Euclid's Elements - 300BC Volume 1-9
Euclid's Elements - 300BC Volume 1-9
Euclid's Elements - 300BC Volume 1-9
Introduction
Euclid's Elements form one of the most beautiful and influential works of science in the history of
humankind. Its beauty lies in its logical development of geometry and other branches of mathematics. It
has influenced all branches of science but none so much as mathematics and the exact sciences. The
Elements have been studied 24 centuries in many languages starting, of course, in the original Greek,
then in Arabic, Latin, and many modern languages.
I'm creating this version of Euclid's Elements for a couple of reasons. The main one is to rekindle an
interest in the Elements, and the web is a great way to do that. Another reason is to show how Java
applets can be used to illustrate geometry. That also helps to bring the Elements alive.
The text of all 13 Books is complete, and all of the figures are illustrated using the Geometry Applet,
even those in the last three books on solid geometry that are three-dimensional. I still have a lot to write
in the guide sections and that will keep me busy for quite a while.
This edition of Euclid's Elements uses a Java applet called the Geometry Applet to illustrate the
diagrams. If you enable Java on your browser, then you'll be able to dynamically change the diagrams.
In order to see how, please read Using the Geometry Applet before moving on to the Table of Contents.
Table of Contents
Prematter
Introduction
Book I. The fundamentals of geometry: theories Book VII. Fundamentals of number theory.
of triangles, parallels, and area. Definitions (22)
Definitions (23) Propositions (39)
Postulates (5)
Common Notions (5) Book VIII. Continued proportions in number
Propositions (48) theory.
Propositions (27)
Book II. Geometric algebra.
Definitions (2) Book IX. Number theory.
Propositions (13) Propositions (36)
Table of contents
● Definitions (23)
● Postulates (5)
● Common Notions (5)
● Propositions (48)
● Guide to Book I
Definitions
Definition 1.
A point is that which has no part.
Definition 2.
A line is breadthless length.
Definition 3.
The ends of a line are points.
Definition 4.
A straight line is a line which lies evenly with the points on itself.
Definition 5.
A surface is that which has length and breadth only.
Definition 6.
The edges of a surface are lines.
Definition 7.
A plane surface is a surface which lies evenly with the straight lines on itself.
Definition 8.
A plane angle is the inclination to one another of two lines in a plane which meet one another
and do not lie in a straight line.
Definition 9.
And when the lines containing the angle are straight, the angle is called rectilinear.
Definition 10.
When a straight line standing 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 called a
perpendicular to that on which it stands.
Definition 11.
An obtuse angle is an angle greater than a right angle.
Definition 12.
An acute angle is an angle less than a right angle.
Definition 13.
A boundary is that which is an extremity of anything.
Definition 14.
A figure is that which is contained by any boundary or boundaries.
Definition 15.
A circle is a plane figure contained by one line such that all the straight lines falling upon it from
one point among those lying within the figure equal one another.
Definition 16.
And the point is called the center of the circle.
Definition 17.
A diameter of the circle is any straight line drawn through the center and terminated in both
directions by the circumference of the circle, and such a straight line also bisects the circle.
Definition 18.
A semicircle is the figure contained by the diameter and the circumference cut off by it. And the
center of the semicircle is the same as that of the circle.
Definition 19.
Rectilinear figures are those which are contained by straight lines, trilateral figures being those
contained by three, quadrilateral those contained by four, and multilateral those contained by
more than four straight lines.
Definition 20.
Of trilateral figures, an equilateral triangle is that which has its three sides equal, an isosceles
triangle that which has two of its sides alone equal, and a scalene triangle that which has its three
sides unequal.
Definition 21.
Further, of trilateral figures, a right-angled triangle is that which has a right angle, an obtuse-
angled triangle that which has an obtuse angle, and an acute-angled triangle that which has its
three angles acute.
Definition 22.
Of quadrilateral figures, a square is that which is both equilateral and right-angled; an oblong that
which is right-angled but not equilateral; a rhombus that which is equilateral but not right-angled;
and a rhomboid that which has its opposite sides and angles equal to one another but is neither
equilateral nor right-angled. And let quadrilaterals other than these be called trapezia.
Definition 23
Parallel straight lines are straight lines which, being in the same plane and being produced
indefinitely in both directions, do not meet one another in either direction.
Postulates
Postulate 1.
To draw a straight line from any point to any point.
Postulate 2.
To produce a finite straight line continuously in a straight line.
Postulate 3.
To describe a circle with any center and radius.
Postulate 4.
That all right angles equal one another.
Postulate 5.
That, if a straight line falling on two straight lines makes the interior angles on the same side less
than two right angles, the two straight lines, if produced indefinitely, meet on that side on which
are the angles less than the two right angles.
Common Notions
Common notion 1.
Things which equal the same thing also equal one another.
Common notion 2.
If equals are added to equals, then the wholes are equal.
Common notion 3.
If equals are subtracted from equals, then the remainders are equal.
Common notion 4.
Things which coincide with one another equal one another.
Common notion 5.
The whole is greater than the part.
Propositions
Proposition 1.
To construct an equilateral triangle on a given finite straight line.
Proposition 2.
To place a straight line equal to a given straight line with one end at a given point.
Proposition 3.
To cut off from the greater of two given unequal straight lines a straight line equal to the less.
Proposition 4.
If two triangles have two sides equal to two sides respectively, and have the angles contained by
the equal straight lines equal, then they also have the base equal to the base, the triangle equals
the triangle, and the remaining angles equal the remaining angles respectively, namely those
opposite the equal sides.
Proposition 5.
In isosceles triangles the angles at the base equal one another, and, if the equal straight lines are
produced further, then the angles under the base equal one another.
Proposition 6.
If in a triangle two angles equal one another, then the sides opposite the equal angles also equal
one another.
Proposition 7.
Given two straight lines constructed from the ends of a straight line and meeting in a point, there
cannot be constructed from the ends of the same straight line, and on the same side of it, two
other straight lines meeting in another point and equal to the former two respectively, namely
each equal to that from the same end.
Proposition 8.
If two triangles have the two sides equal to two sides respectively, and also have the base equal to
the base, then they also have the angles equal which are contained by the equal straight lines.
Proposition 9.
To bisect a given rectilinear angle.
Proposition 10.
To bisect a given finite straight line.
Proposition 11.
To draw a straight line at right angles to a given straight line from a given point on it.
Proposition 12.
To draw a straight line perpendicular to a given infinite straight line from a given point not on it.
Proposition 13.
If a straight line stands on a straight line, then it makes either two right angles or angles whose
sum equals two right angles.
Proposition 14.
If with any straight line, and at a point on it, two straight lines not lying on the same side make
the sum of the adjacent angles equal to two right angles, then the two straight lines are in a
straight line with one another.
Proposition 15.
If two straight lines cut one another, then they make the vertical angles equal to one another.
Corollary. If two straight lines cut one another, then they will make the angles at the point of
section equal to four right angles.
Proposition 16.
In any triangle, if one of the sides is produced, then the exterior angle is greater than either of the
interior and opposite angles.
Proposition 17.
In any triangle the sum of any two angles is less than two right angles.
Proposition 18.
In any triangle the angle opposite the greater side is greater.
Proposition 19.
In any triangle the side opposite the greater angle is greater.
Proposition 20.
In any triangle the sum of any two sides is greater than the remaining one.
Proposition 21.
If from the ends of one of the sides of a triangle two straight lines are constructed meeting within
the triangle, then the sum of the straight lines so constructed is less than the sum of the remaining
two sides of the triangle, but the constructed straight lines contain a greater angle than the angle
contained by the remaining two sides.
Proposition 22.
To construct a triangle out of three straight lines which equal three given straight lines: thus it is
necessary that the sum of any two of the straight lines should be greater than the remaining one.
Proposition 23.
To construct a rectilinear angle equal to a given rectilinear angle on a given straight line and at a
point on it.
Proposition 24.
If two triangles have two sides equal to two sides respectively, but have one of the angles
contained by the equal straight lines greater than the other, then they also have the base greater
than the base.
Proposition 25.
If two triangles have two sides equal to two sides respectively, but have the base greater than the
base, then they also have the one of the angles contained by the equal straight lines greater than
the other.
Proposition 26.
If two triangles have two angles equal to two angles respectively, and one side equal to one side,
namely, either the side adjoining the equal angles, or that opposite one of the equal angles, then
the remaining sides equal the remaining sides and the remaining angle equals the remaining
angle.
Proposition 27.
If a straight line falling on two straight lines makes the alternate angles equal to one another, then
the straight lines are parallel to one another.
Proposition 28.
If a straight line falling on two straight lines makes the exterior angle equal to the interior and
opposite angle on the same side, or the sum of the interior angles on the same side equal to two
right angles, then the straight lines are parallel to one another.
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 sum of the interior angles on
the same side equal to two right angles.
Proposition 30.
Straight lines parallel to the same straight line are also parallel to one another.
Proposition 31.
To draw a straight line through a given point parallel to a given straight line.
Proposition 32.
In any triangle, if one of the sides is produced, then the exterior angle equals the sum of the two
interior and opposite angles, and the sum of the three interior angles of the triangle equals two
right angles.
Proposition 33.
Straight lines which join the ends of equal and parallel straight lines in the same directions are
themselves equal and parallel.
Proposition 34.
In parallelogrammic areas the opposite sides and angles equal one another, and the diameter
bisects the areas.
Proposition 35.
Parallelograms which are on the same base and in the same parallels equal one another.
Proposition 36.
Parallelograms which are on equal bases and in the same parallels equal one another.
Proposition 37.
Triangles which are on the same base and in the same parallels equal one another.
Proposition 38.
Triangles which are on equal bases and in the same parallels equal one another.
Proposition 39.
Equal triangles which are on the same base and on the same side are also in the same parallels.
Proposition 40.
Equal triangles which are on equal bases and on the same side are also in the same parallels.
Proposition 41.
If a parallelogram has the same base with a triangle and is in the same parallels, then the
parallelogram is double the triangle.
Proposition 42.
To construct a parallelogram equal to a given triangle in a given rectilinear angle.
Proposition 43.
In any parallelogram the complements of the parallelograms about the diameter equal one
another.
Proposition 44.
To a given straight line in a given rectilinear angle, to apply a parallelogram equal to a given
triangle.
Proposition 45.
To construct a parallelogram equal to a given rectilinear figure in a given rectilinear angle.
Proposition 46.
To describe a square on a given straight line.
Proposition 47.
In right-angled triangles the square on the side opposite the right angle equals the sum of the
squares on the sides containing the right angle.
Proposition 48.
If in a triangle the square on one of the sides equals the sum of the squares on the remaining two
sides of the triangle, then the angle contained by the remaining two sides of the triangle is right.
The Elements begins with a list of definitions. Some of these indicate little more than certain concepts
will be discussed, such as Def.I.1, Def.I.2, and Def.I.5, which introduce the terms point, line, and
surface. (Note that for Euclid, the concept of line includes curved lines.) Others are substantial
definitions which actually describe new concepts in terms of old ones. For example, Def.I.10 defines a
right angle as one of two equal adjacent angles made when one straight line meets another. Other
definitions look like they're substantial, but actually are not. For instance, Def.I.4 says a straight line "is
a line which lies evenly with the points on itself." No where in the Elements is the defining phrase
"which lies evenly with the points on itself" applicable. Thus, this definition indicates, at most, that
some lines under discussion will be straight lines.
It has been suggested that the definitions were added to the Elements sometime after Euclid wrote them.
Another possibility is that they are actually from a different work, perhaps older. In Def.I.22 special
kinds of quadrilaterals are defined including square, oblong (a rectangle that are not squares), rhombus
(equilateral but not a square), and rhomboid (parallelogram but not a rhombus). Except for squares,
these other shapes are not mentioned in the Elements. Euclid does use parallelograms, but they're not
defined in this definition. Also, the exclusive nature of some of these terms—the part that indicates not a
square—is contrary to Euclid's practice of accepting squares and rectangles as kinds of parallelograms.
Following the list of definitions is a list of postulates. Each postulate is an axiom—which means a
statement which is accepted without proof— specific to the subject matter, in this case, plane geometry.
Most of them are constructions. For instance, Post.I.1 says a straight line can be drawn between two
points, and Post.I.3 says a circle can be drawn given a specified point to be the center and another point
to be on the circumference. The fourth postulate, Post.I.4, is not a constuction, but says that all right
angles are equal.
The Common Notions are also axioms, but they refer to magnitudes of various kinds. The kind of
magnitude that appears most frequently is that of straight line. Other important kinds are rectilinear
angles and areas (plane figures). Later books include other kinds.
In proposition III.16 (but nowhere else) angles with curved sides are compared with rectilinear angles
which shows that rectilinear angles are to be considered as a special kind of plane angle. That agrees
with Euclid's definition of them in I.Def.9 and I.Def.8.
Also in Book III, parts of circumferences of circles, that is, arcs, appear as magnitudes. Only arcs of
equal circles can be compared or added, so arcs of equal circles comprise a kind of magnitude, while
arcs of unequal circles are magnitudes of different kinds. These kinds are all different from straight
lines. Whereas areas of figures are comparable, different kinds of curves are not.
Book V includes the general theory of ratios. No particular kind of magnitude is specified in that book.
It may come as a surprise that ratios do not themselves form a kind of magnitude since they can be
compared, but they cannot be added. See the guide on Book V for more information.
Number theory is treated in Books VII through IX. It could be considered that numbers form a kind of
magnitude as pointed out by Aristotle.
Beginning in Book XI, solids are considered, and they form the last kind of magnitude discussed in the
Elements.
The propositions
Following the definitions, postulates, and common notions, there are 48 propositions. Each of these
propositions includes a statement followed by a proof of the statement. Each statement of the proof is
logically justified by a definition, postulate, common notion, or an earlier proposition that has already
been proven. There are gaps in the logic of some of the proofs, and these are mentioned in the
commenaries after the propositions. Also included in the proof is a diagram illustrating the proof.
Some of the propositions are constructions. A construction depends, ultimately, on the constructive
postulates about drawing lines and circles. The first part of a proof for a constuctive proposition is how
to perform the construction. The rest of the proof (usually the longer part), shows that the proposed
construction actually satisfies the goal of the proposition. In the list of propositions in each book, the
constructions are displayed in red.
Most of the propositions, however, are not constructions. Their statements say that under certain
conditions, certain other conditions logically follow. For example, Prop.I.5 says that if a triangle has the
property that two of its sides are equal, then it follows that the angles opposite these sides (called the
"base angles") are also equal. Even the propositions that are not constructions may have constructions
included in their proofs since auxillary lines or circles may be needed in the explanation. But the bulk of
the proof is, as for the constructive propositions, a sequence of statements that are logically justified and
which culminates in the statement of the proposition.
The various postulates and common notions are frequently used in Book I. Only two of the propositions
rely solely on the postulates and axioms, namely, I.1 and I.4. The logical chains of propositions in Book
I are longer than in the other books; there are long sequences of propositions each relying on the
previous.
1 2 3
3, 4 5, 6
5 7 8
1, 3, 8 9, 11
1, 4, 9 10
8, 10 12
11 13 14, 15
3, 4, 10, 15 16 27
13, 16 17
3, 5, 16 18
5, 18 19
3, 5, 19 20
16, 20 21
3, 20 22
8, 22 23
3, 4, 5, 19, 23 24
4, 24 25
3, 4, 16 26
29 30
23, 27 31
13, 29, 31 32
4, 27, 29 33
4, 26, 29 34 43
4, 29, 34 35
33, 34, 35 36
31, 34, 35 37
31, 34, 36 38
31, 37 39
31, 38 40
34, 37 41
3, 8, 11, 47 48
Select topic
© 1996
D.E.Joyce
Clark University
Definition 1
A point is that which has no part.
The Elements is the prime example of an axiomatic system from the ancient world. Its form has shaped
centuries of mathematics. An axiomatic system should begin with a list of the terms that it will use. This
definition says that one term that will be used is that of point. The next few definitions give some more
terms that will be used. Although there is some description to go along with the terms, that description is
actually never used in the exposition of the axiomatic system. It can, at most, be used to orient the
reader.
The description of a point, "that which has no part," indicates that Euclid will be treating a point as
having no width, length, or breadth, but as an indivisible location.
Later definitions will define terms by means of terms defined before them, but the first few terms in the
Elements are not defined by means of other terms; they're "primitive" terms. Their meaning comes from
properties about them that are assumed later in axioms. In the Elements, the axioms come in two kinds:
postulates and common notions. The first postulate, I.Post.1, for instance, gives some meaning to the
term "point." It states that a straight line may be drawn between any two points. Other postulates add
more meaning to the term "point."
Actually, Euclid failed to notice that he made a number of conclusions without complete justification at
a number of places in the Elements. This usually means that a postulate, that is, a explicit assumption, is
missing.
© 1996
D.E.Joyce
Clark University
Definition 2
A line is breadthless length.
"Line" is the second primitive term in the Elements. The description, "breadthless length," says that a
line will have one dimension, length, but it won't have breadth or depth. In I.Def.5 a surface is defined
with the two dimensions length and breadth, and in XI.Def.1 a solid is defined with the three dimensions
length, breadth, and depth.
One cannot tell from this definition what kind of line is meant by "line," but later a "straight" line
defined to be a special kind of line. One can conclude, then, that "lines" need not be straight. Perhaps
"curve" would be a better translation than "line" since Euclid meant what is commonly called a curve in
modern English, where a curve may or may not be straight.
Also, from the next definition, it is apparent that Euclid's lines may have ends, so they are "line
segments" or "curve segments." But they need not have ends in all cases since the entire circumference
of a circle is an example of a line. Indeed, lines need not be finite in all cases; there are a few instances
in the Elements where a line is not bounded, and that is usually indicated by the language. See, for
example, proposition I.12.
One piece of terminology that Euclid did not mention explicitly in a definition is a phrase to indicate
when a line passes through a point. That would be a "primitive" relation that could hold between a line
and a point. Postulates would be included as well to give meaning to the phrase as they are in modern
treatments of elementary geometry.
Book I introduction
© 1996
D.E.Joyce
Clark University
Definition 3
The ends of a line are points.
This statement can be taken as indicating that between certain lines and points a relation holds, that a
point can be an end of a line. It doesn't say what ends are. It also doesn't indicate how many ends a line
can have. For instance, the circumference of a circle has no ends, but a finite line has its two end points.
Book I introduction
© 1996
D.E.Joyce
Clark University
http://aleph0.clarku.edu/~djoyce/java/elements/bookI/defI3.html5/25/2008 8:55:25 AM
Euclid's Elements, Book I, Definition 4
Definition 4
A straight line is a line which lies evenly with the points on itself.
This statement indicates, at least, that the term "straight line" refers to a kind of line. It is hard to tell
what else it means, if anything. Various commentators have interpreted in a variety of ways. The
definition of plane surface in I.Def.7 uses a similar language that is equally opaque.
There are a some postulates that come a little later in Book I and give meaning to straight lines. I.Post.1
says that a straight line can be drawn between any two points, I.Post.2 says that a straight line can be
extended, and the remaining postulates use the concept of straight line in one way or another.
Book I introduction
© 1996
D.E.Joyce
Clark University
http://aleph0.clarku.edu/~djoyce/java/elements/bookI/defI4.html5/25/2008 8:56:23 AM
Euclid's Elements, Book I, Definition 5
Definition 5
A surface is that which has length and breadth only.
This statement suggests that a surface has two dimensions, but says very little, if anything, since neither
length nor breadth have been defined yet, nor will they be. From the next definition, it is clear that a
surface does not have to be a plane. Other examples of surfaces that appear in the Elements are surfaces
of cones, cylinders, and spheres.
Book I introduction
© 1996
D.E.Joyce
Clark University
http://aleph0.clarku.edu/~djoyce/java/elements/bookI/defI5.html5/25/2008 9:02:27 AM
Euclid's Elements, Book I, Definition 6
Definition 6
The edges of a surface are lines.
As in I.Def.3, this statement describes a certain relationship, but this time between surfaces and lines.
For example, a hemisphere is a surface, and its edge is the circumference of a circle, a kind of line.
This definition cannot actually be used since there are no postulates to go along with it to connect the
edges of a surface in any way to the surface.
Euclid uses the same term for the end of a line in I.Def.3, the edge of a surface in this definition, and the
surface of a solid in XI.Def.2. That term could be translated as "that which is around," "the limits of," or
"the extremities of," but in English the terms "the ends of" a line, "the edges of" a surface, and either
"the surfaces of" or "the faces of" a solid are fairly standard for different dimensions.
