" m Plato's Geometrical Number in Republic 417 Objection (trrrwit), technical term, in geometry 135, 157, 160, 165: in logic (Aristotle) 13 j 0&fon° 61, 151, t88 Otnopides of Chios 34, 36, 116, 171, 195, 37L 4«4 Ofterdinger, L. F. 9 Olympiodorus 39 Oppermann 151 Optics of Euclid 17 Oresme, N. 97 Orontius Finaeus (Oronce Fine) tot, 104 Ozanam, Jaques 107, 108 Paciuolo, Luca 98-91 100, 418 Pamphile 317, 319 Pappus ; contrasts Euclid and Apollonius 3 : on Euclid's Poris/m ro-r4, Surface-loci 15. 16, Dates 8 : on Treasury of Analysis 8, lo, 11, 138: commentary on Elements 14-7, partly preserved in scholia 66: evidence of scholia as to Pappus' text 66-7 1 lemmas in Book x. interpolated from, 67 : on Analysis and Synthesis 1 38-9, 141-1: additional axioms by, ij, 193, 194, 131: on converse of Post. 4 35, sot: proof of I- 3 by, 154: extension of 1. 47 366: semi-algebraical methods in 373, 378: on loci 319, 330: on conchoids 161, 366 : on quadratrix 366 : on isoperimctric figures 16, 17, 333: on paradoxes of Erycinos 17, 390: 30, 39, 133 n., 137, 'S'. "5- 388. 391, 401 Papyrus, Herculanensis No. 106 1 50, 184 ; Oxyrhynchus 50: Fayum 51, 337, 338: Rhind 304, 3ji Paradoxes, in geometry 188 : of Eryrinus 27, 190, 319: an ancient "Budget of Paradoxes 319 Parallelogram { *■ parallelogram mic area), first introduced 335 : rectangular parallelo- gram 370 Parallels: Aristotle on, 190, 191-3: defini tions. l>y "Aganis" 191, by Geminus 191 Posidomus 190, Simplicius 190: as equi' distanls 190-1, 194: direction-theory of, 191-1, 194 : definitions classified 193-4 Veronese's definition and postulate 194: Parallel Postulate, ite Postulate 3: Legendie's attempt to establish theory of 113-9 Paris MSS. of Elements, (p) 49 : (q) 50 Pascb, M. 157, 118, 130 "Peacock's tail," name for III. 8 99, 418 Pediasimus, Joannes 71-3 Feet, T. Eric 331 Peithon 103 Peletarius (Jacques Peletier) 103, 104, 149, 407 Fena 104 Perpendicular (codcro*): definition t8l: "plane" and "solid" 171: perpendicular and obliques 391 Perseus 41, 161-3 Peseh, J. G. van, De Prodi fentibsts 13 sqq., 39 n. Petrus Montaureus {Pierre Mondore) 103 Peyrard and Vatican MS. 190 (P) 46, 47, 103 : 108 Pfleiderer, C. F, 168, 198 Phacnomena of Euclid 16, 17 Phi lip pus of Med ma t, 116 Phillips, George 111 Philo of Byzantium 10, 13: proof of I. 8 163-4 Fhilolaus 34, 331, 371, 399, 411 Philoponus 45, iot -3 Pirekenstein, A. E. Burkh. von 107 Plane (or plane surface): Pinto's definition of, 171 : Produs' and Simplicius' inter- pretation of Euclid's def. 171: possible origin of Euclid's def. 171: Archimedes' assumption 171, 171: other ancient defini- tions of, in Froclus, Heron, Theon of Smyrna, an-NairUI 171-3: "Simson's" definition and Gauss on 171-3: Crelle's tract on, 173-41 other definitions by Fourier 173, Deahni. 174, J. K- Becker 174, Leibniz 176, Beez 176: evolution of, by Bolyai and Lobachewsky 174-51 Enriqucs and Amaldi, Ingrami, Veronese and Hilbert on, 175 "Plane loci" 319-30: Plata Lett of Apol- lonius 14, 159, 330 "Plane problems" 319 Ptanudes, Maximus 71 Plato: 1, 1, 3, 137, ijs-6, IJ9. 184, 187. 