Book I introduction
© 1996
D.E.Joyce
Clark University
http://aleph0.clarku.edu/~djoyce/java/elements/bookI/defI6.html5/25/2008 9:02:54 AM
Euclid's Elements, Book I, Definition 7
Definition 7
A plane surface is a surface which lies evenly with the straight lines on itself.
We see now that a plane surface, usually abbreviated to the single word "plane," is a kind of surface.
Perhaps the remainder of the statement is a definition of content, but, if so, some words are missing.
One interpretation often given is that if a plane surface contains two points, then it contains the line
connecting the two points. If that were the meaning, then it would be just as well to make that the
explicit definition or to make it a postulate. But that does not seem to be Euclid's intent. His proposition
XI.7 has a detailed proof that the line joining two points on two parallel lines lies in the plane of the two
parallel lines. No proof at all would be necessary if that line were by definition or by postulate contained
in a plane that contained its ends.
Note that a plane surface may be infinite, but needn't be infinite. It can be a square, a circle, or any other
plane figure (Def.I.19).
There are no postulates in the Elements for the existance of plane surfaces, either finite or infinite. Post.3
says circles can be drawn, but a ambient plane is implicitly required there. Rectilinear figures are
assumed to exist once the bounding lines have been constructed, but again, a plane is presumed to exist
first. Throughout Books I through IV and Book VI, the books on plane geometry, there is the implicit
assumption of one plane in which all the points, lines, and circles lie. In the books on solid geometry,
Books XI through XIII, there is sometimes mentioned a "plane of reference," and proposition XI.2
claims that two intersecting lines determine a plane as does any triangle (but its proof fails completely).
Book I introduction
© 1996
D.E.Joyce
Clark University
Definition 8
A plane angle is the inclination to one another of two lines in a plane which meet one another and
do not lie in a straight line.
The concept of angle is a very important concept for all of Greek geometry. Many of the propositions
require angles even for their statements.
The two lines are meant to emanate from the same point; two intersecting lines will actually make four
angles.
The concept is also a difficult one, and, surprisingly, broader than our modern concept of angle.
Book I introduction
© 1996
D.E.Joyce
Clark University
Definition 9
And when the lines containing the angle are straight, the angle is called rectilinear.
This continues the previous definition of angle. Nearly all the angles that
appear in the Elements are rectilinear as is the illustrated angle BAC.
Angles usually are named by three points, the middle point the vertex of
the angle. When there is no ambiguity it is sufficient to name the angle by
its vertex, in this example, A.
Angles as magnitudes
As treated by Euclid, rectilinear angles are magnitudes that can be added together. When the sum of
angles happens greater than two right angles, it is continued to be treated as a sum of angles rather than
an individual angle. For instance, in proposition I.32 it is proved that the sum of the interior angles of a
triangle equals two right angles.
Treating angles as magnitudes should not be confused with measuring angles. The angles themselves are
the magnitudes. The only measurement of angles in the Elements is in terms of right angles (defined in
the next definition). Degree measurement and radian measurement were not used until later. In terms of
degrees a right angle is 90°, while in terms of radians a right angle is pi/2 radians.
Throughout ancient Greek mathematics, only positive magnitudes were considered. Zero and negative
magnitudes were not conceived. For the most part, a lack of zero and negative magnitudes complicates
mathematics, but occasionally simplifies it. In any case, the power of a mathematics without zero and
negative magnitudes is no less in the sense that any statement made using the language of zero or
negative magnitudes can be translated into a statement that doesn't use them, although the translated
statement may be longer and less understandable.
Although in modern mathematics, angles can be positive, negative, or zero, and can be greater than a full
circle (360° or 2 pi radians), in the Elements angles are always greater than zero and less than two right
angles (180° or pi radians), except perhaps in one interpretation of proposition III.20 where the central
angle of a circle could be greater than two right angles.
Book I introduction
© 1996
D.E.Joyce
Clark University
Definition 10
When a straight line standing 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 called a
perpendicular to that on which it stands.
In the figure, the two angles DBA and DBC are equal, so they are right angles by definition, and so the
line BD set up on the line AC is perpendicular to it. Later there will be a postulate (Post.4) which states
that all right angles are equal, and after a few propositions, it can be shown that AC is also perpendicular
to BD. There are no postulates that explicitly state perpendiculars exist. Instead a construction for them
is given and proved in proposition I.11.
This is the first mention in the Elements of magnitudes being equal. There are several different kinds of
magnitudes in the Elements besides angles. Lines, plane figures, and solids are also kinds of magnitudes.
Some of the assumptions about magnitudes are stated later as "common notions" C.N., which are often
called "axioms." One thing that magnitudes of the same kind can be is "equal," as the angles in this
definition can be. Nowhere does Euclid explicitly state what it means for angles to be equal, or for that
matter, for lines, plane figures, or solids to be equal, although much can be determined by the way he
uses equality.
Book I introduction
© 1996
D.E.Joyce
Clark University
Definition 19
Rectilinear figures are those which are contained by straight lines, trilateral figures being those
contained by three, quadrilateral those contained by four, and multilateral those contained by
more than four straight lines.
Euclid classifies rectilinear figures by their number of sides in this definition. Classifying them by their
number of angles could lead to complications since an angle has to be less than two right angles, and a
non-convex figure would have an internal angle greater than two right angles.
The modern English names, however, are based an the number of angles (except quadrilateral): triangle,
pentagon, hexagon, heptagon, octagon, etc. From pentagon on up these names derive from the Greek,
but they're rarely used past octagon.
Book I introduction
© 1996
D.E.Joyce
Clark University
Definition 22
Of quadrilateral figures, a square is that which is both equilateral and right-angled; an oblong that
which is right-angled but not equilateral; a rhombus that which is equilateral but not right-angled;
and a rhomboid that which has its opposite sides and angles equal to one another but is neither
equilateral nor right-angled. And let quadrilaterals other than these be called trapezia.
The figure A is, of course, a square. Figure B is an oblong, or a rectangle. Figure C is a rhombus. Figure
D is a trapezium (sometimes called a trapeze or trapezoid). And figure E is a parallelogram.
The only figure defined here that Euclid actually uses is the square. The other names of figures may
have been common at the time of Euclid's writing, or they may have been left over from earlier authors'
versions of the Elements. Euclid makes much use of parallelogram, or parallelogrammic area, which he
does not define, but clearly means quadrilateral with parallel opposite sides. Parallelograms include
rhombi and rhomboids as special cases. And rather than oblong, he uses rectangle, or rectangular
parallelogram, which includes both squares and oblongs.
Squares and oblongs are defined to be "right-angled." Of course, that is intended to mean that all four
angles are right angles. Sometimes Euclid's definitions are too brief, but the intended meaning can easily
be determined from the way the definitions are used. In particular, proposition I.46 constructs a square,
and all four angles are constructed to be right, not just one of them.
Book I introduction
© 1996
D.E.Joyce
Clark University
Definition 23
Parallel straight lines are straight lines which, being in the same plane and being produced
indefinitely in both directions, do not meet one another in either direction.
Book I introduction
© 1996
D.E.Joyce
Clark University
http://aleph0.clarku.edu/~djoyce/java/elements/bookI/defI23.html5/25/2008 9:09:25 AM
Euclid's Elements, Book I, Definitions 11 and 12
Definitions 11 and 12
Def. 11. An obtuse angle is an angle greater than a right angle.
Note that there is no requirement that the angle be rectilinear, indeed, the horn angles mentioned before
are not rectilinear, but they are less than right angles, and so are acute (notwithstanding Proclus' remarks
to the contrary).
With these definition, we see another aspect of magnitudes, namely, two magnitudes of the same kind,
such as two angles, can be compared for size. Euclid frequently uses what is known as the law of
trichotomy: given two magnitudes F and G of the same kind, exactly one of the following three
situations hold, F is less than G, F equals G, or F is greater than G. See the comments after the Common
Notions for more discussion on magnitudes and the law of trichotomy.
Book I introduction
© 1996
D.E.Joyce
Clark University
Definitions 13 and 14
Def. 13. A boundary is that which is an extremity of anything.
These are rather nebulous definitions since they are based on the undefined terms "extremity" and
"contained by." Euclid deals with two kinds of figures in the Elements: plane figures and solid figures.
Plane figures are defined in the upcoming definitions: circles and semicircles in I.Def.15 and I.Def.18,
rectilinear figures in I.Def.19 and particular kinds of rectilinear figures such as triangles and
quadrilaterals following that. Specific solid figures such as spheres, cones, pyramids, and various
polyhedra are defined in Book XI. Plane figures are not solid figures since they are not contained by any
boundaries in space. Thus, implicit to the concept of figure is the ambient plane or space of the figure.
Euclid deals with three kinds of extremities, or boundaries. There are the ends of lines (I.Def.3), the
edges of surfaces (I.Def.3), and the surfaces of solids (XI.Def.2). A finite line has two points as its
boundaries. A circle is defined in I.Def.15 as is a plane figure and has its circumference as its boundary.
A sphere is defined in XI.Def.14 as a solid figure and has a spherical surface as its boundary.
The modern subject of topology studies space in a different way than geometry does. The geometric
concepts of straightness, distance, and angle are excluded from topology, but the concept of boundary is
central to topology. In topology, a sphere remains a sphere even when it's squeezed or stretched.
Not everything has a boundary. For instance, the circumference of a circle has no boundary. Also a
spherical surface has no boundary. In topology, a finite region with no boundary is called a cycle.
Circles and spherical surfaces are cycles. In general, if something is a boundary, it has no boundary
itself. So boundaries are cycles. But not all cycles are boundaries.
Topology uses observation to distinguish various spaces. For instance, on a spherical surface, every
circle is the boundary of a region on that surface. But on a toroidal surface (rotate a circle around a line
in the plane of the circle that doesn't meet the circle), there are circles (for instance, that circle mentioned
parenthetically) that don't bound any region on the surface. Thus, spherical surfaces are topologically
different from toroidal surfaces.
The definition of figure needs to be fleshed out. In order to be a figure, a region must be bounded, that
is, held in by a boundary. For instance, an infinite plane is unbounded, so it is not intended to be a
figure. Neither is the region between two parallel lines even though that region has the two parallel lines
as its extremities.
Other figures may be considered if other ambient spaces are allowed, although Euclid only uses plane
and solid figures. For a one-dimensional example, a line segment could be considered to be a figure in
an infinite line with its endpoints as its boundary. Also, a hemisphere could be considered to be a figure
on the surface of a sphere with the equator as its boundary.
Book I introduction
© 1996
D.E.Joyce
Clark University
Def. 16. And the point is called the center of the circle.
Def. 17. A diameter of the circle is any straight line drawn through the center and terminated in
both directions by the circumference of the circle, and such a straight line also bisects the circle.
Def. 18. A semicircle is the figure contained by the diameter and the circumference cut off by it.
And the center of the semicircle is the same as that of the circle.
A definition such as this describes what circles are. Definitions do not guarantee the existence of the
things they define. The existence of circles follows from a postulate, namely, Post.3.
Note that a circle for Euclid is a two-dimensional figure. But in modern mathematics, usually the word
"circle" refers to what Euclid calls the circumference of a circle.
The center of the circle in the diagram is the point C. It's interesting
that the English word "center" derives from the Greek word which also
means a prod or a poker, and it refers to the fixed leg of a compass
used to draw a circle.
The (curved) line ABD that contains the circle is its circumference.
Euclid typically names a circle by three points on its circumference.
Perhaps a better translation than "circumference" would be "periphery"
since that is the Greek word while "circumference" derives from the
Latin.
Euclid doesn't have a term for "radius" other than "that from the center," but "radius" is such a useful
word that it is used here to translate "that from the center," such as the radius CD.
An example diameter is the line AB which passes through the center. Of course, a diameter is twice a
radius, and since the radii are all equal to each other by definition, the diameters also all equal to each
other.
That diameters "also bisects the circle" should not be part of the definition, but either assumed as a
postulate or proved as a proposition. It depends on the fact that circles are drawn on planes, and planes
have constant curvature. The analogous figure on a surface of nonconstant curvature does not have this
property. For such figures the two "semicircles" on either side of a "diameter" need not be equal.
Although circles are used throughout Book I, the proper theory of circles doesn't begin until Book III.
That book begins with more definitions relating to circles including the equality of circles, when circles
touch (are tangent to) lines and other circles, and so forth.
Book I introduction
Definitions 20 and 21
Def. 20. Of trilateral figures, an equilateral triangle is that which has its three sides equal, an
isosceles triangle that which has two of its sides alone equal, and a scalene triangle that which
has its three sides unequal.
Def. 21. Further, of trilateral figures, a right-angled triangle is that which has a right angle, an
obtuse-angled triangle that which has an obtuse angle, and an acute-angled triangle that which
has its three angles acute.
Definition 20 classifies triangles by their symmetries, while definition 21 classifies them by the kinds of
angles they contain.
The scalene triangle C has no symmetries, but the isosceles triangle B has a bilateral symmetry. The
equilateral triangle A not only has three bilateral symmetries, but also has 120°-rotational symmetries.
Equilateral triangles are constructed in the very first proposition of the Elements, I.1. An alternate
characterization of isosceles triangles, namely that their base angles are equal, is demonstrated in
propositions I.5 and I.6.
Since triangle D has a right angle, it is a right triangle. Proposition I.17 states that the sum of any two
angles in a triangle is less than two right angles, therefore, no triangle can contain more than one right
angle. Furthermore, there can be at most one obtuse angle, and a right angle and an obtuse angle cannot
occur in the same triangle.
Triangle E is an obtuse triangle since it has an obtuse angle, while triangle F is an acute triangle since all
its angles are acute.
Book I introduction
© 1996
D.E.Joyce
Clark University
Postulate 1
To draw a straight line from any point to any point.
This first postulate says that given any two points such as A and B, there is a
line AB which has them as endpoints. This is one of the constructions that
may be done with a straightedge (the other being described in the next
postulate).
Although it doesn't explicitly say so, there is a unique line between the two points. Since Euclid uses this
postulate as if it includes the uniqueness as part of it, he really ought to have stated the uniqueness
explicitly.
The last three books of the Elements cover solid geometry, and for those, the two points mentioned in
the postulate may be any two points in space. Proposition XI.1 claims that if part of a line is contained in
a plane, then the whole line is. In the books on plane geometry, it is implicitly assumed that the line AB
joining A to B lies in the plane of discussion.
Book I introduction
© 1996
D.E.Joyce
Clark University
Postulate 2
To produce a finite straight line continuously in a straight line.
Here we have the second ability of a straightedge, namely, to extend a given line AB to CD. This
postulate does not say how far a line can be extended. Sometimes it is used so that the extension equals
some other line. Other times it is extended arbitrarily far.
As with the first postulate, it is implicitly assumed in the books on plane geometry that when a line is
extended, it remains in the plane of discussion. The first proposition on solid geometry, proposition
XI.1, claims that line can't be only partly in a plane. The central step in the proof of that proposition is to
show that a line cannot be extended in two ways, that is, there is only one continuation of a line. The
proof is hardly convincing. Rather, this postulate should include a clause to that effect.
Other uses of a straightedge can be imagined. For instance, it might be marked at two points on it, then
fit into a diagram so that the two points fall on two lines, perhaps curved. This operation is an example
of "neusis" or "verging" where lines are adjusted to fit the diagram. For instance, Archimedes, who lived
in the century after Euclid, used neusis in several constructions in his work On Spirals. In the Book of
Lemmas, attributed by Thabit ibn-Qurra to Archimedes, neusis is used to trisect an angle.
With the help of Euclid's propositions here in Book I, we can show that angle EGC is one-third of angle
ABC. Since the lines GH, HB, and BE are equal, therefore the triangles GHB and HBE are isosceles.
Therefore, by I.5, angle HGB equals HBG, and angle BHE equals angle BEH.
By I.32, the exterior angle BHE of triangle GHB equals the sum of the equal angles HGB and HBG,
therefore angle BHE is double angle HGB. And angle BEH equals BHE, so it is also double angle HGB.
Again by I.32, the exterior angle ABC of triangle BEG is the sum of angles HGB and BEH. But angle
BEH is double angle HGB, therefore angle ABC is triple angle GHB.
The ancient Greek geometers believed that angle trisection required tools beyond those given in Euclid's
postulates. They were right, but it wasn't proved until Wantzel in 1837 proved that a 60°-angle cannot be
so trisected using only Euclidean tools.
Euclid has no postulate for neusis constructions, and since neusis constructions can trisect angles, we
conclude from Wantzel's theorem that another postulate is required to justify neusis constructions.
Book I introduction
© 1996
D.E.Joyce
Clark University
Postulate 3
To describe a circle with any center and radius.
This is the third assumed construction in the Elements. It corresponds to drawing a circle with a
compass.
Circles were defined in Def.I.15 and Def.I.16 as plane figures with the property
that there is a certain point, called the center of the circle, such that all straight
lines from the center to the boundary are equal. That is, all the radii are equal.
The given data are (1) a point A to be the center of the circle, (2) another point B
to be on the circumference of the circle, and (3) a plane in which the two points
lie. In the first few books of the Elements, there is but one plane under
consideration and needn't be mentioned, but in the last three books which
develop solid geometry, the plane has to be specified.
Note that this postulate does not allow for the compass to be moved. The usual way that a compass is
used is that is is opened to a given width, then the pivot is placed on the drawing surface, then a circle is
drawn as the compass is rotated around the pivot. But this postulate does not allow for transferring
distances. It is as if the compass collapses as soon as it's removed from the plane. Proposition I.3,
however, gives a construction for transferring distances. Therefore, the same constructions that can be
made with a regular compass can also be made with Euclid's collapsing compass.
Book I introduction
© 1996
D.E.Joyce
Clark University
Postulate 4
That all right angles equal one another.
Book I introduction
© 1996
D.E.Joyce
Clark University
http://aleph0.clarku.edu/~djoyce/java/elements/bookI/post4.html5/25/2008 9:11:43 AM
Euclid's Elements, Book I, Postulate 5
Postulate 5
That, if a straight line falling on two straight lines makes the interior angles on the same side less
than two right angles, the two straight lines, if produced indefinitely, meet on that side on which
are the angles less than the two right angles.
Of course, this is a postulate for plane geometry. It should include the condition that the two straight
lines lie in a plane, otherwise, skew lines in space would satisfy the hypotheses. Also, without an
ambient plane, the term "that side [of the straight line]" has no meaning.
In the diagram, if angle ABE plus angle BED is less than two
right angles (180°), then lines AC and DF will meet when
extended in the direction of A and D.
This postulate is usually called the "parallel postulate" since it can be used to prove properties of parallel
lines. Euclid develops the theory of parallel lines in propositions through I.31.
The parallel postulate is historically the most interesting postulate. Geometers throughout the ages have
tried to show that it could be proved from the remaining postulates so that it wasn't necessary to assume
it. The process tried was to assume its falsehood, then derive a contradiction. Many strange conclusions
follow from denying the parallel postulate, and several geometers found such great absurdities that they
concluded that the parallel postulate did follow from the rest.
Nevertheless, these apparent absurdities are not contradictions. In the early nineteenth century, Bolyai,
Lobachevsky, and Gauss found ways of dealing with this "non-Euclidean" geometry by means of
analysis and accepted it as a valid kind of geometry, although very different from Euclidean geometry.
This hyperbolic geometry, as it is called, is just as consistent as Euclidean geometry and has many uses.
Thus, we know now that we must include the parallel postulate to derive Euclidean geometry. For more
on noneuclidean geometries, see the notes on hyperbolic geometry after I.29 and elliptic geometry after
I.16.
Euclid does not use this parallel postulate until Proposition I.29, but nearly all of the rest of Book I
depends on it. For more commentary about the postulate see the Guides to I.29 and I.30.
Book I introduction
© 1996, 2003
D.E.Joyce
Clark University
Proposition 1
But AC was proved equal to AB, therefore each of the straight lines AC and BC equals AB.
And things which equal the same thing also equal one another, therefore AC also equals BC. C.N.1
Therefore the three straight lines AC, AB, and BC equal one another.
Therefore the triangle ABC is equilateral, and it has been constructed on the given finite I.Def.20
straight line AB.
Q.E.F.
This proposition is a very pleasant choice for the first proposition in the Elements. The construction of
the triangle is clear, and the proof that it is an equilateral triangle is evident. Of course, there are two
choices for the point C, but either one will do.
Euclid could have chosen proposition I.4 to come first, since it doesn't logically depend on the previous
three, but there are some good reasons for putting I.1 first. For one thing, the Elements ends with
constructions of the five regular solids in Book XIII, so it is a nice aesthetic touch to begin with the
construction of a regular triangle. More important, though, is I.1 is needed in I.2, and that in I.3.
Propositions I.2 and I.3 give constructions for moving lines, and I.4, although not logically dependent on
I.2 or I.3, does use the concept of superposition which involves, in some sense, moving points and lines.
The abbreviations in the right column refer to postulates, definitions, common notions, and previously
proved propositions. Each indicates a justification of a construction or conclusion in a sentence to its
left. They are not part of Euclid's Elements, but it is a tradition to include them as a guide to the reader.
Sometimes the justification is quoted as C.N.1 is quoted here, but usually it is left to the reader to
determine the justification.
The Q.E.F. at the end of the proof is an abbreviation for the Latin words "quod erat faciendum" which
means "which was to be done." A few of the propositions, as this one and the next two, solve problems
by constructions. These are the ones that end with Q.E.F. (They're also printed in red here in the listings
of propositions for each book.)
The rest of the proofs end with Q.E.D. instead, an abbreviation for "quod erat demonstrandum" which
means "which was to be demonstrated." It's convenient to have a standard way to indicate the end of a
proof. These Latin abbreviations are a bit of an anachronism. It would be less of an anachronism to use
abbreviations for the original Greek phrase, or abbreviations for a modern English phrase since the rest
of this version of the Elements is in English. But by now, Q.E.F. and Q.E.D. are traditional. In recent
decades a small square has become common as a symbol to indicate the end of a proof.
It is surprising that such a short, clear, and understandable proof can be so full of holes. These are
logical gaps where statements are made with insufficient justification. Having the first proof in the
Elements this proposition has probably received more criticism over the centuries than any other.
Why does the point C exist? Near the beginning of the proof, the point C is mentioned where the
circles are supposed to intersect, but there is no justification for its existence. The only one of Euclid's
postulate that says a point exists the parallel postulate, and that postulate is not relevant here. Thus, there
is no assurance that the point C actually exists. Indeed, there are models of geometry in which the circles
do not intersect. Thus, other postulates not mentioned by Euclid are required. In Book III, Euclid takes
some care in analyzing the possible ways that circles can meet, but even with more care, there are
missing postulates.
Why is ABC a plane figure? After concluding the three straight lines AC, AB, and BC are equal, what is
the justification that they contain a plane figure ABC? Recall that a triangle is a plane figure bounded by
contained by three lines. These lines have not been shown to lie in a plane and that the entire figure lies
in a plane. It is proposition XI.1 that claims that all parts of a line lie in a plane, and XI.2 that claims that
the entire triangle lie in a plane. Logically, they should precede I.1. The reason they don't, of course, is
that those propositions belong to solid geometry, and plane geometry is developed first in the Elements,
also, no doubt, plane geometry developed first historically.
Why does ABC contain an equilateral triangle? Proclus relates that early on there were critiques of
the proof and describes that of Zeno of Sidon, an Epicurean philosopher of the early first century B.C.E.
(not to be confused with Zeno of Elea famous of the paradoxes who lived long before Euclid), and
whose criticisms, Proclus says, were refuted in a book by Posidonius. The critique is sound, however,
and the refutation faulty.
The possibilities that haven't been excluded are much more numerous than Zeno's example. The sides
could meet numerous times and the region they contain could look like a necklace of bubbles. What
needs to be shown (or assumed as a postulate) is that two infinitely extended straight lines can meet in at
most one point.
Use of Proposition 1
The construction in this proposition is directly used in propositions I.2, I.9, I.10, I.11, XI.11, and XI.22.
Book I introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 2
To place a straight line equal to a given straight line with one end at a given point.
Let A be the given point, and BC the given straight line.
It is required to place a straight line equal to the given straight line BC with one end at the
point A.
And in these DA equals DB, therefore the remainder AL equals the remainder BG. C.N.3
But BC was also proved equal to BG, therefore each of the straight lines AL and BC equals
BG. And things which equal the same thing also equal one another, therefore AL also equals C.N.1
BC.
Therefore the straight line AL equal to the given straight line BC has been placed with one
end at the given point A.
Q.E.F.
This is a very clever construction to solve what seems to be a simple problem. One would like simply to
slide the line BC along so that one end coincides with the point A. But there is no motion in the
geometry of Euclid. There is something like motion used in proposition I.4, but nothing is actually
moved there. The only basic constructions that Euclid allows are those described in Postulates 1, 2, and
3. Euclid then builds new constructions (such as the one in this proposition) out of previously described
constructions. So at this point, the only constructions available are those of the three postulates and the
construction in proposition I.1, and Euclid uses all four here.
Another, different, expectation is that one might use a compass to transfer the distance BC over to the
point A. It is clear from Euclid's use of postulate 3 that the point to be used for the center and a point that
will be on the circumference must be constructed before applying the postulate; postulate 3 is not used to
transfer distance. Sometimes postulate 3 is likened to a collapsing compass, that is, when the compass is
lifted off the drawing surface, it collapses.
It could well be that in some earlier Greek geometric theory abstracted compasses that could transfer
distances. If that speculation is correct, then this proposition would be a late addition to the theory. The
construction of the proposition allows a weaker postulate (namely postulate 3) to be assumed.
Construction steps
Use of Proposition 2
Note that this constuction assumes that all the point A and the line BC lie in a plane. It may also be used
in space, however, since Proposition XI.2 implies that A and BC do lie in a plane.
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 3
To cut off from the greater of two given unequal straight lines a straight line equal to the less.
Place AD at the point A equal to the straight line C, and describe the circle DEF with center I.2
A and radius AD. Post. 3
Now, since the point A is the center of the circle DEF, therefore AE equals AD. I.Def.15
But C also equals AD, therefore each of the straight lines AE and C equals AD, so that AE C.N.1
also equals C.
Therefore, given the two straight lines AB and C, AE has been cut off from AB the greater
equal to C the less.
Q.E.F.
Now it is clear that the purpose of Proposition 2 is to effect the construction in this proposition.
According to Proclus (410-485 C.E.) in his Commentary on Book I, Hippocrates of Chios (fl. ca. 430 B.
C.E.) was the first to write an Elements. Leon and Theudius also wrote versions before Euclid (fl. ca.
295 B.C.E.). These other Elements have all been lost since Euclid's replaced them. It is conceivable that
in some of these earlier versions the construction in proposition I.2 was not known, so this proposition
would instead have been a postulate (a stronger version of Post.3). Once the construction in I.2 was
discovered, the current weaker Post.3 would do. Then again, I.2 might go back to the time of
Hippocrates.
Construction steps
This construction takes one more step beyond that of I.2, and that is the final circle, the circle shown in
the diagram accompanying this proposition. Altogether, therefore, five circles and two lines are required
for this construction.
Frequently, though, one end of the line C is already placed at A, and then the construction of I.2 isn't
required. In that case, only one circle needs to be drawn.
Use of Proposition 3
This proposition begins the geometric arithmetic of lines. Explicitly, it allows lines to be subtracted, but
it can also be used to compare lines for equality and to add lines, that is, one line can be placed alongside
another to determine if they are equal, or if not, which is greater. In other words, this construction
justifies the law of trichotomy for lines.
The construction is use more often in the Elements than any other starting with proposition I.5. It is used
in all the books on geometry, that is in Books I through IV, VI, and XI through XIII.
Naturally a construction of this sort is needed in the solid geometry of Books XI through XIII.
Surprisingly, the construction given here also works in solid geometry, even the lines AB and C don't lie
in the same plane. Since the point A and the line C lie in one plane, the construction of I.2 produces a
line AD equal to C in that plane. Now AD and AB also lie in one plane, but not the same one, and the
circle AEF can be drawn there.
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 4
If two triangles have two sides equal to two sides respectively, and have the angles contained
by the equal straight lines equal, then they also have the base equal to the base, the triangle
equals the triangle, and the remaining angles equal the remaining angles respectively, namely
those opposite the equal sides.
Let ABC and DEF be two triangles having the two sides AB and AC equal to the two
sides DE and DF respectively, namely AB equal to DE and AC equal to DF, and the
angle BAC equal to the angle EDF.
I say that the base BC also equals the base EF, the triangle ABC equals the triangle DEF,
and the remaining angles equal the remaining angles respectively, namely those opposite
the equal sides, that is, the angle ABC equals the angle DEF, and the angle ACB equals
the angle DFE.
Again, AB coinciding with DE, the straight line AC also coincides with DF, because the
angle BAC equals the angle EDF. Hence the point C also coincides with the point F,
because AC again equals DF.
But B also coincides with E, hence the base BC coincides with the base EF and equals it. C.N.4
Thus the whole triangle ABC coincides with the whole triangle DEF and equals it. C.N.4
And the remaining angles also coincide with the remaining angles and equal them, the
angle ABC equals the angle DEF, and the angle ACB equals the angle DFE.
Therefore if two triangles have two sides equal to two sides respectively, and have the angles
contained by the equal straight lines equal, then they also have the base equal to the base, the triangle
equals the triangle, and the remaining angles equal the remaining angles respectively, namely those
opposite the equal sides.
Q.E.D.
This is the first of the congruence propositions for triangles. Euclid did not explicitly use the concept of
congruence, although it would have simplified his exposition a bit. The definition of congruence would
include the hypotheses and conclusions of this proposition, that is, two triangles ABC and DEF are
congruent if angles A, B, and C are equal to angles D, E, and F respectively, and sides AB, BC, and AC
are equal to sides DE, EF, and DF respectively, and the triangle ABC equals the triangle DEF (by which
is meant that they have the same area). In the books on solid geometry, Euclid uses the phrase "similar
and equal" for congruence, but similarity is not defined until Book VI, so that phrase would be out of
place in the first part of the Elements.
For more discussion of congruence theorems see the note after proposition I.26, the last of the
congruence propositions.
Euclid frequently refers to one side of a triangle as its "base," leaving the other two named "sides." Any
one of the sides might be chosen as the base, but once chosen, it remains the base for the rest of the
discussion. This is simply a linguistic device to save words.
The method of proof used in this proposition is sometimes called "superposition." It apparently is not a
method that Euclid prefers since he so rarely uses it, only here in I.4 and in I.8 and III.24, but not in
many other propositions in which he could have used it.
It is not entirely clear what is meant by "superposing a triangle on a triangle" means. It has been
variously interpreted as actually moving one triangle to cover the other or as simply associating parts of
one triangle with parts of the other. For the two triangles illustrated in the figure, you can actually slide
one over the other in a continuous motion within the plane. Note, however, that if one triangle is the
mirror image of the other, then any continuous motion would require moving one triangle outside of the
plane. But the triangles don't have to be same plane to begin with, and they often are not in the same
plane when this proposition is invoked in the books on solid geometry.
Whatever the intended meaning of superposition may be, there are no postulates to allow any
conclusions based on superposition. One possibility is to add postulates based on a group of
transformations of space, or if restricted to plane geometry, on a group of transformations of the plane.
Charles Dodgson (a.k.a. Lewis Carroll) would have said that using group theory is not appropriate to an
elementary exposition of Euclidean geometry. Heath has described a more elementary conservative basis
in his commentary on this proposition.
Yet another alternative is to simply take this proposition as a postulate, or part of it as a postulate. For
instance, Hilbert in his Foundations of Geometry takes as given that under the hypotheses of this
proposition that the remaining angles equal the remaining angles. Then, Hilbert proves that the base
equals the base.
Use of Proposition 4
Of the various congruence theorems, this one is the most used. This proposition is used frequently in
Book I starting with the next two propositions, and it is often used in the rest of the books on geometry,
namely, Books II, III, IV, VI, XI, XII, and XIII.
Although the two triangles in this proposition appear to be in the same plane, that is not necessary. In
Proposition XI.4 and many others in Book XI this proposition is applied to pairs of triangles in different
planes.
Book I introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 5
In isosceles triangles the angles at the base equal one another, and, if the equal straight lines
are produced further, then the angles under the base equal one another.
I.
Let ABC be an isosceles triangle having the side AB equal to the side AC, and let the straight Def.20
lines BD and CE be produced further in a straight line with AB and AC. Post.2
I say that the angle ABC equals the angle ACB, and the angle CBD equals the angle BCE.
Take an arbitrary point F on BD. Cut off AG from AE the greater equal to AF the less, and I.3.
join the straight lines FC and GB. Post.1
Since the whole AF equals the whole AG, and in these AB equals AC, therefore the C.N.3
remainder BF equals the remainder CG.
But FC was also proved equal to GB, therefore the two sides BF and FC equal the two sides
CG and GB respectively, and the angle BFC equals the angle CGB, while the base BC is
common to them. Therefore the triangle BFC also equals the triangle CGB, and the I.4
remaining angles equal the remaining angles respectively, namely those opposite the equal
sides. Therefore the angle FBC equals the angle GCB, and the angle BCF equals the angle
CBG.
Accordingly, since the whole angle ABG was proved equal to the angle ACF, and in these
the angle CBG equals the angle BCF, the remaining angle ABC equals the remaining angle C.N.3
ACB, 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 in isosceles triangles the angles at the base equal one another, and, if the equal straight
lines are produced further, then the angles under the base equal one another.
Q.E.D.
There are two conclusions for this proposition, first that the internal base angles ABC and ACB are equal,
second that the external base angles FBC and GCB are equal. From the diagram it looks like it would be
easy to prove the second conclusion from the first by simply subtracting the equal angles ABC and ACB
the straight angles ABF and ACG, respectively. But Euclid doesn't accept straight angles, and even if he
did, he hasn't proved that all straight angles are equal. Proposition I.13 would be enough, since it implies
the sum of angles ABC and FBC equals two right angles, and the sum of angles ACB and GCB also
equals two right angles, and so the two sums are equal effectively saying all straight angles are equal.
Unfortunately, such an argument would be circular. I.13 depends on I.11, I.11 on I.8, I.8 on I.7, and I.7
on I.5. Thus, I.13 cannot be used in the proof of I.5. It may appear that I.7 only depends on the first
conclusion of I.5, but a case of I.7 that Euclid does not discuss relies on the second conclusion of I.5.
This proposition has been called the Pons Asinorum, or Asses' Bridge. Whether this name is due to its
difficulty (which it isn't) or the resemblance of its figure to a bridge is not clear. Very few of the
propositions in the Elements are known by names.
Pappus' proof
Pappus (fl. ca. 320 C.E.) gave a much shorter proof of the first conclusion, but it is also conceptually
more difficult. The two triangles BAC and CAB have two sides equal to two sides, namely side BA of the
first triangle equals side CA of the second triangle, and side AC of the first triangle equal to side AB of
the second, and the contained angles are equal, namely angle BAC of the first triangle equals angle CAB
of the second, therefore, by I.4, the corresponding parts of the two triangles are equal, in particular, the
angle B in the first triangle equals the angle C of the second.
The difficulty lies in treating one triangle as two, or in making a correspondence between a triangle and
itself, but not the correspondence of identity. There is nothing wrong with this proof formally, but it
might be more difficult for a student just learning geometry.
Use of Proposition 5
This proposition is used in Book I for the proofs of several propositions starting with I.7 It is also used
frequently in Books II, III, IV, VI, and XIII.
Book I introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 6
If in a triangle two angles equal one another, then the sides opposite the equal angles also
equal one another.
Let ABC be a triangle having the angle ABC equal to the angle ACB.
Therefore if in a triangle two angles equal one another, then the sides opposite the equal angles also
equal one another.
Q.E.D.
Converses of propositions
This is the converse of (part of) the previous proposition I.5. Proposition I.6 says that if angle B equals
angle C, then side AB equals side AC. Proposition I.5 says that if side AB equals side AC, then angle B
equals angle C. In general, the converse of a proposition of the form "If P, then Q" is the proposition "If
Q, then P." When both a proposition and its converse are valid, Euclid tends to prove the converse soon
after the proposition, a practice that has continued to this day.
A proposition and its converse are not logically equivalent. There are examples where "If P, then Q" is
valid, but "If Q, then P" is not valid. An example from the Elements is proposition III.5 which states "If
two circles cut one another, then they do not have the same center." The converse would be "If two
circles do not have the same center, then they cut one another" which is certainly not valid since if one
circle lies entirely outside the other, then they don't have the same center.
Proofs by contradiction
This is the first "proof by contradiction," also called "reductio ad absurdum," in the Elements. In this
proof, in order to prove AB equals AC, Euclid assumes they are unequal and derives a contradiction,
namely, that the triangle ACB equals a part of itself, triangle DBC, which contradicts C.N.5, the whole is
greater than the part. The contradiction is that triangle ACB both equals and does not equal triangle
DBC.
In general, to prove a statement of the form "P" with a proof by contradiction, begin with an assumption
"not P" and derive some contradiction "Q and not Q," and finally conclude "P."
Euclid often uses proofs by contradiction, but he does not use them to conclude the existence of
geometric objects. That is, he does not use them in constructions. But he does use them to show what
has been constructed is correct.
The proof uses the law of trichotomy for lines. "If AB does not equal AC, then one of them is greater."
There are three cases: AB < AC, AB = AC, or AB > AC. If the middle possibility is excluded, then only
the two others remain, so one of the lines is greater. The law of trichotomy is not explicitly stated as a
Common Notion, but it is the sort of property of magnitudes listed as Common Notions.
Proposition I.3 can be read as a construction to determine whether one line is less than, equal to, or
greater than another. Using I.3, one line is laid along another, and it will fall short, fall equal, or extend
beyond the other. For this proposition I.6, the construction simplifies since the two lines AB and AC
already have one end in common.
The other part of the law of trichotomy is also used in the proof, the part that says only one of the three
cases can occur. "... the triangle DBC equals the triangle ACB, the less equals the greater, which is
absurd." C.N.5, the whole is greater than the part, allows the conclusion that triangle DBC (the part) is
less than triangle ACB (the whole). But the contradiction arises because only one of the two cases
DBC = ACB and DBC < ACB can occur.
Use of Proposition 6
This proposition is not used in the proofs of any of the later propositions in Book I, but it is used in
Books II, III, IV, VI, and XIII.
Book I introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 7
Given two straight lines constructed from the ends of a straight line and meeting in a point,
there cannot be constructed from the ends of the same straight line, and on the same side of it,
two other straight lines meeting in another point and equal to the former two respectively,
namely each equal to that from the same end.
I.5
Since AC equals AD, therefore the angle ACD equals the angle ADC. Therefore the angle ADC C.
is greater than the angle DCB. Therefore the angle CDB is much greater than the angle DCB. N.5
C.N.
I.5
Again, since CB equals DB, therefore the angle CDB also equals the angle DCB. But it was
also proved much greater than it, which is impossible. C.N.
Therefore given two straight lines constructed from the ends of a straight line and meeting in a point,
there cannot be constructed from the ends of the same straight line, and on the same side of it, two
other straight lines meeting in another point and equal to the former two respectively, namely each
equal to that from the same end.
Q.E.D.
In order to conclude "the angle ADC is greater than the angle DCB" it is necessary for angle ADC to be
greater than angle DCB, but that won't happen unless the point D lies outside the triangle ABC. Euclid
hasn't considered the case when D lies inside triangle ABC as well as other special cases. This is not
unusual as Euclid frequently treats only one case. Commentators over the centuries have inserted other
cases in this and other propositions. It is usually easy to modify Euclid's proof for the remaining cases.
In this proposition for the case when D lies inside triangle ABC, the second conclusion of I.5 may be
used to justify the proof.
Hidden justifications
The sentences
... the angle ACDequals the angle ADC.Therefore the angle ADCis greater than the angle DCB.
Therefore the angle CDBis much greater than the angle DCB.
use several properties of magnitudes. C.N.5 justifies the unstated angle ACD > DCB since DCB is part
of ACD. The statement that ADC is greater than the angle DCB is justified by the property of magnitudes
justifies the last statement "CDB is much greater than the angle DCB." Transitivity is another property
not listed as a Common Notion.
As in the proof of the last proposition and many to come, the law of trichotomy is also used. Here it's
used to reach the final contradiction.