103, 331, 411, 417 : supposed invention of Analysis by, 134: def. of straight line 165- 6 : def. of plane surface 171! generation of cosmic figures by putting together triangles 136, 413: rule for rational right-angled triangles 336, 337, 339, 360, 385: "rational diameter of 3 " 399, 413 "Platonic" figures 3, 413-4 Play fair, John 103, 111: " Playfair's " Avium 310: used to prove I. 19, 311, and Eucl. Post. 5, 313: comparison of Axiom with Post. 3, 313-4 Pliny 333 Plutarch 19, 37. '77. 343. 3S>> 4' 7 Point: Pythagorean definition of, 155: inter- ENGLISH INDEX 4*9 pretation of Euclid's definition 155: Plato's view of, and Aristotle's criticism 155-6: attributes of, according to Aristotle 1 56 : terms for {ffriypij, oijfietor) 156: other definitions by " Herundes," Posidonius 156, Simp] [ci us 11}: negative character of Euclid's def, 156: is it sufficient? 156: motion of, produces line 1 57 : an-Nairfil on, r J7 : modem explanations by abstrac- tion 157 Folybius 331 Polygon: sum of interior angles (Proclus 1 proof) 313 : sum of exterior angles 311 **Pens asinorum" 415-6: " Pontaux Ants" ibid. Porism : two senses [3: (1) = corollary 134, 178-9 1 interpolated Ponsms (corollaries) 60-1, 38 r : (1) as used in Porisms of Euclid, distinguished from theorems and problems 10, r t : account of the Perisms given by Pappus ro-13: modem restorations by Simson and Chasles 14: views of Heiberg ri, 14, and Zculben 15 Porphyry 17: commentary on Euclid 34: Symmiita Ha, 34, 44: 136, 17;, 183, 187 Posidonius, the Stoic 30, 17, i8n., 30, 189, 197 : book directed against the Epicurean Zeno 34, 43 : on parallels 40, 100: defini- tion of figure 41, 183 Postulate, distinguished from axiom, by Aristotle nB-9, by Proclus (Gemmus and "others") 1*1-3: from hypothesis, by Aristotle tio-l, by Proclus [31-41 postulates in Archimedes no, 113: Euclid's view of, reconcileable with Aristotle's mo-jo, 114: postulates do not confine us to ruler and compasses 114: Postulate* 1, 1, significance of, 195-0 : famous "Postulate of Archimedes" 134 Postulate 4 : significance of, 100 : proofs of, resting on other postulates r *oo-i, 131* converse true only when angles rectilineal (Pappus) lot Postulate 5 : due to Euclid himself 101 : Proclus on, IM-3 : attempts to prove, Ptolemy SO4-6, Proclus 106-8, Nasiraddln at-Tusl io8-ro, Wallis 11c— 1, Ssccheri 1 1 1 -1 , Lambert 1 1 1 -3 : subst i tu tes for, "Playfair's" axiom (in Proclus) 110, others by Proclus 107, 110, Posidonius and Gemmus 110, Legendre 113, Hi, 110, Wallis 110, Cm hot, Laplace, Loreni, W, Bolyai, Gauss, Worpitiky, Clairaut, Veronese, Ingrami 110 1 Post, j proved from, and compared with, "Playfair's" axiom 313-4: 1. 