Use of Proposition 7
Book I introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 8
If two triangles have the two sides equal to two sides respectively, and also have the base equal
to the base, then they also have the angles equal which are contained by the equal straight
lines.
Let ABC and DEF be two triangles having the two sides AB and AC equal to the two sides DE
and DF respectively, namely AB equal to DE and AC equal to DF, and let them have the base
BC equal to the base EF.
Then, BC coinciding with EF, therefore BA and AC also coincide with ED and DF, for, if the
base BC coincides with the base EF, and the sides BA and AC do not coincide with ED and DF
but fall beside them as EG and GF, then given two straight lines constructed on a straight line
and meeting in a point, there will have been constructed on the same straight line 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 end with it.
Therefore it is not possible that, if the base BC is applied to the base EF, the sides BA and AC
do not coincide with ED and DF. Therefore they coincide, so that the angle BAC coincides C.N.4
with the angle EDF, and equals it.
Therefore if two triangles have the two sides equal to two sides respectively, and also have the base
equal to the base, then they also have the angles equal which are contained by the equal straight lines.
Q.E.D.
This, the "side-side-side" congruence theorem, is the second of Euclid's three congruence theorems for
triangles. See the note on congruence theorems after proposition I.26.
As in the proof of I.4, this proof employs the hazy method of superposition.
Use of Proposition 8
This proposition is used for the a few of the propositions in Book I starting with the next one. It is also
used several times in the Books III, IV, XI, and XIII.
As in I.4 the two triangles need not lie in one plane. Propositions such as XI.4 in Book XI apply this
theorem to the case when the two triangles are not coplanar.
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 9
I say that the angle BAC is bisected by the straight line AF.
I.
And the base DF equals the base EF, therefore the angle DAF equals the angle EAF. Def.20
I.8
Therefore the given rectilinear angle BAC is bisected by the straight line AF.
Q.E.F.
Construction steps
On angle trisection
Angle bisection is an easy construction to make using Euclidean tools of straightedge and compass.
Also, line bisection is quite easy (see the next proposition I.10), and division of a line into any number
of equal parts is not especially difficult (see proposition VI.9).
Dividing an angle into an odd number of equal parts is not so easy, in fact, it is impossible to trisect a
60°-angle using Euclidean tools (the Postulates 1 through 3). Euclid's predecessors employed a variety
higher curves for this purpose. Archimedes, after Euclid, created two constructions: his spiral could
divide an angle into any number of parts, and his neusis construction could trisect angles (see the note on
Post.2). By Pappus' time it was believed that angle trisection was not possible using Euclidean tools, but
that wasn't proven until 1837 when Wantzel published his proof.
Nevertheless, amateur geometers continue to search in vain for such a construction and frequently bother
mathematicians with their purported solutions. Their solutions are of two forms. Sometimes they simply
construct approximate trisections. Other times they use neusis or some other other tool that goes beyond
Euclid's tools.
Students of geometry are cautioned not to waste their time on this problem and, if they do, not to bother
others with their purported solutions. Much better would be to study Galois theory, the mathematics that
proves the impossibility of angle trisection.
Use of Proposition 9
The construction of this proposition is used in the next one and a few propositions in Books IV, VI, and
XIII.
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 10
Since CA equals CB, and CD is common, therefore the two sides CA and CD equal the two I.
sides CB and CD respectively, and the angle ACD equals the angle BCD, therefore the base Def.20
AD equals the base BD. I.4
While this construction divides a line into two equal parts, the construction in proposition VI.9 divides a
line into any given number of equal parts.
Construction steps
This method for bisecting lines takes less actual work than it appears to. It is really no more than the
double-equilateral-triangle.
Actually, only two circles and the straight line CE need to be drawn. The straight lines AC, CB, AE, and
EB, aren't necessary for the construction; they are only used to show that the construction is correct.
Use of Proposition 10
The construction of this proposition in Book I is used in propositions I.12, I.16, and I.42. It is also used
in several propositions in the Books II, III, IV, X, and XIII.
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 11
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.
It is required to draw a straight line at right angles to the straight line AB from the point C.
Since CD equals CE, and CF is common, therefore the two sides CD and CF equal the two I.
sides CE and CF respectively, and the base DF equals the base EF. Therefore the angle DCF Def.20
equals the angle ECF, and they are adjacent angles. I.8
But, when a straight line standing on a straight line makes the adjacent angles equal to one I.Def.10
another, each of the equal angles is right, therefore each of the angles DCF and FCE is right.
Therefore the straight line CF has been drawn at right angles to the given straight line AB
from the given point C on it.
Q.E.F.
This and the next proposition both construct a perpendicular to a line through a given point. The
difference is that the given point lies on the line in this proposition but doesn't in the next.
Construction steps
Use of Proposition 11
Thia construction is used in propositions I.13, I.46, I.48, and numerous propositions in Books II, III, VI,
VI, XI, XII, and XIII.
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 12
To draw a straight line perpendicular to a given infinite straight line from a given point not on
it.
Let AB be the given infinite straight line, and C the given point which is not on it.
It is required to draw a straight line perpendicular to the given infinite straight line AB from
the given point C which is not on it.
Since GH equals HE, and HC is common, therefore the two sides GH and HC equal the two I.
sides EH and HC respectively, and the base CG equals the base CE. Therefore the angle Def.15
CHG equals the angle EHC, and they are adjacent angles. I.8
But, when a straight line standing 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 called I.Def.10
a perpendicular to that on which it stands.
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.
Again, the double-equilateral-triangle construction is used, but this time the preparation of the starting
line EG is different. The point D is taken on the other side of the line AB to insure that circle meets the
line AB in at least two points, E and G. If D is taken on the line AB, it might be taken at H, and the
resulting circle would touch the line only at H; and if D is taken on the same side of AB, then the circle
could miss the line entirely.
Euclid does not precede this proposition with propositions investigating how lines meet circles. He is
much more careful in Book III on circles in which the first dozen or so propositions lay foundations. For
instance, Proposition III.10 states that a circle does not cut a circle at more than two points. Even so,
some propositions are missing. One is needed for this proposition to justify the existence of the two
points C and E where the line AB meets circle with center C and radius CD. Such a proposition would
state "A circle whose center is on one side of a line and on whose circumference lies a point on the other
side of the line meets the line at two points."
Incidentally, Proclus explains in his commentary on Book I that the problem of constructing the
perpendicular was investigated by Oenopides of Chios who lived sometime in the middle of the fifth
century B.C.E., a century and a half before Euclid.
Use of Proposition 12
The construction of this proposition is not used in Book I, but it is used on occasion in the remaining
geometric books, namely, Books II through IV, VI, and XI through XIII.
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 13
If a straight line stands on a straight line, then it makes either two right angles or angles whose
sum equals two right angles.
I say that either the angles CBA and ABD are two
right angles or their sum equals two right angles.
Now, if the angle CBA equals the angle ABD, then they are two right angles. I.Def.10
But, if not, draw BE from the point B at right angles to CD. Therefore the angles CBE and I.11
EBD are two right angles.
Since the angle CBE equals the sum of the two angles CBA and ABE, add the angle EBD to
each, therefore the sum of the angles CBE and EBD equals the sum of the three angles CBA, C.N.2
ABE, and EBD.
Again, since the angle DBA equals the sum of the two angles DBE and EBA, add the angle
ABC to each, therefore the sum of the angles DBA and ABC equals the sum of the three C.N.2
angles DBE, EBA, and ABC.
But the sum of the angles CBE and EBD was also proved equal to the sum of the same three
angles, and things which equal the same thing also equal one another, therefore the sum of
the angles CBE and EBD also equals the sum of the angles DBA and ABC. But the angles C.N.1
CBE and EBD are two right angles, therefore the sum of the angles DBA and ABC also
equals two right angles.
Therefore if a straight line stands on a straight line, then it makes either two right angles or angles
whose sum equals two right angles.
Q.E.D.
With this proposition, we begin to see what the arithmetic of magnitudes means to Euclid, in particular,
how to add angles. Euclid says that the angle CBE equals the sum of the two angles CBA and ABE. So,
one way a sum of angles occurs is when the two angles have a common vertex (B in this case) and a
common side (BA in this case), and the angles lie on opposite sides of their common side. Thus, addition
of angles can be performed by joining adjacent angles.
But that's not the only addition that occurs here. Euclid also says that the sum of the angles CBE and
EBD equals the sum of the three angles CBA, ABE, and EBD. That sum being mentioned is a straight
angle, which is not to be considered as an angle according to Euclid. It is a formal sum equal to two right
angles. In other propositions formal sums of four right angles occur. These and larger formal sums are
not angles themselves, merely sums of angles. Only if an angle sum is less than two right angles can it
be identified with a single angle.
Use of Proposition 13
This proposition is used in the proofs of the next two propositions and several others in this book as well
as a few propositions in Books IV and VI.
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 14
If with any straight line, and at a point on it, two straight lines not lying on the same side make
the sum of the adjacent angles equal to two right angles, then the two straight lines are in a
straight line with one another.
With any straight line AB, and at the point B on it, let the two straight lines BC and BD not
lying on the same side make the sum of the adjacent angles ABC and ABD equal to two right
angles.
Subtract the angle CBA from each. Then the remaining angle ABE equals the remaining angle
ABD, the less equals the greater, which is impossible. Therefore BE is not in a straight line C.N.3
with CB.
Similarly we can prove that neither is any other straight line except BD. Therefore CB is in a
straight line with BD.
Therefore if with any straight line, and at a point on it, two straight lines not lying on the same side
make the sum of the adjacent angles equal to two right angles, then the two straight lines are in a
straight line with one another.
Q.E.D.
This is a proposition in plane geometry. If A, B, C, and D do not lie in a plane, then CBD cannot be a
straight line. An ambient plane is necessary to talk about the sides of the line AB
The qualifying sentence, "Similarly we can prove that neither is any other straight line except BD," is
meant to take care of the cases when E does not lie inside the angle ABD.
Use of Proposition 14
This proposition is used in propositions I.45, I.47, and a few in Books VI and XI.
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 15
If two straight lines cut one another, then they make the vertical angles equal to one another.
Let the straight lines AB and CD cut one another at the point E.
I say that the angle CEA equals the angle DEB, and the angle BEC equals the angle AED.
Since the straight line AE stands on the straight line CD making the
angles CEA and AED, therefore the sum of the angles CEA and AED I.13
equals two right angles.
Again, since the straight line DE stands on the straight line AB making
the angles AED and DEB, therefore the sum of the angles AED and I.13
DEB equals two right angles.
But the sum of the angles CEA and AED was also proved equal to two right angles, therefore Post.4
the sum of the angles CEA and AED equals the sum of the angles AED and DEB. Subtract the C.N.1
angle AED from each. Then the remaining angle CEA equals the remaining angle DEB. C.N.3
Similarly it can be proved that the angles BEC and AED are also equal.
Therefore if two straight lines cut one another, then they make the vertical angles equal to one
another.
Q.E.D.
Corollary
From this it is manifest that, if two straight lines cut one another, then they make the angles at the point
of section equal to four right angles.
Although the term "vertical angles" is not defined in the list of definitions at the beginning of Book I, its
meaning is clear form its use in this proposition.
A corollary that follows a proposition is a statement that immediately follows from the proposition or
the proof in the proposition. It is possible that this and the other corollaries in the Elements are
interpolations inserted after Euclid wrote the Elements. During the writing, he could have either bundled
the corollary into the proposition or made it a separate proposition.
Notes
Proclus includes another corollary: If any number of straight lines intersect one another at one point,
then the sum of all the angles so formed equals four right angles.
Use of Proposition 15
This proposition is used in the next one, a few others in this book, II.10, IV.15
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 16
In any triangle, if one of the sides is produced, then the exterior angle is greater than either of
the interior and opposite angles.
Let ABC be a triangle, and let one side of it BC be produced to D.
I say that the exterior angle ACD is greater than either of the interior and opposite angles CBA
and BAC.
I.10
Bisect AC at E. Join BE, and produce it in a straight Post.1
line to F. Post.2
I.3
Make EF equal to BE, join FC, and draw AC through Post.1
to G. Post.2
Since AE equals EC, and BE equals EF, therefore the two sides AE and EB equal the two sides
CE and EF respectively, and the angle AEB equals the angle FEC, for they are vertical angles.
I.15
Therefore the base AB equals the base FC, the triangle ABE equals the triangle CFE, and the I.4
remaining angles equal the remaining angles respectively, namely those opposite the equal
sides. Therefore the angle BAE equals the angle ECF.
But the angle ECD is greater than the angle ECF, therefore the angle ACD is greater than the C.N.5
angle BAE.
Similarly, if BC is bisected, then the angle BCG, that is, the angle ACD, can also be proved to I.15
be greater than the angle ABC.
Therefore in any triangle, if one of the sides is produced, then the exterior angle is greater than either
of the interior and opposite angles.
Q.E.D.
In the later proposition I.32, after he invokes the parallel postulate Post.5, Euclid shows the stronger
result that the exterior angle of a triangle equals the sum of the interior, opposite angles.
Elliptic geometry
There are geometries besides Euclidean geometry. Two of the more important geometries are elliptic
geometry and hyperbolic geometry, which were developed in the nineteenth century. The first 15
propositions in Book I hold in elliptic geometry, but not this one. (For more on hyperbolic geometry, see
the note after Proposition I.29.)
Plane elliptic geometry is closely related to spherical geometry, but it differs in that antipodal points on
the sphere are identified. Thus, a "point" in an elliptic plane is a pair of antipodal points on the sphere. A
"straight line" in an elliptic plane is an arc of great circle on the sphere. When a "straight line" is
extended, its ends eventually meet so that, topologically, it becomes a circle. This is very different from
Euclidean geometry since here the ends of a line never meet when extended.
Elliptic geometry satisfies some of the postulates of Euclidean geometry, but not all of them under all
interpretations. Usually, Post.1, to draw a straight line from any point to any point, is interpreted to
include the uniqueness of that line. But in elliptic geometry a completed "straight line" is topologically a
circle so that any pair of points on it divide it into two arcs. Therefore, in elliptic geometry exactly two
"straight lines" join any two given "points."
Also, Post.2, to produce a finite straight line continuously in a straight line, is sometimes interpreted to
include the condition that its ends don't meet when extended. Under that interpretation, elliptic geometry
fails Postulate 2.
Elliptic geometry fails Post.5, the parallel postulate, as well, since any two "straight lines" in an elliptic
plane meet. That is, any two great circles on the sphere meet at a pair of antipodal points.
Finally, a completed "straight line" in the elliptic plane does not divide the plane into two parts as
infinite straight lines do in the Euclidean plane. A completed "straight line" in the elliptic plane is a great
circle on the sphere. Any two "points" not on that "straight line" include two points in the same
hemisphere, and they can be joined by an arc that doesn't meet the great circle. Therefore two "points"
lie on the same side of the completed "straight line."
The proof of this particular proposition fails for elliptic geometry, and the statement of the proposition is
false for elliptic geometry. In particular, the statement "the angle ECD is greater than the angle ECF" is
not true of all triangles in elliptic geometry. The line CF need not be contained in the angle ACD. All the
previous propositions do hold in elliptic geometry and some of the later propositions, too, but some need
different proofs.
Use of Proposition 16
This proposition is used in the proofs of the next two propositions, a few others in this book, and a
couple in Book III.
Book I introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 17
In any triangle the sum of any two angles is less than two right angles.
Let ABC be a triangle.
I say that the sum of any two angles of the triangle ABC is less than two right angles.
Produce BC to D. Post.2
Since the angle ACD is an exterior angle of the triangle ABC, I.16
therefore it is greater than the interior and opposite angle ABC.
Add the angle ACB to each. Then the sum of the angles ACD and
C.N.
ACB is greater than the sum of the angles ABC and BCA.
But the sum of the angles ACD and ACB is equal to two right angles. Therefore the sum of the I.13
angles ABC and BCA is less than two right angles.
Similarly we can prove that the sum of the angles BAC and ACB is also less than two right
angles, and so the sum of the angles CAB and ABC as well.
Therefore in any triangle the sum of any two angles is less than two right angles.
Q.E.D.
The statements
... the angle ACD... is greater than the interior and opposite angle ABC.Add the angle ACBto
each. Then the sum of the angles ACDand ACBis greater than the sum of the angles ABCand BCA.
If x>y,then x + z>y + z.
This proposition is strengthened in Proposition I.32 to say the sum of all three angles in a triangle equals
two right angles.
Use of Proposition 17
This proposition is used in III.16 and a couple other propositions of Books III, and a few in Books VI
and XI.
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 18
I say that the angle ABC is also greater than the angle BCA.
But the angle ADB equals the angle ABD, since the side AB equals AD, therefore the angle
ABD is also greater than the angle ACB. Therefore the angle ABC is much greater than the I.5
angle ACB.
Therefore in any triangle the angle opposite the greater side is greater.
Q.E.D.
On word order
In this translation of Euclid's Elements the order of the words differs from the original Greek. In each of
Euclid's Greek sentences, the data, that is the geometric objects given or already constructed, appear
first, and the remaining geometric objects appear later. This is possible in Greek since it is an inflected
language and the word order is very flexible. On the other hand, the word order in English is intrinsic to
the syntax and semantics of the sentence and is not very flexible.
Take, for instance, the statements of this and the next proposition. Very literal translations of these are
(I.18) "In any triangle, the greater side [as subject] the greater angle [as object] subtends," and (I.19) "In
any triangle, the greater angle [as object] the greater side [as subject] subtends."
Heath keeps the word order in his translation but makes the second statement passive: (I.18) "In any
triangle the greater side subtends the greater angle," and (I.19) "In any triangle the greater angle is
subtended by the greater side." Without the understanding that the data come first, these two sentences
are logically equivalent.
In this translation the original word order is abandoned in order to make for more readable sentences and
to clarify the meaning. Thus, (I.18) "In any triangle the angle opposite the greater side is greater," and
(I.19) "In any triangle the side opposite the greater angle is greater."
It may sound like these two propositions really do say the same thing, but they don't. They're actually
disguised converses of each other. I.18 says "if side AC > side AB, then angle ABC > angle BCA" (but it
hasn't yet been shown that there is no other way for angle ABC to be greater), while I.19 says "if angle
ABC > angle BCA, then side AC > side AB."
Use of Proposition 18
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 19
Now AC does not equal AB, for then the angle ABC would equal the I.5
angle ACB, but it does not. Therefore AC does not equal AB.
Neither is AC less than AB, for then the angle ABC would be less than I.18
the angle ACB, 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 in any triangle the side opposite the greater angle is greater.
Q.E.D.
As mentioned before, this proposition is a disguised converse of the previous one. As Euclid often does,
he uses a proof by contradiction involving the already proved converse to prove this proposition. It is not
that there is a logical connection between this statement and its converse that makes this tactic work, but
some kind of symmetry involved. In this case, if one side is less than another, then the other is greater
than the one, and the previous proposition applies. So the relevant symmetry is between "less" and
"greater."
Although some of the geometric underpinnings of trigonometry appear in the Elements, trigonometry
itself does not. Trigonometry makes its appearance among later Greek mathematics where the the basic
trigonometric function is the chord, which is related to the sine.
Without going into details, the law of sines contains more precise
information about the relation between angles and sides of a
triangle than this and the last proposition did. The law of sines
states that
In other words, the sine of an angle in a triangle is proportional to the opposite side. (Proportions aren't
defined in the Elements until Book V.)
Use of Proposition 19
This proposition is used in the proofs of propositions I.20, I.24, and some others in Book III.
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 20
In any triangle the sum of any two sides is greater than the remaining one.
Let ABC be a triangle.
I say that in the triangle ABC the sum of any two sides is greater than the remaining one, that
is, the sum of BA and AC is greater than BC, the sum of AB and BC is greater than AC, and the
sum of BC and CA is greater than AB.
Post.2
Draw BA through to the point D, and make DA equal to CA. Join DC. I.3
Post.1
Since DA equals AC, therefore the angle ADC also equals the angle I.5
ACD. Therefore the angle BCD is greater than the angle ADC. C.N.5
Since DCB is a triangle having the angle BCD greater than the angle
BDC, and the side opposite the greater angle is greater, therefore DB I.19
is greater than BC.
But DA equals AC, therefore the sum of BA and AC is greater than BC.
Similarly we can prove that the sum of AB and BC is also greater than CA, and the sum of BC
and CA is greater than AB.
Therefore in any triangle the sum of any two sides is greater than the remaining one.