30 is logical equivalent of, 110 Foils, Robert 111, 146 Prime (of numbers): two senses of, 146 Principles, First n 7- (14 Problem, distinguished from theorem 114-8: problems classified according to number of solutions (a) one solution, ordered (Tmry- iti'D.) {i) a definite number, intermediate fjUta) f>) an infinite number of solutions, unordered (iraim) 118 : in widest sense anything propounded (possible or not) but generally a construction which is possible 118-9: another classification (1} problem in excess (rXfordfor), asking too much 119, (1) deficient problem (AXtWr wp6fi\ifua) t giving too little 119 Proclus : details of career 19-30 : remarks on earlier commentators 19, 33, 45 ; com- mentary on End. I, sources of, 19-45, object and character of, 31-1 : com- mentary probably not continued, though continuation intended 31-3 : books quoted by name in, 34 : famous " sum- mary " 37-8 : list of writers quoted 44 : his own contributions 44-5 : character of MS- used by, 61, 63 : on the nature of elements and things elementary 1 14-6 : on advantages of Euclid's Elements, and their object 115-6: on first principles, hypotheses, postulates, axioms 111-4: on difficulties in three distinctions between postulates and axioms 113 : on theorems and problems 124-9 : attempt to prove Postulate 5 106-8 : on Eucl. I. 47, 350, 411: on discovery of five regular solids 413: commentary on Plato's Republic, allusion in, to "side-" and "diagonal-" numbers in connexion with Eucl. 11. 9, 10 399-400 Proof ' fdwAotifir), necessary part of pro- position 110-30 Proposition t formal divisions of, 119-131 Protarchus 5 Psellui, Michael, scholia by, 70, 71 Pstudaria of Euclid 7: Psettaographtmataln. Pseudoboethius 91 Ptolemy I. : I, 1: story of Euclid and Ptolemy 1 Ptolemy II. Philadelphia 10 Ptolemy VII. (Euergetes II.), Physcon ic Ptolemy, Claudius 11, 30 n.\ Harmonica of , and commentary on 17: an Parallel-Pos- tulate 18*., 34, 43, 45 : attempt to prove it 104-6 Punch on "Pens Asinoruin" 416 Pythagoras 4 »., 36 : supposed discoverer of the irrational 351, 41 1, 411, of application 0/ areas 343-4, of theorem of I. 47 343-4, 350-4, 411. 41s, of construction of five regular solids 413-4; story of sacrifice 37, 3+3, 350 : probable method of discovery of I. 47 and proof of, 351-5 : suggestions by Bretschneider and Ilankel 3 J 4, by Zeuthen 355-6: rule for forming right-angled tri- angles in rational numbers 331, 356-9, 383 Pythagoreans 10, 36, 155, 188, 179, 411-414 : term for surface (xpout) 169: angles of tri- angle equal to two right angles, theorem and proof 3 17-10: three polygons which in contact fill space round point 318 : method of application of areas (including exceeding and falling short) 343, 3B4, 403 : gnomon Pythagorean 351 : "rational " and "ir- rational diameter of 5" 399-400, 413 43° ENGLISH INDEX Qirllzade ar-Rumi j«., 90 Q.E.D. (or F.) 57 al-Qifli ^n., 94 Quadratic eq nation, geometrical solution of, 383-5, 386-8 : solution assumed by Hippo- crates 386-7 Quadrserix 365-6, 330 Quadrature [TeTpaywvtrfi&i), definitions of, '♦? . . Quadrilaterals, varieties of, 188-90 Quintilian 333 Qusta b- Luqa al-Ba'labakkl, translator of "Books xiv, xv" 76, 87, 88 Radius, no Greek word for, 190 Ramus, Petals (Pierre de la Ramee) 104 Ratdolt, Erhard 78, 97 Rational {far&t) : (of ratios) 137 : " rational diameter of 5 "399-400, 413: rational right-angled triangles, see right-angled triangles Rauchfuss, tee Dasypodius Rausenberger, O. 157, 175, 313 ar-RazI, Abu Yusuf Ya'qub b. Muh. 86 Rectangle : = rectangular