Q.E.D.
This proposition is known as "the triangle inequality." It is part of the statement that the shortest path
between two points is a straight line, but there are many other conceivable paths besides broken lines.
A minimum distance
This proposition on the triangle inequality, along with I.15 on vertical angles, allows us to solve a
problem on minimum distance, described and solved by Heron of Alexandria.
The solution is that the shortest path will be the path AEB
where angle of incidence, namely, angle AEC, equals the
angle of reflection, namely, angle BED.
First, we should show how to construct the bent line where the angle of incidence equals the angle of
reflection. Draw a perpendicular BF from the point B to the line CD (I.12), and extend it to B' so that
FB' = BF (Post.2, I.3). Draw AB' and let E be the point where AB' intersects CD. (There will be a point
of intersection since A and B' are on opposite sides of CD.) Draw BE.
Now, triangles BFE and B'FE are congruent since they have two sides and the included angle equal
(I.4), the included angles being right angles. Therefore, angles BFE and BF'E are equal. The vertical
angle AEC across from angle B'ED also equals these angles (I.15). Thus, the angle of incidence AEC
equals the angle of reflection BED.
We still have to show that the distance AE + EB is less than any distance AP + EP for any point P other
than E that lies on the line CD. Let P be such a point and draw lines AP, BP, and B'P. Then by
proposition I.20, above, AP + EP is less than AB'. But AB' = AE + EB', and EB' = EB, therefore AP + EP
is less than AE + EB.
Thus, the shortest bent line between two points on the same side of a line that meets that line is the one
where the angle of incidence equals the angle of reflection.
Q.E.D.
Use of Proposition 20
This proposition is used in the next two propositions, several in Book III, and XI.20.
Book I introduction
© 1996, 2002
D.E.Joyce
Clark University
Proposition 21
If from the ends of one of the sides of a triangle two straight lines are constructed meeting
within the triangle, then the sum of the straight lines so constructed is less than the sum of the
remaining two sides of the triangle, but the constructed straight lines contain a greater angle
than the angle contained by the remaining two sides.
From the ends B and C of one of the sides BC of the triangle ABC, let the two straight lines BD
and DC be constructed meeting within the triangle.
I say that the sum of BD and DC is less than the sum of the remaining two sides of the triangle
BA and AC, but BD and DC contain an angle BDC greater than the angle BAC.
Since in any triangle the sum of two sides is greater than the
remaining one, therefore, in the triangle ABE, the sum of the I.20
two sides AB and AE is greater than BE.
Again, since, in the triangle CED, the sum of the two sides CE and ED is greater than CD, add I.20
DB to each, therefore the sum of CE and EB is greater than the sum of CD and DB. C.N.
But the sum of BA and AC was proved greater than the sum of BE and EC, therefore the sum C.N.
of BA and AC is much greater than the sum of BD and DC.
Again, since in any triangle the exterior angle is greater than the interior and opposite angle, I.16
therefore, in the triangle CDE, the exterior angle BDC is greater than the angle CED.
For the same reason, moreover, in the triangle ABE the exterior angle CEB is greater than the
I.16
angle BAC. But the angle BDC was proved greater than the angle CEB, therefore the angle C.N.
BDC is much greater than the angle BAC.
Therefore if from the ends of one of the sides of a triangle two straight lines are constructed meeting
within the triangle, then the sum of the straight lines so constructed is less than the sum of the
remaining two sides of the triangle, but the constructed straight lines contain a greater angle than the
angle contained by the remaining two sides.
Q.E.D.
Pappus and others before him noticed that if the lines are not drawn from the ends of the side, then the
sum of the the constructed straight lines can be greater than the sum of the remaining two sides of the
triangle. In fact that sum can be made almost as large as twice the longest side of the triangle.
Use of Proposition 21
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 22
To construct a triangle out of three straight lines which equal three given straight
lines: thus it is necessary that the sum of any two of the straight lines should be I.20
greater than the remaining one.
Let the three given straight lines be A, B, and C, and let the sum of any two of these be
greater than the remaining one, namely, A plus B greater than C, A plus C greater than B, and
B plus C greater than B.
I say that the triangle KFG has been constructed out of three straight lines equal to A, B, and
C.
I.
Since the point F is the center of the circle DKL, therefore FD equals FK. But FD equals A, Def.16
therefore KF also equals A. C.N.1
I.
Again, since the point G is the center of the circle LKH, therefore GH equals GK. But GH Def.16
equals C, therefore KG also equals C. C.N.1
And FG also equals B, therefore the three straight lines KF, FG, and GK equal the three
straight lines A, B, and C.
Therefore out of the three straight lines KF, FG, and GK, which equal the three given
straight lines A, B, and C, the triangle KFG has been constructed.
Q.E.F.
The qualifier in the statement of the proposition, "thus it is necessary that the sum of any two of the
straight lines should be greater than the remaining one," refers to the triangle inequality, Proposition
I.20. This condition is, indeed, necessary. It is also sufficient, but Euclid failed to show that sufficiency.
This construction is actually a generalization of the very first proposition I.1 in which the three lines are
all equal. There too, as was noted, Euclid failed to prove that the two circles intersected.
Use of Proposition 22
The construction in this proposition is used for the construction in proposition I.23. It is also used in
XI.22
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 23
To construct a rectilinear angle equal to a given rectilinear angle on a given straight line and
at a point on it.
Let the angle DCE be the given rectilinear angle, AB the given straight line, and A the point on
it.
Since the two sides DC and CE equal the two sides FA and AG respectively, and the base DE I.8
equals the base FG, therefore the angle DCE equals the angle FAG.
Therefore on the given straight line AB, and at the point A on it, the rectilinear angle FAG has
been constructed equal to the given rectilinear angle DCE.
Q.E.F.
As Proclus and Heath point out, a very minor variant of the construction in I.22 is needed to make the
triangle AFG. The problem is that in I.22 the triangle is placed not at the end of line, but somewhere
beyond that, and in I.23, the triangle needs to be placed right at the end A of the line.
Construction steps
In order to make CE equal to CD, one circle is required. Next, in order to transfer the distance CD to A,
four circles (not shown) are required as per Propositions I.2 and I.1, and four more circles (also not
shown) to transfer ED to G. Finally, two more circles are required, one with center A and radius CD
(which has been transferred), the other with center G and radius ED. These last two circles meet at the
point F, and the line AF is the other side of the required angle.
In all there are ten circles and one line that must be drawn. The lines in the intermediate stages may be
suppressed as usual since they're only needed to verify the construction is correct.
Use of Proposition 23
The construction in this proposition is used in the next one and a couple others in Book I. It is also used
frequently in the later books. It is also used frequently in BookS III and VI and occasionally in Books IV
and XI.
Although it may appear that the triangles are to be in the same plane, that is not necessary. Indeed, the
construction in this proposition is used to construct an angle in a different plane in proposition XI.31.
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 24
If two triangles have two sides equal to two sides respectively, but have one of the angles
contained by the equal straight lines greater than the other, then they also have the base
greater than the base.
Let ABC and DEF be two triangles having the two sides AB and AC equal to the two sides DE
and DF respectively, so that AB equals DE, and AC equals DF, and let the angle at A be
greater than the angle at D.
Again, since DF equals DG, therefore the angle DGF equals the angle DFG. Therefore the I.5
angle DFG is greater than the angle EGF.
Therefore the angle EFG is much greater than the angle EGF.
Since EFG is a triangle having the angle EFG greater than the angle EGF, and side opposite I.19
the greater angle is greater, therefore the side EG is also greater than EF.
Therefore if two triangles have two sides equal to two sides respectively, but have one of the angles
contained by the equal straight lines greater than the other, then they also have the base greater than
the base.
Q.E.D.
Use of Proposition 24
This proposition is used in the next proposition as well as a few in Book III and XI.22.
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 25
If two triangles have two sides equal to two sides respectively, but have the base greater than
the base, then they also have the one of the angles contained by the equal straight lines greater
than the other.
Let ABC and DEF be two triangles having two sides AB and AC equal to two sides DE and DF
respectively, namely AB to DE, and AC to DF, and let the base BC be greater than the base EF.
I say that the angle BAC is also greater than the angle EDF.
Now the angle BAC does not equal the angle EDF, for
then the base BC would equal the base EF, but it is not. I.4
Therefore the angle BAC does not equal the angle EDF.
Neither is the angle BAC less than the angle EDF, for then
the base BC would be less than the base EF, but it is not. I.24
Therefore the angle BAC 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 if two triangles have two sides equal to two sides respectively, but have the base greater
than the base, then they also have the one of the angles contained by the equal straight lines greater
than the other.
Q.E.D.
The conclusions of this proposition and the previous are partial converses of each other. Together they
say that if two triangles have two sides equal to two sides respectively, then the base greater than the
base if and only if the one of the angles contained by the equal straight lines greater than the other.
Use of Proposition 25
This proposition is not used in the rest of Book I, but it is used in XI.20 and XI.23.
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 26
If two triangles have two angles equal to two angles respectively, and one side equal to one
side, namely, either the side adjoining the equal angles, or that opposite one of the equal
angles, then the remaining sides equal the remaining sides and the remaining angle equals the
remaining angle.
Let ABC and DEF be two triangles having the two angles ABC and BCA equal to the two
angles DEF and EFD respectively, namely the angle ABC to the angle DEF, and the angle
BCA to the angle EFD, and let them also have one side equal to one side, first that adjoining
the equal angles, namely BC equal to EF.
I say that the remaining sides equal the remaining sides respectively, namely AB equals DE
and AC equals DF, and the remaining angle equals the remaining angle, namely the angle BAC
equals the angle EDF.
Since BG equals DE, and BC equals EF, the two sides GB and BC equal the two sides DE and
EF respectively, and the angle GBC equals the angle DEF, therefore the base GC equals the I.4
base DF, the triangle GBC equals the triangle DEF, and the remaining angles equal the
remaining angles, namely those opposite the equal sides. Therefore the angle GCB equals the
C.N.1
angle DFE. But the angle DFE equals the angle ACB by hypothesis. Therefore the angle BCG
equals the angle BCA, the less equals the greater, which is impossible.
But BC also equals EF. Therefore the two sides AB and BC equal the two sides DE and EF
respectively, and the angle ABC equals the angle DEF. Therefore the base AC equals the base I.4
DF, and the remaining angle BAC equals the remaining angle EDF.
Since BH equals EF, and AB equals DE, the two sides AB and BH equal the two sides DE and
EF respectively, and they contain equal angles, therefore the base AH equals the base DF, the I.4
triangle ABH equals the triangle DEF, and the remaining angles equal the remaining angles,
namely those opposite the equal sides. Therefore the angle BHA equals the angle EFD.
C.
But the angle EFD equals the angle BCA, therefore, in the triangle AHC, the exterior angle N.1
BHA equals the interior and opposite angle BCA, which is impossible. I.16
But AB also equals DE. Therefore the two sides AB and BC equal the two sides DE and EF
respectively, and they contain equal angles. Therefore the base AC equals the base DF, the I.4
triangle ABC equals the triangle DEF, and the remaining angle BAC equals the remaining
angle EDF.
Therefore if two triangles have two angles equal to two angles respectively, and one side equal to one
side, namely, either the side adjoining the equal angles, or that opposite one of the equal angles, then
the remaining sides equal the remaining sides and the remaining angle equals the remaining angle.
Q.E.D.
There are two statements in this theorem which are different only in their hypotheses. In one, the known
side lies between the two angles, in the other, the known side lies opposite one of the angles. If this
proposition had come after proposition I.32 which states the sum of the angles in a triangle equals two
right angles, then these two hypotheses could have been merged into one, since then if two angles are
known, then is the third. But proposition I.32 depends on the parallel postulate Post.5, which, it is
apparent, Euclid did not want to use unless necessary. Thus, this proposition, I.26, appears where it is
with two distinct hypotheses.
On congruence theorems
This is the last of Euclid's congruence theorems for triangles. Euclid's congruence theorems are I.4 (side-
angle-side), I.8 (side-side-side), and this one, I.26 (side and two angles). Calling them congruence
theorems is anachronistic, since Euclid did not explicitly use the concept of congruence. We would say
that two triangles ABC and DEF are congruent if the angles A, B, and C equal the angles D, E, and F
respectively, and the sides AB, BC, and AC equal the sides DE, EF, and DF respectively, and the triangle
ABC equals the triangle DEF (by which is meant that they have the same area).
The remaining congruence theorem, side-side-angle, includes some ambiguous cases. Suppose triangles
ABC and DEF are such that sides AB and BC are equal to sides DE and EF respectively, and angle A
equals angle D. If it is also known that AB is less than or equal to BC, then it follows that the two
triangles are congruent. If, however, AB is greater than BC, then the two triangles need not be congruent.
Euclid does not include any form of a side-side-angle congruence theorem, but he does prove one
special case, side-side-right angle, in the course of the proof of proposition III.14.
Although Euclid does not include a side-side-angle congruence theorem, he does have a side-side-angle
similarity theorem, namely proposition VI.7. The analogous congruence theorem could be stated as
follows: If two triangles have one angle equal to one angle, two sides adjoining the equal angles equal,
namely, one side adjoining the equal angles, and one opposite the equal angles, and the remaining angles
either both less or both not less than a right angle, then the remaining side equals the remaining side and
the remaining angles equal the remaining angles.
Use of Proposition 26
This proposition is used in the proofs of proposition I.34 and several propositions in Books III, IV, XI,
XII, and XIII.
As in propositions I.4 and I.8, it appears that the triangles are in the same plane, but, again, that is not
necessary. Indeed, this proposition is invoked in proposition XI.35 when two triangles do not lie in the
same plane.
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 27
If a straight line falling on two straight lines makes the alternate angles equal to one another,
then the straight lines are parallel to one another.
Let the straight line EF falling on the two straight lines AB and CD make the alternate
angles AEF and EFD equal to one another.
Then, in the triangle GEF, the exterior angle AEF equals the interior and opposite angle I.16
EFG, which is impossible.
But straight lines which do not meet in either direction are parallel. Therefore AB is I.Def.23
parallel to CD.
Therefore if a straight line falling on two straight lines makes the alternate angles equal to one
another, then the straight lines are parallel to one another.
Q.E.D.
There is implicitly assumed an ambient plane. The term "alternate angles" doesn't have a meaning unless
the lines all lie in a plane.
Note that Euclid does not consider two other possible ways that the two lines could meet, namely, in the
directions A and D or toward B and C.
Although this is the first proposition about parallel lines, it does not require the parallel postulate Post.5
as an assumption. This proposition I.27 and the parallel postulate can be made to look more similar if
they are reworded (with the help of I.13).
Proposition 1.27.
If a straight line falls on two straight lines, then if the alternate angles are equal, then the straight
lines do not meet.
Post.5.
If a straight line falls on two straight lines, then if the alternate angles are not equal, then the
straight lines meet [on a certain side of the line].
If the remark about the side is dropped, then the conclusions are logical inverses of each other, and the
logical inverse of a statement is logically equivalent to the converse.
This little table summarizes the logical relations between similarly looking statements.
Statement If P then Q.
Converse If Q then P. Not logically equivalent to the statement.
Contrapositive If not Q then not P. Logically equivalent to the statement.
Inverse If not P then not Q. Logically equivalent to the converse.
Although the contrapositive is logically equivalent to the statement, Euclid always proves the
contrapositive separately using a proof by contradiction and the original statement. Similarly, the inverse
is proved using the converse. Sometimes all four statements appear in separate propositions as in
propositions X.5 through X.8. Other times the four appear as four statements in one proposition as in
X.9. More often than not, however, the contrapositive and inverse make no appearance, and, of course,
the converse only appears when it can be proved.
Use of Proposition 27
At this point, parallel lines have yet to be constructed. That occurs in proposition I.31 which uses this
proposition to verify that lines constructed there are parallel. This proposition is also used in the next
Book I introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 28
If a straight line falling on two straight lines makes the exterior angle equal to the interior and
opposite angle on the same side, or the sum of the interior angles on the same side equal to two
right angles, then the straight lines are parallel to one another.
Let the straight line EF falling on the two straight lines AB and CD make the exterior angle
EGB equal to the interior and opposite angle GHD, or the sum of the interior angles on the
same side, namely BGH and GHD, equal to two right angles.
Since the angle EGB equals the angle GHD, and the angle EGB I.15
equals the angle AGH, therefore the angle AGH equals the C.
angle GHD. And they are alternate, therefore AB is parallel to N.1
CD. I.27
Next, since the sum of the angles BGH and GHD equals two I.13
right angles, and the sum of the angles AGH and BGH also C.
equals two right angles, therefore the sum of the angles AGH N.1
and BGH equals the sum of the angles BGH and GHD. Post.4
C.
Subtract the angle BGH from each. Therefore the remaining angle AGH equals the remaining N.3
angle GHD. And they are alternate, therefore AB is parallel to CD. I.27
Therefore if a straight line falling on two straight lines makes the exterior angle equal to the interior
and opposite angle on the same side, or the sum of the interior angles on the same side equal to two
right angles, then the straight lines are parallel to one another.
Q.E.D.
This proposition states two useful minor variants of the previous proposition. The three statements differ
only in their hypotheses which are easily seen to be equivalent with the help of proposition I.13.
Use of Proposition 28
This proposition is used in IV.7, VI.4, and a couple times in Book XI.
Book I introduction
© 1996
D.E.Joyce
Clark University
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 sum of the interior
angles on the same side equal to two right angles.
Let the straight line EF fall on the parallel straight lines AB and CD.
I say that it makes the alternate angles AGH and GHD equal, the exterior angle EGB equal to
the interior and opposite angle GHD, and the sum of the interior angles on the same side,
namely BGH and GHD, equal to two right angles.
If the angle AGH does not equal the angle GHD, then one of
them is greater. Let the angle AGH be greater.
Add the angle BGH to each. Therefore the sum of the angles
AGH and BGH is greater than the sum of the angles BGH and
GHD.
But sum of the angles AGH and BGH equals two right angles. Therefore the sum of the angles I.13
BGH and GHD is less than two right angles.
But straight lines produced indefinitely from angles less than two right angles meet. Therefore
AB and CD, if produced indefinitely, will meet. But they do not meet, because they are by Post.5
hypothesis parallel.
Therefore the angle AGH is not unequal to the angle GHD, and therefore equals it.
Again, the angle AGH equals the angle EGB. Therefore the angle EGB also equals the angle I.15
GHD. C.N.1
Add the angle BGH to each. Therefore the sum of the angles EGB and BGH equals the sum of C.N.2
the angles BGH and GHD.
But the sum of the angles EGB and BGH equals two right angles. Therefore the sum of the I.13
angles BGH and GHD also equals two right angles. C.N.1
Therefore 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 sum of the interior angles
on the same side equal to two right angles.
Q.E.D.
The statement of this proposition includes three parts, one the converse of I.27, the other two the
converse of I.28. Like those propositions, this one assumes an ambient plane containing all the three
lines.
This is the first proposition which depends on the parallel postulate. As such it does not hold in
hyperbolic geometry.
Hyperbolic geometry
Two important geometries alternative to Euclidean geometry are elliptic geometry and hyperbolic
geometry. Elliptic geometry was discussed in the note after Proposition I.16, that being the first
proposition which doesn't hold in elliptic geometry. This, I.29, is the first which doesn't hold in
hyperbolic geometry.
These three geometries can be distinguished by the number of lines parallel to a given line passing
through a given point. For elliptic geometry, there is no such parallel line; for Euclidean geometry
(which may be called parabolic geometry), there is exactly one; and for hyperbolic geometry, there are
infinitely many.
It is not possible to illustrate hyperbolic geometry with correct distances on a flat surface since a flat
surface is Euclidean. Poincaré, however, described a useful model of hyperbolic geometry where the
"points" in a hyperbolic plane are taken to be points inside a fixed circle (but not the points on the
circumference). The "lines" in the hyperbolic plane are the parts of circles orthogonal, that is, at right
angles to the fixed circle. And in this model, "angles" in the hyperbolic plane are angles between these
arcs, or, more precisely, angles between the tangents to the arcs at the point of intersection. Since
"angles" are just angles, this model is called a conformal model. Distances in the hyperbolic plane,
however, are not measured by distances along the arcs. There is a more complicated relation between
distances so that near the edge of the fixed circle a very short arc models a very long "line."