parallelogram 370: "rectangle contained by" 370 Rectilineal angle: definitions classified 179— 81: rectilineal figure 187: "rectilineal segment" 196 Rtductio ad absurdum 1 34 : described by Aristotle and Prochts r3 386-7 ; on Eademufi 1 style 35 , 38 : on parallel* 1 90-1*: «, 167, 17 J, 184, 185, ig;, 103, «3* 114, 413 Siinson. Robert: on Euclid's Pviismi 14: ENGLISH INDEX 431 on M vitiations * in Elements due to Theon 46, 103, 104, 106, in, [48: definition of plane 171-3: 185, rS6, 355, 150, 287, *93- *9*- 3"> l* 8 * &i* 3 8 7i 4<»3 Sind b. All Abu \Tuyib 86 Smith, D. E, 363/417 Smith and Bryant 404 ■* Solid loci " 3191 330 : Solid Loci of Aris- tarns 1 -ft, 319 "Solid problems" 319, 330 Speusippug 115 .S/Aamfd, early treatise on, 17 Spiral, "single-turn" 131-3*1., 164- j: in Pappus = eylindrical helix 165 Spiral of Archimedes 16, 167 Spire (tore) or JT/iWf surface 163, 170; varieties of 163 Spine curves or sections, discovered by Perseus 161, 161-4 Steenstra, Pybo 109 Steiner, Jakob 193 StcinmeU, MoriU toi St ein&chn eider, M. 8«., 76 sqq. Stephaiius Gracilis joi-i Stephen Clericus 47 Stobaeus 3 Stoic ♦'axioms 11 41, 111 : illustrations [B*ty- fnn\ 319 StoU, 0. 35 8 Stone, E. 105 Straight line : pre- Euclidean (Platonic) de- finition 165-6 h Archimedes* assumption respecting, 166; Euclid's definition, inter- preted by Proclus and SimpLicius 166-7 : language and construction of, 167, and conjecture as to origin 168 J other defj' ni lions 168-9, m Heron 1 68, by Leib- niz 169^ by Legend re (69: two straight lines cannot enclose a space 10,5-6* can* not have a common segment 196-9 : one or two cannot make a figure 169, 183; division of straight line into any number of equal parts (an-Nairlzl) 326 Stromer, Marten 113 Studemnnd, W. 92 n. St Vincent, Gregory of, 401, 404 Subtend, meaning and const ruction 249 , 183, 350 Suidas 370 Sulaiman b. 'Usma (or Oqba) 85, 90 Superposition: Euclid's dislike of method of, 235, 349 : apparently assumed by Aris- totleas legitimate 216: used by Archimedes 335 : objected to by Peletarius 349: no use theoretically, but merely furnishes practical test of equality 117 : Bertrand Russell on, «7j *49 Surface : Pythagorean term for, xpoid { = col- our, or skin) 169: terms for, in Plato and Aristotle 169; iri^dwtta in Euclid (not i-KlTrtSor) 169: alternative definition of, in Aristotle 170: produced by motion of line 170: divisions or sections of solids are surfaces 170, 171: classifications of surfaces by Heron and Gemmu* 170: com- posite, incomposite, simple, mixed 170: spin) surfaces 163, 170: homoeomeric (uniform) 170: spheroids 170: plane sur- face, see plane: loci on surfaces 319, 330 Surface-loci of Euclid 15, 16, 330: Pappus' lemmas on, 15, 16 Suter, H. 8 tt., 17 w., 18 71., 35 «,, 78 n., 85-90 Suvoroff, Fr, 113 Swmden, J. II. van 169 Synthesis, see Analysis and Synthesis Syrianus 30, 44, 176, 178 Tacquet, Andre' 103, iq;, in Taittirlya-Samhita 363 Tannery, P. 7 «♦, 37-40, 44* *** J *3i "■» 113, 114, 125, 132, 305, 353, 412, 417 TWrikh al-HuktiMtl 4 n. TartagUa, Niccolo 3, 103, 106 Taurinus, F, A, 219 Taurus 62, J 84 Taylor, H. M. 248, 377-8, 404 Taylor, Th- 259 Thabit b. Qurra, translator of Elements 42, 75-80, 82, 84, 87, 94 : proof of I. 47 3^4-5 Thales 36, 37, 185, 252, 253, 278, 3i7t 318, 319 : on distance of ship from snore 3=>+- 5 Tbeaetetus 1, 37 Theodorus Antiochita 71 Theodorus Cabas i las 72 Theodorus Metochita 3 Theodorus of Cyrene 4 n, 412-3 Theognis 371 Theon of Alexandria: edition of Elements 46 : changes made by, 46 ; Simson on "vitiations" by, 46: principles for detect- ing his alterations, by comparison of P, ancient papyri and " Theomne" ms$, 51- 3: character of changes by, 54-8 Theon of Smyrna in, 357, 358, 371, 39 s Theorem and problem, distinguished by Speusippus 125, Amphinomus 125, 128, Menaechmus 125, Zenodotus, Fosidonius 1 16. Euclid r 26, Carpus 137, 128: views of Proclus 127-8, and of Geminus 128: "general 11 and " not -general" (or partial) theorems (Proclus) 325 Theudius of Magnesia II Jf Tbibaut, B- F. a*' Thibaut, C- : On Sulvasutras 360, 363-4 Thompson, Thomas Perronet 112 Thucydides 333 Tibbon, Moses b. 76 Tiraboschi 94 «. Title I, K, 38, 39 Todhunter, I, J12 189, 2461 258, 377, 283, 293, 298, 307 Tonstall, Cuthbert 100 Tore 163 Transformation of areas 346-7, 410 Trapezium 1 Euclid's definition his own 189: further division into trapezia and trape- zoids (Posidonius, Heron) 189-90 ; a 43* ENGLISH INDEX theorem on area of parallel -trapezium 338-9 Trtasury tf Attatyat (AttLkxibiuoat Ttaror} 8, 10. 1 [, 138 Trendelenburg 1461*., 1 48, 149 Treullein, P. 358-6° Triangle: seven species of, 188: "four- sided " triangle, called alto "barb- like" {&juioit&4i) and (by Zenodorus) xoikoytlt- vtor 17, 188 : construction of isosceles and scalene triangles 343 Trisection of an angle 165-7 at-Tusl, lee Nasiraddin Unger, E. S. 108, 169 Vacca, Giovanni 113 Vachtchenko-Zakhartchenko it] Vailati, G. 144 n., Mi". Valerius Maiimus 3 Valla, G., Dt expdendii et fugitndii rtbtu 73. 98 Van Swinden 169 Vatican MS. 190 (P) 46, 47 Vaui, Cam de 10 Verona palimpsest 91 Veronese, G, 157, 168, 175, 180, 193-+, 195, 101, u(4-7, "8, 149, 318 Vtrtxal (angles) 178 Viennese MS, (V) 48, 49 Vinci, Llonando da 365-6 Vitali, G. 13; Vitruvius 351 ; Vitruvius and Heron 10, 11 Viviani. Vincenzo 107, 401 Vogt, Heinrich 360, 364, 411-414 Vooght, C. J. 108 Wachsmuth, C. 31 n., 73 Wallis, John 103: edited Coram, on Ptol- emy's Harmenua 17 : attempt to prove Post. 5 iro-i Weber (H.) and Wellstein (J.) 157 Weissenborn, H. 78 n., 91 »., 94 s., 95, 96. 97 *■> +'8 Whiston, W. Ill Williamson, James in, 193 Witt, H. A. 113 Woepcke, F., discovered De divistottibus in Arabic and published translation 9: on Pappus' commentary on Elcmmlt 15, 66, 77 : 8s «., 86, 87 Xenocraies 168, 413 Ximenes, Leonardo 107 Xylan der 107 Yahya b. Khalid b. Barmak 75 Yahya b. Muh. b. 'Abdan b. Abdalwahid (tbn a]-Lubudl) 90 Yrinus = Heron 11 Yuhanni b. Yusuf b. al-Haritb b. el-Bitrtq al-Qaas 76, 87 Zamberti, Bartolomeo 98-100, 101, 104, 106 Zeno the Epicurean 34, 196, 197, 199, 141 Zenodorus j6, 17, 1 88, 333 Zenodotus 116 Zeuthen, H. G. 15, 1(3, 139, 141, 146ft., IS'. 35S-*. ite- 3*3- 3 8 ;, 39". 399