In the diagram, AB is a "line" in the hyperbolic plane, that is, a circle orthogonal to the circumference of
the shaded disk which represents the hyperbolic plane. A "point" C lies in that plane. Two "lines" are
shown passing through C, one gets close to the line AB in the direction of A, the other gets close in the
direction of B. But these two "lines" don't intersect AB since the arcs representing them only intersect on
the circumference of the disk, and points on the circumference don't represent "points" in the hyperbolic
plane.
These two parallel "lines" are called the asymptotic parallels of AB since they approach AB at one end or
the other. There are infinitely many parallels between them. (In much of the literature on hyperbolic
geometry, the word "parallels" is used for what are called "asymptotic parallels" here, while
"nonintersecting lines" is used for what are called "parallels" here.)
Use of Proposition 29
This proposition is used in very frequently in Book I starting with the next proposition. It is also used
frequently in Book II, VI, and XI, and once in Book XII.
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 30
Straight lines parallel to the same straight line are also parallel to one another.
Let each of the straight lines AB and CD be parallel to EF.
Let the straight line GK fall upon them. Since the straight
line GK falls on the parallel straight lines AB and EF, I.29
therefore the angle AGK equals the angle GHF.
But the angle AGK was also proved equal to the angle GHF. Therefore the angle AGK also C.N.1
equals the angle GKD, and they are alternate.
Therefore straight lines parallel to the same straight line are also parallel to one another.
Q.E.D.
For this proposition it is supposed that the three lines lie in one plane. Proposition XI.9 applies to the
case where the three lines do not lie in a plane.
Playfair's axiom
A number of the propositions in the Elements are equivalent to the parallel postulate Post.5 in the sense
that if the rest of the postulates are assumed and any one of these propositions is assumed, then the
parallel postulate can be proved as a proposition. This one I.30, the last I.29, either part of I.32, and
almost any later one. Thus, Euclid had many statements to choose from to take as a postulate.
In many modern expositions of synthetic geometry, Playfair's axiom (John Playfair, 1748-1819) is
chosen as that postulate instead of Euclid's parallel postulate Post.5. Playfair's axiom states that there is
at most one line parallel to a given line passing through a given point. (That there is at least one follows
from the next proposition I.31 which doesn't depend on the parallel postulate.)
Two advantages of Playfair's axiom over Euclid's parallel postulate are that it is a simpler statement, and
it emphasizes the distinction between Euclidean and hyperbolic geometry.
Two disadvantages are that it does not have the historical importance of Euclid's parallel postulate, and
the proof of the parallel postulate from Playfair's axiom is nonconstructive. That proof is a proof by
contradiction that begins assuming that a point does not exist, deriving a contradiction, and concluding
that the point must exist, but does not construct it. It may well be that Euclid chose to make the
construction an assumption of his parallel postulate rather rather than choosing some other equivalent
statement for his postulate.
Elegance in mathematics
Euclid's Elements form one of the most beautiful works of science in the history of humankind. This
beauty lies more in the logical development of geometry rather than in geometry itself. It is not the
diagrams that excite our interest; rather it is the concepts, the way the concepts interconnect, and the way
Euclid selected and presented these concepts and their interconnections. The Elements are elegant.
Elegance in mathematics is characterized by simplicity and clarity. An elegant presentation is easy for
the reader to follow. But elegance is not only in the presentation, it is in the selection of definitions and
proofs. The elegant definition is the one that makes the rest of the theory easy. The elegant proof is the
one that is easiest to follow, one that is designed just right to fit the goal. Extraneous concepts should not
be involved. Even the goals need to be adjusted to the right level of generality to cover the concepts, but
not so abstract that the abstraction itself obscures the goal.
One of the criticisms of Euclid's parallel postulate was that it isn't simple. The statement of this
proposition, I.30, is much simpler, and Playfair's axiom is much simpler. As they're each logically
equivalent to Euclid's parallel postulate, if elegance were the primary goal, then Euclid would have
chosen one of them in place of his postulate. Perhaps the reasons mentioned above explain why Euclid
used Post.5 instead.
Use of Proposition 30
Book I introduction
Proposition 31
To draw a straight line through a given point parallel to a given straight line.
Let A be the given point, and BC the given straight line.
Take a point D at random on BC. Join AD. Construct the angle DAE equal to the angle ADC Post.1
on the straight line DA and at the point A on it. Produce the straight line AF in a straight line I.23
with EA. Post.2
Since the straight line AD falling on the two straight lines BC and EF makes the alternate I.27
angles EAD and ADC equal to one another, therefore EAF is parallel to BC.
Therefore the straight line EAF has been drawn through the given point A parallel to the given
straight line BC.
Q.E.F.
The parallel line EF constructed in this proposition is the only one passing through the point A. If there
were another, then the interior angles on one side or the other of AD it makes with BC would be less
than two right angles, and therefore by the parallel postulate Post.5, it would meet BC, a contradiction.
Incidentally, this construction also works in hyperbolic geometry, although different parallel lines
through A are constructed for different points D.
Construction steps
The construction needed is that of I.23 to construct an angle. That construction required ten circles and
one line in general. In the specific case needed here, however, one of the distances does not have to be
transferred, and that eliminates the need to construct four of the circles. Therefore this construction
Use of Proposition 31
This construction is frequently used in the remainder of Book I starting with the next proposition. It is
also frequently used in Books II, IV, VI, XI, XII, and XIII.
Book I introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 32
In any triangle, if one of the sides is produced, then the exterior angle equals the sum of the
two interior and opposite angles, and the sum of the three interior angles of the triangle equals
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 ACD equals the sum of the two interior and
opposite angles CAB and ABC, and the sum of the three interior angles of the
triangle ABC, BCA, and CAB equals two right angles.
Again, since AB is parallel to CE, and the straight line BD falls upon them, I.29
therefore the exterior angle ECD equals the interior and opposite angle ABC.
But the angle ACE was also proved equal to the angle BAC. Therefore the
whole angle ACD equals the sum of the two interior and opposite angles BAC
and ABC.
Add the angle ACB to each. Then the sum of the angles ACD and ACB equals C.N.2
the sum of the three angles ABC, BCA, and CAB.
But the sum of the angles ACD and ACB equals two right angles. Therefore I.13
the sum of the angles ABC, BCA, and CAB also equals two right angles. C.N.1
Therefore in any triangle, if one of the sides is produced, then the exterior angle equals the sum of the
two interior and opposite angles, and the sum of the three interior angles of the triangle equals two
right angles.
Q.E.D.
Corollaries of Proclus
Corollary 1. The sum of the interior angles of a convex rectilinear figure equals twice as many angles as
the figure has sides, less four.
Corollary 2. The sum of the exterior angles of any convex rectilinear figure together equal four right
angles.
Use of Proposition 32
Although this proposition isn't used in the rest of Book I, it is frequently used in the rest of the books on
geometry, namely Books II, III, IV, VI, XI, XII, and XIII. The corollaries, however, are not used in the
Elements.
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 33
Straight lines which join the ends of equal and parallel straight lines in the same directions are
themselves equal and parallel.
Let AB and CD be equal and parallel, and let the straight lines AC
and BD join them at their ends in the same directions.
Since AB is parallel to CD, and BC falls upon them, therefore the alternate angles ABC and I.29
BCD equal one another.
Since AB equals CD, and BC is common, the two sides AB and BC equal the two sides DC and
CB, and the angle ABC equals the angle BCD, therefore the base AC equals the base BD, the
triangle ABC equals the triangle DCB, and the remaining angles equals the remaining angles I.4
respectively, namely those opposite the equal sides. Therefore the angle ACB equals the angle
CBD.
Since the straight line BC falling on the two straight lines AC and BD makes the alternate I.27
angles equal to one another, therefore AC is parallel to BD.
Therefore straight lines which join the ends of equal and parallel straight lines in the same directions
are themselves equal and parallel.
Q.E.D.
The qualifier "in the same directions" in the statement of this proposition is necessary since without it
the lines AD and BC could join the endpoints of the parallel lines, and AD and BC are not parallel but
intersect. But these words of Euclid words are informal, and it would take some work to determine
In general, given four points A, B, C, and D, exactly one of the three pairs of lines, AB and CD, AC and
BD, and AD and BC, intersects. (If extended to infinite lines, all three pairs of lines might intersect, but
as line segments only one pair does.) This statement belongs to the fundamental part of plane geometry
that includes betweenness and sides of lines that wasn't developed until the late nineteenth century.
Use of Proposition 33
This proposition is used in I.36, I.45, and a few prpositions in Books XI through XIII.
Book I introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 34
In parallelogrammic areas the opposite sides and angles equal one another, and the diameter
bisects the areas.
Let ACDB be a parallelogrammic area, and BC its diameter.
I say that the opposite sides and angles of the parallelogram ACDB equal one another, and the
diameter BC bisects it.
Therefore ABC and DCB are two triangles having the two angles ABC and BCA equal to the
two angles DCB and CBD respectively, and one side equal to one side, namely that adjoining
the equal angles and common to both of them, BC. Therefore they also have the remaining I.26
sides equal to the remaining sides respectively, and the remaining angle to the remaining
angle. Therefore the side AB equals CD, and AC equals BD, and further the angle BAC equals
the angle CDB.
Since the angle ABC equals the angle BCD, and the angle CBD equals the angle ACB, C.N.2
therefore the whole angle ABD equals the whole angle ACD.
And the angle BAC was also proved equal to the angle CDB.
Therefore in parallelogrammic areas the opposite sides and angles equal one another.
Since AB equals CD, and BC is common, the two sides AB and BC equal the two sides DC and
CB respectively, and the angle ABC equals the angle BCD. Therefore the base AC also equals I.4
DB, and the triangle ABC equals the triangle DCB.
Therefore in parallelogrammic areas the opposite sides and angles equal one another, and the
diameter bisects the areas.
Q.E.D.
In this proposition Euclid uses the term "parallelogrammic area" rather than the word "parallelogram"
which first occurs in the next proposition. Proclus indicated that the word "parallelogram" was created
by Euclid.
This proposition begins the study of areas of rectilinear figures. It is a modest beginning, but it allows
the comparison of triangles and parallelograms so that problems and results concerning one can be
converted to problems and results concerning the other.
Use of Proposition 34
This proposition is used in the next four propositions and some others in Book I, several in Book II, a
few in Books IV, VI, X, XI, and XII.
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 35
Parallelograms which are on the same base and in the same parallels equal one another.
Let ABCD and EBCF be parallelograms on the same base BC and in the same parallels AF and
BC.
For the same reason EF equals BC, so that AD also equals EF. And DE is common, therefore C.N.1
the whole AE equals the whole DF. C.N.2
Subtract DGE from each. Then the trapezium ABGD which remains equals the trapezium C.N.3
EGCF which remains.
Add the triangle GBC to each. Then the whole parallelogram ABCD equals the whole C.N.2
parallelogram EBCF.
Therefore parallelograms which are on the same base and in the same parallels equal one another.
Q.E.D.
We see how Euclid treats figures as magnitudes by adding as subtracting them. The triangles EAB and
FDC are shown directly to be equal. Then the triangle DGE, which is contained in each, is subtracted
from each, and Euclid concludes that the remaining trapezia ABGD and EGCF are therefore equal.
These trapezia are not congruent, but they do have the same area. Next, the triangle GBC is added to
each trapezium to conclude the two parallelograms ABCD and EBCF are equal.
These are the same kinds of cut-and-paste operations that Euclid used on lines and angles earlier in Book
I, but these are applied to rectilinear figures. In later books cut-and-paste operations will be applied to
other kinds of magnitudes such as solid figures and parts of circumferences of circles.
Use of Proposition 35
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 36
Parallelograms which are on equal bases and in the same parallels equal one another.
Let ABCD and EFGH be parallelograms which are on the equal bases BC and FG and in the
same parallels AH and BG.
But they are also parallel, and EB and HC join them. But straight lines joining equal and
parallel straight lines in the same directions are equal and parallel, therefore EBCH is a I.33
parallelogram.
And it equals ABCD, for it has the same base BC with it and is in the same parallels BC and I.35
AH with it.
For the same reason also EFGH equals the same EBCH, so that the parallelogram ABCD also C.N.1
equals EFGH.
Therefore parallelograms which are on equal bases and in the same parallels equal one another.
Q.E.D.
This proposition is a generalization of the previous proposition I.35, and its proof depends directly on it.
Euclid could have bundled the two propositions into one. Then the special case of I.35 would have been
proven first and then used to prove the general case of I.36. In an introductory book like Book I this
separation makes it easier to follow the logic, but in later books special cases are often bundled into the
general proposition.
Use of Proposition 36
This proposition is used in I.38, a few propositions in Books II and VI, and XI.29
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 37
Triangles which are on the same base and in the same parallels equal one another.
Let ABC and DBC be triangles on the same base BC and in the same parallels AD and BC.
Moreover the triangle ABC is half of the parallelogram EBCA, for the diameter AB bisects it. I.34
And the triangle DBC is half of the parallelogram DBCF, for the diameter DC bisects it.
Therefore triangles which are on the same base and in the same parallels equal one another.
Q.E.D.
In this proposition the triangles have the same base while in the next one the triangles have equal bases.
Since the proofs are the same except that this depends on I.35 while the next depends on I.36, and the
next is more general, there is no purpose to include this proposition.
The justification of the last conclusion is missing. From the statement that the doubles of two
magnitudes are equal, we want to conclude that the magnitudes themselves are equal. Although Euclid
included no such common notion, others inserted it later. See the commentary on Common Notions for a
proof of this halving principle based on other properties of magnitudes.
Use of Proposition 37
Book I introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 38
Triangles which are on equal bases and in the same parallels equal one another.
Let ABC and DEF be triangles on equal bases BC and EF and in the same parallels BF and
AD.
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.
Therefore triangles which are on equal bases and in the same parallels equal one another.
Q.E.D.
The idea of the argument is clear: since parallelograms on equal bases and in the same parallels are
equal by I.36, and the triangles are half the parallelograms by I.34, therefore the triangles are also equal.
Use of Proposition 38
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 39
Equal triangles which are on the same base and on the same side are also in the same parallels.
Let ABC and DBC be equal triangles which are on the same base BC and on the same side of
it. Join AD.
Post.1
I say that AD is parallel to BC.
If not, draw AE through the point A parallel to the straight line BC, I.31
and join EC. Post.1
Therefore the triangle ABC equals the triangle EBC, for it is on the I.37
same base BC with it and in the same parallels.
But ABC equals DBC, therefore DBC also equals EBC, the greater equals the less, which is C.N.1
impossible.
Similarly we can prove that neither is any other straight line except AD, therefore AD is
parallel to BC.
Therefore equal triangles which are on the same base and on the same side are also in the same
parallels.
Q.E.D.
This is a partial converse to proposition I.37, only partial since the two triangles ABC and DBC have to
be on the same side of the line BC. If they weren't, then of course AD would not be parallel to BC but
instead cross it at the midpoint
Use of Proposition 39
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 40
Equal triangles which are on equal bases and on the same side are also in the same parallels.
Let ABC and CDE be equal triangles on equal bases BC and CE and on the same side.
I.31
If not, draw AF through A parallel to BE, and join FE. Post.1
Therefore the triangle ABC equals the triangle FCE, for they are on equal bases BC and CE I.38
and in the same parallels BE and AF.
But the triangle ABC equals the triangle DCE, therefore the triangle DCE also equals the
triangle FCE, the greater equals the less, which is impossible. Therefore AF is not parallel to C.N.1
BE.
Similarly we can prove that neither is any other straight line except AD, therefore AD is
parallel to BE.
Therefore equal triangles which are on equal bases and on the same side are also in the same
parallels.
Q.E.D.
The setting out of this proposition is not up to Euclid's standards. There is no justification for assuming
that the point C is a common vertex of the two triangles. Fortunately, the proof works just as well if C is
For some of the propositions and many of the lemmas and corollaries in the Elements, there is evidence
that Euclid did not write them, but they were added later. The process of incorporating new material in
textbooks was almost automatic when the books were copied by hand instead of printed. Scholars wrote
comments (called "scholia") in the margins of the texts, and copyists (some of whom were later
scholars) would include those comments as part of the text in their new copies.
Heiberg could show by means of an early papyrus fragment that this proposition was an early
interpolation. For others, such as I.37 there is no direct evidence, only a doubt that a mathematician of
Euclid's caliber would have included them.
Unlike the other propositions in Book I, this one is not used later in the Elements.
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 41
If a parallelogram has the same base with a triangle and is in the same parallels, then the
parallelogram is double the triangle.
Let the parallelogram ABCD have the same base BC with the triangle EBC, and let it be in the
same parallels BC and AE.
Then the triangle ABC equals the triangle EBC, for it is on the I.37
same base BC with it and in the same parallels BC and AE.
But the parallelogram ABCD is double the triangle ABC, for the diameter AC bisects it, so that I.34
the parallelogram ABCD is also double the triangle EBC.
Therefore if a parallelogram has the same base with a triangle and is in the same parallels, then the
parallelogram is double the triangle.
Q.E.D.
This partially generalizes I.34, that a parallelogram is twice the triangle by its diameter and two of its
sides. A slightly more general statement would be that If a parallelogram has an equal base with a
triangle and is in the same parallels, then the parallelogram is double the triangle.
Use of Proposition 41
This proposition is used in the next one, I.47, VI.1, and X.38.
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 42
It is required to construct D a parallelogram equal to the triangle ABC in the rectilinear angle.
Since BE equals EC, therefore the triangle ABE also equals the triangle AEC, for they are on
equal bases BE and EC and in the same parallels BC and AG. Therefore the triangle ABC is I.38
double the triangle AEC.
But the parallelogram FECG is also double the triangle AEC, for it has the same base with it I.41
and is in the same parallels with it, therefore the parallelogram FECG equals the triangle ABC. C.N.1
Therefore the parallelogram FECG has been constructed equal to the given triangle ABC, in
the angle CEF which equals D.
Q.E.F.
The idea of the construction is as follows. First make a triangle half the size of the given triangle. Next
skew the half-size triangle to make one of its angles the desired angle without changing its area.
Complete the resulting half-size triangle to a parallelogram. That's the desired parallelogram equal to the
original triangle in the desired angle.
Application of areas
With this proposition Euclid moves to the next phase in his study of areas, the application of areas.
Before this, he has exhibited various situations when triangles or parallelograms have equal areas, or
when a triangle has half the area of a parallelogram. But now he's interested in constructing another
figure with the same area as a given figure.
In this proposition, he constructs a parallelogram that has a given angle and has the same area as a given
triangle. But his goals are coming up, application of areas in I.45 and quadrature in II.14. In proposition
I.45, given a rectilinear figure an equal parallelogram is constructed on a given side within a given
angle. This kind of construction is called "applying" an area to a side. The area is sort of laid along the
line. It may be that before Euclid the area was always applied to a rectangle along the line, but Euclid
generalized the construction to parallelograms. This extra generalization is not often used.
Later, in proposition II.14 a square is constructed equal to a given rectilinear figure, a process called
"quadrature" (making into a square) of the figure. This square is a canonical measure of the area.
Use of Proposition 42
This construction is used as part of the constructions in the two propositions following the next one.
Book I introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 43
In any parallelogram the complements of the parallelograms about the diameter equal one
another.
Let ABCD be a parallelogram, and AC its diameter, and about AC let EH and FG be
parallelograms, and BK and KD the so-called complements.
Now, since the triangle AEK equals the triangle AHK, and KFC equals KGC, therefore the C.N.2
triangle AEK together with KGC equals the triangle AHK together with KFC.
And the whole triangle ABC also equals the whole ADC, therefore the remaining complement C.N.3
BK equals the remaining complement KD.
Therefore in any parallelogram the complements of the parallelograms about the diameter equal one
another.
Q.E.D.
The meaning of the statement has to be found in its use. The term "the parallelograms about the
diameter" refers to the two parallelograms having the same angles as the original parallelogram and with
diameters AK and KC which are two parts of a diameter AC of the original parallelogram. The
"complements" are the two parallelograms left over after removing those two parallelograms from the
original parallelogram.
Use of Proposition 43
The immediate purpose of this proposition is to change the shape of a parallelogram (one of the
complements) into an equal parallelogram with the same angles (the other complement). That's how it is
used in the next proposition. It is also used in several propositions in Book II, and a couple in Book VI.
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 44
To a given straight line in a given rectilinear angle, to apply a parallelogram equal to a given
triangle.
Let AB be the given straight line, D the given rectilinear angle, and C the given triangle.
It is required to apply a parallelogram equal to the given triangle C to the given straight line
AB in an angle equal to D.
Post.2
Draw FG through to H, and draw AH through A I.31
parallel to either BG or EF. Join HB. Post.1
Since the straight line HF falls upon the parallels AH and EF, therefore the sum of the angles
AHF and HFE equals two right angles. Therefore the sum of the angles BHG and GFE is less I.29
than two right angles. And straight lines produced indefinitely from angles less than two right Post.5
angles meet, therefore HB and FE, when produced, will meet.
Let them be produced and meet at K. Draw KL through the point K parallel to either EA or I.31
FH. Produce HA and GB to the points L and M.
Then HLKF is a parallelogram, HK is its diameter, and AG and ME are parallelograms, and I.43
LB and BF are the so-called complements about HK. Therefore LB equals BF.
Since the angle GBE equals the angle ABM, while the angle GBE equals D, therefore the I.15
angle ABM also equals the angle D. C.N.1
Therefore the parallelogram LB equal to the given triangle C has been applied to the given
straight line AB, in the angle ABM which equals D.
Q.E.F.
There are two steps in this construction. The first uses proposition I.42 to construct some parallelogram
with the correct angle equal to the given triangle. The second uses I.43 to change its length to the proper
length.
To "apply an area to a line in an angle" means just what this construction accomplishes, namely, to
construct a parallelogram equal to that area with one side as the given line and one angle equal to the
given angle.
In practice the angle is often a right angle. The given line may be thought of as a "unit" line. Then the
length of the resulting rectangle represents the the area.
Use of Proposition 44
Besides being used in the next proposition, this construction is used in VI.25 to make a figure similar to
one rectilinear figure but equal to another.
Book I introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 45
It is required to construct a parallelogram equal to the rectilinear figure ABCD in the given
angle E.
Join DB. Construct the parallelogram FH equal to the triangle ABD in the angle HKF which Post.1
equals E. Apply the parallelogram GM equal to the triangle DBC to the straight line GH in the I.42
angle GHM which equals E. I.44
Since the angle E equals each of the angles HKF and GHM, therefore the angle HKF also C.N.1
equals the angle GHM.
Add the angle KHG to each. Therefore the sum of the angles FKH and KHG equals the sum C.N.2
of the angles KHG and GHM.
But the sum of the angles FKH and KHG equals two right angles, therefore the sum of the I.29
angles KHG and GHM also equals two right angles. C.N.1
Thus, with a straight line GH, and at the point H on it, two straight lines KH and HM not lying
on the same side make the adjacent angles together equal to two right angles, therefore KH is I.14
in a straight line with HM.
Since the straight line HG falls upon the parallels KM and FG, therefore the alternate angles I.29
MHG and HGF equal one another.
Add the angle HGL to each. Then the sum of the angles MHG and HGL equals the sum of the C.N.2
angles HGF and HGL.
I.29
But the sum of the angles MHG and HGL equals two right angles, therefore the sum of the C.N.1
angles HGF and HGL also equals two right angles. Therefore FG is in a straight line with GL. I.14
I.34
Since FK is equal and parallel to HG, and HG equal and parallel to ML also, therefore KF is
I.30
also equal and parallel to ML, and the straight lines KM and FL join them at their ends. C.N.1
Therefore KM and FL are also equal and parallel. Therefore KFLM is a parallelogram. I.33
Since the triangle ABD equals the parallelogram FH, and DBC equals GM, therefore the C.N.2
whole rectilinear figure ABCD equals the whole parallelogram KFLM.
Therefore the parallelogram KFLM has been constructed equal to the given rectilinear figure
ABCD in the angle FKM which equals the given angle E.
Q.E.F.
With this construction any rectilinear area can be applied to a line in an angle, that is, it can be
transformed into a parallelogram with whatever angle you want and with one side whatever you want.
That is a satisfactory solution to the question "what's the area of this figure?"
But the question "what's the area of a circle?" is not answered in the Elements. See the note on squaring
the circle after proposition II.14 for more discussion of this question.
Use of Proposition 45
This construction is used in propositions II.14, VI.25, and XI.32. Like many of the other constructions in
Book I, it is used to make constructions in different planes as is done in XI.32.
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 46
Draw AC at right angles to the straight line AB from the point A on I.11
it. Make AD equal to AB. Draw DE through the point D parallel to I.3
AB, and draw BE through the point B parallel to AD. I.31
But AB equals AD, therefore the four straight lines BA, AD, DE, and EB equal one another.
Therefore the parallelogram ADEB is equilateral.
Since the straight line AD falls upon the parallels AB and DE, therefore the sum of the angles I.29
BAD and ADE equals two right angles.
But the angle BAD is right, therefore the angle ADE is also right.
And in parallelogrammic areas the opposite sides and angles equal one another, therefore I.34
each of the opposite angles ABE and BED is also right. Therefore ADEB is right-angled.
Q.E.F.
We now have the second regular polygon, the first being the equilateral triangle of proposition I.1.
Regular polygons with 5, 6, and 15 sides are constructed in Book IV.
Consruction steps
Next, EB is to be drawn through B parallel to AD. In general that construction given in I.31 takes six
circles, but in this case if EB is drawn perpendicular to AB, then it will be parallel to AD, too, and that
construction only takes three circles with radii BA, AG, and GA.
This abbreviation of Euclid's construction requires six circles and four lines. There are alternate
constructions that are a bit shorter. For instance, E may be found as the other intersection of the circles
of radii BA and DA.
Use of Proposition 46
The construction of a square given in this proposition is used in the next proposition, numerous
propositions in Book II, and others in Books VI, XII, and XIII.
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 47
In right-angled triangles the square on the side opposite the right angle equals the sum of 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 equals the sum of the squares on BA and AC.
I.46
Describe the square BDEC on BC, and the squares GB and HC on BA and AC. Draw AL I.31
through A parallel to either BD or CE, and join AD and FC. Post.1
I.
Since the angle DBC equals the angle FBA, for each is right, add the angle ABC to each, Def.22
therefore the whole angle DBA equals the whole angle FBC. Post.4
C.N.2
Since DB equals BC, and FB equals BA, the two sides AB and BD equal the two sides FB I.
and BC respectively, and the angle ABD equals the angle FBC, therefore the base AD equals Def.22
the base FC, and the triangle ABD equals the triangle FBC. I.4
Now the parallelogram BL is double the triangle ABD, for they have the same base BD and
are in the same parallels BD and AL. And the square GB is double the triangle FBC, for they I.41
again have the same base FB and are in the same parallels FB and GC.
Similarly, if AE and BK are joined, the parallelogram CL can also be proved equal to the C.N.2
square HC. Therefore the whole square BDEC equals the sum of the two squares GB and HC.
And the square BDEC is described on BC, and the squares GB and HC on BA and AC.
Therefore the square on BC equals the sum of the squares on BA and AC.
Therefore in right-angled triangles the square on the side opposite the right angle equals the sum of
the squares on the sides containing the right angle..
Q.E.D.
This proposition is generalized in VI.31 to arbitrary similar figures placed on the sides of the triangle
ABC. If the rectilinear figures on the sides of the triangle are similar, then that on the hypotenuse is the
sum of the other two figures.
A bit of history
This proposition, I.47, is often called the "Pythagorean theorem," called so by Proclus and others
centuries after Pythagoras and even centuries after Euclid. The statement of the proposition was very
likely known to the Pythagoreans if not to Pythagoras himself. The Pythagoreans and perhaps
Pythagoras even knew a proof of it. But the knowledge of this relation was far older than Pythagoras.
More than a millennium before Pythagoras, the Old Babylonians (ca. 1900-1600 B.C.E) used this
relation to solve geometric problems involving right triangles. Moreover, the tablet known as Plimpton
322 shows that the Old Babylonians could construct all the so-called Pythagorean triples, those triples of
numbers a, b, and c such that a2 + b2 = c2 which describe triangles with integral sides. (The smallest of
these is 3, 4, 5.) For more on Pythagorean triples, see X.29.Lemma 1.
The rule for computing the hypotenuse of a right triangle was well known in ancient China. It is used in
the Zhou bi suan jing, a work on astronomy and mathematics compiled during the Han period, and in the
later important mathematical work Jiu zhang suan shu [Nine Chapters] to solve right triangles.
The Zhou bi includes a very interesting diagram known as the "hypotenuse diagram." This diagram may
not have been in the original text but added by its primary
commentator Zhao Shuang sometime in the third century C.
E. A particular case of this proposition is illustrated by this
diagram, namely, the 3-4-5 right triangle.
The Zhou bi has recently been translated into English with an excellent commentary. See Astronomy and
mathematics in ancient China: the Zhou bi suan jing, by Christopher Cullin, Cambridge University
Press, 1996.
According to Proclus, the specific proof of this proposition given in the Elements is Euclid's own. It is
likely that older proofs depended on the theories of proportion and similarity, and as such this
proposition would have to wait until after Books V and VI where those theories are developed. It
appears that Euclid devised this proof so that the proposition could be placed in Book I.
Euclid presents a proof based on proportion and similarity in the lemma for proposition X.33. Compare
it, summarized here, to the proof in I.47.
Let ABC be a right-angled triangle with a right angle at A. Draw AM perpendicular to BC.
Next VI.17 converts this proportion to a statement about areas, namely, the rectangle CB by BM
(which is the parallelogram BL in the proof of I.47) equals the square on AB. For the same reason the
rectangle BC by CM (which is the parallelogram CL in the proof of I.47) also equals the square on
AC. Therefore the sum of the two rectangles CB by BM and BC by CM, which is the square on BC,
equals the sum of the squares on AC and BC. Q.E.D.
(Actually, the final sentence is not part of the lemma, probably because Euclid moved that statement to
the first Book as I.47.)
So, although Euclid's proof in I.47 may be more complicated than some others, we can see how it well it
corresponds to a simpler proof that depends on the theories of proportion and similarity.
Generalizations of I.47
Propositions II.12 and II.13 consider triangles other than right triangles. In II.12 the right angle is
replaced by an obtuse angle, while in II.13 the right angle is replaced by an acute angle. The resulting
statements are actually geometric forms of the law of cosines.
Proposition VI.31 generalizes the figures that can be placed on the sides of the right triangle to any three
similar figures instead of the three squares here in I.47.
This proposition is used in the next one, which its converse, in propositions II.9 through II.14 in Book II,
and several propositions in the rest of the books on geometry.
Book I introduction
© 1996
D.E.Joyce
Clark University
Proposition 48
If in a triangle the square on one of the sides equals the sum of the squares on the remaining
two sides of the triangle, then the angle contained by the remaining two sides of the triangle is
right.
In the triangle ABC let the square on one side BC equal the sum of the squares on the sides BA
and AC
I.11
Draw AD from the point A at right angles to the straight line AC. Make I.3
AD equal to BA, and join DC. Post.1
Since DA equals AB, therefore the square on DA also equals the square
on AB.
Add the square on AC to each. Then the sum of the squares on DA and AC equals the sum of C.N.2
the squares on BA and AC.
But the square on DC equals the sum of the squares on DA and AC, for the angle DAC is right, I.47
and the square on BC equals the sum of the squares on BA and AC, for this is the hypothesis,
therefore the square on DC equals the square on BC, so that the side DC also equals BC. C.N.1
Since DA equals AB, and AC is common, the two sides DA and AC equal the two sides BA and
AC, and the base DC equals the base BC, therefore the angle DAC equals the angle BAC. But I.8
the angle DAC is right, therefore the angle BAC is also right.
Therefore if in a triangle the square on one of the sides equals the sum of the squares on the
remaining two sides of the triangle, then the angle contained by the remaining two sides of the
triangle is right.
Q.E.D.
Book I introduction
© 1996
D.E.Joyce
Clark University
Table of contents
● Definitions (2)
● Propositions (14)
● Guide to Book II
● Logical structure of Book II
Definitions
Definition 1.
Any rectangular parallelogram is said to be contained by the two straight lines containing the
right angle.
Definition 2
And in any parallelogrammic area let any one whatever of the parallelograms about its diameter
with the two complements be called a gnomon.
Propositions
Proposition 1.
If there are two straight lines, and one of them is cut into any number of segments whatever, then
the rectangle contained by the two straight lines equals the sum of the rectangles contained by the
uncut straight line and each of the segments.
Proposition 2.
If a straight line is cut at random, then the sum of the rectangles contained by the whole and each
of the segments equals the square on the whole.
Proposition 3.
If a straight line is cut at random, then the rectangle contained by the whole and one of the
segments equals the sum of the rectangle contained by the segments and the square on the
aforesaid segment.
Proposition 4.
If a straight line is cut at random, the square on the whole equals the squares on the segments plus
twice the rectangle contained by the segments.
Proposition 5.
If a straight line is cut into equal and unequal segments, then the rectangle contained by the
unequal segments of the whole together with the square on the straight line between the points of
section equals the square on the half.
Proposition 6.
If a straight line is bisected and a straight line is added to it in a straight line, then the rectangle
contained by the whole with the added straight line and the added straight line together with the
square on the half equals the square on the straight line made up of the half and the added straight
line.
Proposition 7.
If a straight line is cut at random, then the sum of the square on the whole and that on one of the
segments equals twice the rectangle contained by the whole and the said segment plus the square
on the remaining segment.
Proposition 8.
If a straight line is cut at random, then four times the rectangle contained by the whole and one of
the segments plus the square on the remaining segment equals the square described on the whole
and the aforesaid segment as on one straight line.
Proposition 9.
If a straight line is cut into equal and unequal segments, then the sum of the squares on the
unequal segments of the whole is double the sum of the square on the half and the square on the
straight line between the points of section.
Proposition 10.
If a straight line is bisected, and a straight line is added to it in a straight line, then the square on
the whole with the added straight line and the square on the added straight line both together are
double the sum of the square on the half and the square described on the straight line made up of
the half and the added straight line as on one straight line.
Proposition 11.
To cut a given straight line so that the rectangle contained by the whole and one of the segments
equals the square on the remaining segment.
Proposition 12.
In obtuse-angled triangles the square on the side opposite the obtuse angle is greater than the sum
of the squares on the sides containing the obtuse angle by twice the rectangle contained by one of
the sides about the obtuse angle, namely that on which the perpendicular falls, and the straight
line cut off outside by the perpendicular towards the obtuse angle.
Proposition 13.
In acute-angled triangles the square on the side opposite the acute angle is less than the sum of
the squares on the sides containing the acute angle by twice the rectangle contained by one of the
sides about the acute angle, namely that on which the perpendicular falls, and the straight line cut
off within by the perpendicular towards the acutc angle.
Proposition 14.
To construct a square equal to a given rectilinear figure.
Guide to Book II
The subject matter of Book II is usually called "geometric algebra." The first ten propositions of Book II
can be easily interpreted in modern algebraic notation. Of course, in doing so the geometric flavor of the
propositions is lost. Nonetheless, restating them algebraically can aid in understanding them. The
equations are all quadratic equations since the geometry is plane geometry.
II.1. If y = y1 + y2 + ... + yn, then xy = x y1 + x y2 + ... + x yn. This can be stated in a single identity as
II.2. If x = y + z, then x2 = xy + xz. This can be stated in various ways in an identity of two variables. For
instance,
(y + z)2 = (y + z) y + (y + z) z,
or
x2 = xy + x (x – y).
(y + z)y = yz + y2,
and
xy = y(x – y) + y2.
(y + z)2 = y2 + z2 + 2yz.
x2 + z2 = 2xz + (x – z)2.
The remaining four propositions are of a slightly different nature. Proposition II.11 cuts a line into two
parts which solves the equation a (a – x) = x2 geometrically. Propositions II.12 and II.13 are
recognizable as geometric forms of the law of cosines which is a generalization of I.47. The last
propostion II.14 constructs a square equal to a given rectilinear figure thereby completeing the theory of
areas begun in Book I.
The proofs of the propositions in Book II heavily rely on the propositions in Book I involving right
angles and parallel lines, but few others. For instance, the important congruence theorems for triangles,
namely I.4, I.8, and I.26, are not invoked even once. This is understandable considering Book II is
mostly algebra interpreted in the theory of geometry.
The first ten propositions in Book II were written to be logically independent, but they could have easily
been written in logical chains which, perhaps, would have shortened the exposition a little. The
remaining four propositions each depend on one of the first ten.
6 11
4 12
7 13
5 14
Elements
Introduction
© 1996 http://aleph0.clarku.edu/~djoyce/java/elements/elements.html
D.E.Joyce
Dept. Math. & Comp. Sci.
Clark University
Definitions
Def. 1. Any rectangular parallelogram is said to be contained by the two straight lines containing
the right angle.
Def. 2. And in any parallelogrammic area let any one whatever of the parallelograms about its
diameter with the two complements be called a gnomon.
Guide
According to the first definition, the rectangle ABCD illustrated on the left is contained by the lines AB
and BC, and this rectangle can be called the rectangle AB by BC. Of course, it could also be called the
rectangle BC by CD, or two other names.
On the right, in the parallelogram EFGH, there is a diameter EG with a parallelogram LNGO about it
and the two complements KLOF and MHNL, and these three parallelograms together make up the
gnomon. In other words a gnomon is an L-shaped figure made by removing a parallelogram from a
larger similar parallelogram. (The "g" in "gnomon" is silent.)
Euclid illustrated gnomons by arcs of circles around the inner vertex. In this example, the gnomon is
called the gnomon PQR.
© 1996 http://aleph0.clarku.edu/~djoyce/java/elements/elements.html
D.E.Joyce
Dept. Math. & Comp. Sci.
Clark University
Definitions
Def. 1. Any rectangular parallelogram is said to be contained by the two straight lines containing
the right angle.
Def. 2. And in any parallelogrammic area let any one whatever of the parallelograms about its
diameter with the two complements be called a gnomon.
Guide
According to the first definition, the rectangle ABCD illustrated on the left is contained by the lines AB
and BC, and this rectangle can be called the rectangle AB by BC. Of course, it could also be called the
rectangle BC by CD, or two other names.
On the right, in the parallelogram EFGH, there is a diameter EG with a parallelogram LNGO about it
and the two complements KLOF and MHNL, and these three parallelograms together make up the
gnomon. In other words a gnomon is an L-shaped figure made by removing a parallelogram from a
larger similar parallelogram. (The "g" in "gnomon" is silent.)
Euclid illustrated gnomons by arcs of circles around the inner vertex. In this example, the gnomon is
called the gnomon PQR.
© 1996 http://aleph0.clarku.edu/~djoyce/java/elements/elements.html
D.E.Joyce
Dept. Math. & Comp. Sci.
Clark University
Proposition 1
If there are two straight lines, and one of them is cut into any number of segments whatever,
then the rectangle contained by the two straight lines equals the sum of the rectangles
contained by the uncut straight line and each of the segments.
Let A and BC be two straight lines, and let BC be cut at random at the points D
and E.
I say that the rectangle A by BC equals the sum of the rectangle A by BD, the
rectangle A by DE, and the rectangle A by EC.
Therefore the rectangle A by BC equals the sum of the rectangle A by BD, the
rectangle A by DE, and the rectangle A by EC.
Therefore if there are two straight lines, and one of them is cut into any number of segments
whatever, then the rectangle contained by the two straight lines equals the sum of the rectangles
contained by the uncut straight line and each of the segments.
Q.E.D.
The phrase "the rectangle contained by the two straight lines" means any rectangle constructed with two
sides equal to the two given sides. In some sense this is the product of the two lines. When the sides
have names, such as A and BC, we will refer to that rectangle by "the rectangle A by BC" since that is a
little clearer than Euclid's terse "the rectangle A, BC."
BC = BD + DE + EC,
then
(A by BC) = (A by BD) + (A by DE) + (A by EC).
In modern algebraic notation this could be stated as follows: If y = y1 + y2 + ... + yn, then
xy = x y1 + x y2 + ... + x yn. This can also be stated in a single equation as
Here x and the various yi's are all lines, and n is an arbitrary number. In modern terminology this identity
is called the distributive law for multiplication over addition.
This proposition is not specifically invoked in the rest of the Elements. The next two propositions,
however, are special cases of it, and they are each explicitly used once.
Select topic
Book II introduction
© 1996 http://aleph0.clarku.edu/~djoyce/java/elements/elements.html
D.E.Joyce
Dept. Math. & Comp. Sci.
Clark University
Proposition 2
If a straight line is cut at random, then the sum of the rectangles contained by the whole and
each of the segments equals the square on the whole.
Now AE is the square on AB; AF is the rectangle BA by AC, for it is contained by DA and II.Def.1
AC, and AD equals AB; and CE is the rectangle AB by BC, for BE equals AB.
Therefore the sum of the rectangle BA by AC and the rectangle AB by BC equals the square
on AB.
Therefore if a straight line is cut at random, then the sum of the rectangles contained by the whole
and each of the segments equals the square on the whole.
Q.E.D.
This proposition is actually a special case of II.1. In II.1 Euclid shows that the product of one line by a
sum of any number of lines is the sum of the products of that line by each of the lines. In this
proposition, there are just two of those lines and their sum equals the one line. Rather than using II.1 to
prove II.2, Euclid proves II.2 directly. This suggests that II.1 may have been inserted into the Elements
after II.2 was included, either by Euclid or someone else.
In modern algebraic notation this proposition says that if y = y1 + y2, then xy = x y1 + x y2. This can also
be stated in a single equation as
This proposition is used in the proof of proposition XIII.10 which shows that a certain relationship holds
for the sides of a regular pentagon, regular hexagon, and regular decagon that are all inscribed in the
same circle, namely, the square on the side of the pentagon equals the sum of the squares on the side of a
hexagon and on the side of a decagon.
Select topic
Book II introduction
© 1996 http://aleph0.clarku.edu/~djoyce/java/elements/elements.html
D.E.Joyce
Dept. Math. & Comp. Sci.
Clark University
Proposition 3
If a straight line is cut at random, then the rectangle contained by the whole and one of the
segments equals the sum of the rectangle contained by the segments and the square on the
aforesaid segment.
Let the straight line AB be cut at random at C.
Now AE is the rectangle AB by BC, for it is contained by AB and BE, and BE equals BC; AD is
the rectangle AC by CB, for DC equals CB; and DB is the square on CB.
Therefore the rectangle AB by BC equals the sum of the rectangle AC by CB and the square on
BC.
Therefore if a straight line is cut at random, then the rectangle contained by the whole and one of the
segments equals the sum of the rectangle contained by the segments and the square on the aforesaid
segment.
Q.E.D.
This proposition is another special case of II.1. In modern algebraic notation it says that if x = y + z, then
xy = y2 + yz. Identities that are logically equivalent to this implication can be found by eliminating one
of the three variables x, y, or z. Here are two of them.
(y + z) y = y2 + yz,
and
xy = y2 + y (x – y).
This proposition refers to lines and rectangles, but the analogous statement for numbers is used in a
proposition in one of the Euclid's books on number theory, namely, of proposition IX.15.
Select topic
Book II introduction
© 1996 http://aleph0.clarku.edu/~djoyce/java/elements/elements.html
D.E.Joyce
Dept. Math. & Comp. Sci.
Clark University
Proposition 4
If a straight line is cut at random, then the square on the whole equals the sum of the squares
on the segments plus twice the rectangle contained by the segments.
Let the straight line AB be cut at random at C.
I say that the square on AB equals the sum of the squares on AC and CB plus twice the rectangle
AC by CB.
But the angle ADB equals the angle ABD, since the side BA
I.5
also equals AD. Therefore the angle CGB also equals the I.6
angle GBC, so that the side BC also equals the side CG.
But CB equals GK, and CG to KB. Therefore GK also equals KB. Therefore CGKB is equilateral. I.34
Since CG is parallel to BK, the sum of the angles KBC and GCB equals two right angles. I.29
But the angle KBC is right. Therefore the angle BCG is also right, so that the opposite angles I.34
CGK and GKB are also right.
Therefore CGKB is right-angled, and it was also proved equilateral, therefore it is a square, and
it is described on CB.
For the same reason HF is also a square, and it is described on HG, that is AC. Therefore the I.34
squares HF and KC are the squares on AC and CB.
Now, since AG equals GE, and AG is the rectangle AC by CB, for GC equals CB, therefore GE
also equals the rectangle AC by CB. Therefore the sum of AG and GE equals twice the rectangle I.43
AC by CB.
But the squares HF and CK are also the squares on AC and CB, therefore the sum of the four
figures HF, CK, AG, and GE equals the sum of the squares on AC and CB plus twice the
rectangle AC by CB.
But HF, CK, AG, and GE are the whole ADEB, which is the square on AB.
Therefore the square on AB equals the the sum of the squares on AC and CB plus twice the
rectangle AC by CB.
Therefore if a straight line is cut at random, the square on the whole equals the squares on the
segments plus twice the rectangle contained by the segments.
Q.E.D.
The statement of the proposition can be interpreted in modern notation as saying that if x = y + z, then
x2 = y2 + z2 + 2yz. More simply, as an identity, it says
(y + z)2 = y2 + z2 + 2yz.
This is one of the more frequently used propositions of Book II. It is used in II.12 later in this book,
frequently in BookX, and in XIII.2. Also, the analogous statement for numbers is used in IX15,
Select topic
Book II introduction
© 1996 http://aleph0.clarku.edu/~djoyce/java/elements/elements.html
D.E.Joyce
Dept. Math. & Comp. Sci.
Clark University
Proposition 5
If a straight line is cut into equal and unequal segments, then the rectangle contained by the
unequal segments of the whole together with the square on the straight line between the points
of section equals the square on the half.
Let a straight line AB be cut into equal segments at C and into unequal segments at D.
I say that the rectangle AD by DB together with the square on CD equals the square on CB.
Then, since the complement CH equals the complement HF, add DM to each. Therefore the I.43
whole CM equals the whole DF.
But CM equals AL, since AC is also equal to CB. Therefore AL also equals DF. Add CH to I.36
each. Therefore the whole AH equals the gnomon NOP. II.Def.2
But AH is the rectangle AD by DB, for DH equals DB, therefore the gnomon NOP also
equals the rectangle AD by DB.
Add LG, which equals the square on CD, to each. Therefore the sum of the gnomon NOP and
LG equals the sum of the rectangle AD by DB and the square on CD.
But the gnomon NOP together with LG is the whole square CEFB, which is described on CB.
Therefore the rectangle AD by DB together with the square on CD equals the square on CB.
Therefore if a straight line is cut into equal and unequal segments, then the rectangle contained by the
unequal segments of the whole together with the square on the straight line between the points of
section equals the square on the half.
Q.E.D.
In the figure there is a part of a circle denoted with the points NOP. This is supposed to indicate the
gnomon which is three parts of the square BCEF, the only part left out for the gnomon being the square
EGHL. Later versions of the Elements did not have this curve in the figure. Instead, they named the
same gnomon as LBG. Either way is sufficient to specify the gnomon.
We can represent a rectangle algebraically as xy where the sides are x and y. In the diagram above, take x
as the line AD and y as the line DH, so that xy is the rectangle AH. According to this proposition, this
product, or rectangle, is the difference of two squares, the large one being the square of (x + y)/2, which
is the square on the line BC in the diagram, and the small one being the square of (x – y)/2, which is the
square on the line LH (which equals the square on the line CD).
x+y y–x
2
xy = ( ) – ( )2
2 2
But Euclid was resticted to geometric arguments. The argument isn't difficult. The original rectangle AH
is the sum of the rectangles AL and CH. By proposition I.43, the rectangles CH and HF are equal. And,
of course, the rectangles AL and CM are equal. Therefore, AH = AL + CH = CM + HF = CB2 – LH2, as
required.
This proposition is set up to help in the solution of a quadratic problem of the following form.
Find two numbers xand yso that their sum is a known value band their product is a known value
c2.
In terms of the single variable x, this is equivalent to solving the quadratic equation, x(b – x) = c2. This
equation can be written in a standard form as
x2 + c2 = bx.
If b is represented as the line AB in the diagram, and with x = AD and y = BD, the first condition
x + y = b is satisfied. This proposition says that the product xy equals the square on BC (which is b/2)
minus the square on CD. Thus, the remaining condition reduces to finding CD so that (b/2)2 – CD2 = c2.
By I.47, if a right triangle is constructed with one side equal to b/2 and another equal to c, then the
hypotenuse will equal the required value for CD. Algebraically, the solutions AD for x and BD for y
have the values
This analysis yields a construction to solve the quadratic problem stated above.
To cut a given straight line so that the rectangle contained by the unequal segments equals a given
square. Thus the given square must not be greater than the square described on the half of the given
straight line.
Then, as described above, AB has been cut at G so that AG times GB equals the given square.
This proposition is not found in the Elements, but a generalization is. After II.14 the given square could
be replaced by any given rectilinear figure, since II.14 constructs a square equal to a given rectilinear
figure. But the full generalization is not given until proposition VI.28. Not only has the given square
become a general rectilinear figure, but all the rectangles and squares have been replaced by
parallelograms. That requires generalizing II.4 to parallelograms. That's done in VI.25 which constructs
a parallelgram similar to a given parallelogram and equal to a given rectilinear figure. It also requires a
few technical propositions to carry out the proof.
Select topic
Book II introduction
Proposition 6
If a straight line is bisected and a straight line is added to it in a straight line, then the
rectangle contained by the whole with the added straight line and the added straight line
together with the square on the half equals the square on the straight line made up of the half
and the added straight line.
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 AD by DB together with the square on CB equals the square
on CD.
Then, since AC equals CB, AL also equals CH. But CH equals HF. Therefore AL I.36
also equals HF. I.43
Add CM to each. Therefore the whole AM equals the gnomon NOP. II.Def.2
But AM is the rectangle AD by DB, for DM equals DB. Therefore the gnomon NOP
also equals the rectangle AD by DB.
Add LG, which equals the square on BC, to each. Therefore the rectangle AD by DB
together with the square on CB equals the gnomon NOP plus LG.
But the gnomon NOP and LG are the whole square CEFD, which is described on
CD.
Therefore the rectangle AD by DB together with the square on CB equals the square
on CD.
Therefore if a straight line is bisected and a straight line is added to it in a straight line, then the
rectangle contained by the whole with the added straight line and the added straight line together
with the square on the half equals the square on the straight line made up of the half and the added
straight line.
Q.E.D.
This proposition is remarkably similar to the last one, II.5, except the point D does not lie on the line AB
but on that line extended.
Let b denote the line AB, x denote AD, and y denote BD as in II.5. Then x – y = b (as opposed to
x + y = b as in II.5). According to this proposition the rectangle AD by DB, which is the product xy, is
the difference of two squares, the large one being the square on the line CD, that is the square of x – b/2,
and the small one being the square on the line CB, that is, the square of b/2. Algebraically,
This equation is easily verified with modern algebra, but it's also easily verified in geometry, as done
here in the proof.
The geometric proof is primarily an exercise in cutting and pasting. The rectangle AB by DB is the
rectangle AM, which is the sum of the rectangles AL and CM. But the rectangles AL, CH, and HF are all
equal. Therefore, the rectangle AB by DB equals the gnomon formed by the rectangles CM and HF. That
gnomon is the square CF minus the square LG, but the latter equals the square on BC. Thus, the
rectangle AB by DB equals the square on DB minus the ssquare on CB.
As was II.5, this proposition is set up to help in the solution of a quadratic problem:
Find two numbers xand yso that their difference x – yis a known value band their product is a
In terms of x alone, this is equivalent to solving the quadratic equation x(x – b) = c2. Since this
proposition says that x(x – b) = (x – b/2)2 – (b/2)2, the problem reduces to solving the equation
c2 = (x – b/2)2 – (b/2)2,
that is, finding CD so that CD2 = (b/2)2 + c2. By I.47, if a right triangle is constructed with one side
equal to b/2 and another equal to c, then the hypotenuse will equal the required value for CD.
Algebraically, the solutions AD for x and BD for y have the values
This analysis yields a construction to solve the quadratic problem stated above.
To apply a rectangle equal to a given square to a given straight line but exceeding it by a square.
Then, as described above, AB has been extended to D so that AD times BD equals the given square.
This construction is not found in the Elements, but a generalization of it to parallelograms is proposition
VI.29.
Select topic
Book II introduction
Proposition 7
If a straight line is cut at random, then the sum of the square on the whole and that on one of
the segments equals twice the rectangle contained by the whole and the said segment plus the
square on the remaining segment.
Let a straight line AB be cut at random at the point C.
I say that the sum of the squares on AB and BC equals twice the rectangle AB by BC plus the
square on CA.
Describe the square ADEB on AB, and let the figure be I.46
drawn. I.31
But the sum of AF and CE equals the gnomon KLM plus the square CF, therefore the gnomon
KLM plus the square CF is double AF.
But twice the rectangle AB by BC is also double AF, for BF equals BC, therefore the gnomon
KLM plus the square CF equal twice the rectangle AB by BC.
Add DG, which is the square on AC, to each. Therefore the gnomon KLM plus the sum of the
squares BG and GD equals twice the rectangle AB by BC plus the square on AC.
But the gnomon KLM plus the sum of the squares BG and GD equals the whole ADEB plus CF,
which are squares described on AB and BC.
Therefore the sum of the squares on AB and BC equals twice the rectangle AB by BC plus the
square on CA.
Therefore if a straight line is cut at random, then the sum of the square on the whole and that on one
of the segments equals twice the rectangle contained by the whole and the said segment plus the
square on the remaining segment.
Q.E.D.
We can interpret this algebraically with x for AB, y for AC, and z for CB. Then the proposition says that
if x = y + z, then x2 + z2 = 2xz + y2. This can be rewritten as various identities depending on which
variable is eliminated, the simplest being
x2 + z2 = 2xz + (x – z)2.
This proposition is used later in Book II to prove proposition II.13, and it is used repeatedly in Book X.
Select topic
Book II introduction
© 1996 http://aleph0.clarku.edu/~djoyce/java/elements/elements.html
D.E.Joyce
Dept. Math. & Comp. Sci.
Clark University
Proposition 8
If a straight line is cut at random, then four times the rectangle contained by the whole and one
of the segments plus the square on the remaining segment equals the square described on the
whole and the aforesaid segment as on one straight line.
Let a straight line AB be cut at random at the point C.
I say that four times the rectangle AB by BC plus the square on AC equals the square described
on AB and BC as on one straight line.
And, since BC equals BD, and GK equals KN, therefore CK also equals KD, and GR equals RN. I.36
But CK equals RN, for they are complements of the parallelogram CP. Therefore KD also
equals GR. Therefore the four areas DK, CK, GR, RN equal one another. Therefore the four are I.43
quadruple of CK.
Again, since CB equals BD, while BD equals BK, that is CG, and CB equals GK, that is GQ, I.34
therefore CG also equals GQ.
And, since CG equals GQ, and QR equals RP, AG also equals MQ, and QL equals RF. I.36
But MQ equals QL, for they are complements of the parallelogram ML, therefore AG also
equals RF. Therefore the four areas AG, MQ, QL, RF equal one another. Therefore the four are I.43
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 by BD, for BK equals BD, therefore four times the rectangle
AB by BD is quadruple of AK.
But the gnomon STU was also proved to be quadruple of AK, therefore four times the rectangle
AB by BD equals the gnomon STU.
Add OH, which equals the square on AC, to each. Therefore four times the rectangle AB by BD
plus the square on AC equals the gnomon STU plus OH.
But the gnomon STU and OH are the whole square AEFD, which is described on AD. Therefore
four times the rectangle AB by BD plus the square on AC equals the square on AD.
Therefore four times the rectangle AB by BC together with the square on AC equals the square
on AD, that is to the square described on AB and BC as on one straight line.
Therefore if a straight line is cut at random, then four times the rectangle contained by the whole and
one of the segments plus the square on the remaining segment equals the square described on the
whole and the aforesaid segment as on one straight line.
Q.E.D.
Select topic
Book II introduction
© 1996 http://aleph0.clarku.edu/~djoyce/java/elements/elements.html
D.E.Joyce
Dept. Math. & Comp. Sci.
Clark University
Proposition 9
If a straight line is cut into equal and unequal segments, then the sum of the squares on the
unequal segments of the whole is double the sum of the square on the half and the square on
the straight line between the points of section.
Let a straight line AB be cut into equal segments at C, and into unequal segments at D.
I say that the sum of the squares on AD and DB is double the sum of the squares on AC and CD.
And, since the angle at C is right, the sum of the remaining angles EAC and AEC equals one I.32
right angle.
And they are equal, therefore each of the angles CEA and CAE is half of a right angle.
For the same reason each of the angles CEB and EBC is also half of a right angle, therefore the
whole angle AEB is right.
And, since the angle GEF is half of a right angle, and the angle EGF is right, for it equals the I.29
interior and opposite angle ECB, the remaining angle EFG is half of a right angle. Therefore the I.32
angle GEF equals the angle EFG, so that the side EG also equals GF. I.6
Again, since the angle at B is half of a right angle, and the angle FDB is right, for it is again I.29
equal to the interior and opposite angle ECB, the remaining angle BFD is half of a right angle. I.32
Therefore the angle at B equals the angle DFB, so that the side FD also equals the side DB. I.6
Now, since AC equals CE, the square on AC also equals the square on CE, therefore the sum of
the squares on AC and CE is double the square on AC.
But the square on EA equals the sum of the squares on AC and CE, for the angle ACE is right, I.47
therefore the square on EA is double the square on AC.
Again, since EG equals GF, the square on EG also equals the square on GF. Therefore the sum
of the squares on EG and GF is double the square on GF.
I.47
But the square on EF equals the sum of the squares on EG and GF, therefore the square on EF
is double the square on GF.
But GF equals CD, therefore the square on EF is double the square on CD. I.34
But the square on EA is also double of the square on AC, therefore the sum of the squares on AE
and EF is double the sum of the squares on AC and CD.
And the square on AF equals sum of the squares on AE and EF, for the angle AEF is right. I.47
Therefore the square on AF is double the sum of the squares on AC and CD.
But the sum of the squares on AD and DF equals the square on AF, for the angle at D is right, I.47
therefore the sum of the squares on AD and DF is double the sum the squares on AC and CD.
Therefore the sum of the squares on AD and DB is double the sum of the squares on AC and CD.
Therefore if a straight line is cut into equal and unequal segments, then the sum of the squares on the
unequal segments of the whole is double the sum of the square on the half and the square on the
straight line between the points of section.
Q.E.D.
Start with the given line AB bisected at C and cut at another point D. The construction will result in a
line AF whose square is simultaneously equal to both AD2 + DB2 and 2(AC2+CD2).
Draw CE at right angles to AB and equal to half of it. Finish the diagram by drawing parallel lines and
connecting points. That results in four isosceles right triangles, ACE, ECB, EGF, and FDB, as well as
two other right triangles, AEB and AEF. Then,
Up until this proposition, Euclid has only used cut-and-paste proofs, and such a proof can easily be made
for this proposition as well. It would start with the same line AB bisected at C and also cut at D. Then
lines at right angles and parallel to line AB would be constructed to make squares and rectangles of
various sizes.
The equation, AD2 + DB2 = 2(AC2+CD2), can be interpreted in various ways depending on which parts
of the given information are taken as basic. These interpretations are more of an aid to the modern reader
than as intrinsic aspects of the proposition, since they are interpretations in modern symbolic algebra.
For instance, when AC and CB are set to y while CD is set to z, then the algebraic identity
results.
Alternatively, when AC and CB are set to y while BC is set to x,, then we get the identity
There are yet other interpretiations. This proposition is used in Book X to prove a lemma for X.60, and
in that lemma, w = AD and x = DB are given first, so that
Select topic
Book II introduction
Proposition 10
If a straight line is bisected, and a straight line is added to it in a straight line, then the square
on the whole with the added straight line and the square on the added straight line both
together are double the sum of the square on the half and the square described on the straight
line made up of the half and the added straight line as on one straight line.
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 sum of the squares on AD and DB is double the sum of the squares on AC and
CD.
But straight lines produced from angles less than two right angles meet. Therefore EB and FD, Post.5
if produced in the direction B and D, will meet.
Then, since AC equals CE, the angle EAC also equals the angle AEC. The angle at C is right, I.5
therefore each of the angles EAC and AEC is half of a right angle. I.32
For the same reason each of the angles CEB and EBC is also half of a right angle, therefore the
angle AEB is right.
And, since the angle EBC is half of a right angle, the angle DBG is also half of a right angle. I.15
But the angle BDG is also right, for it equals the angle DCE, since they are alternate. I.29
Therefore the remaining angle DGB is half of a right angle. Therefore the angle DGB equals I.32
the angle DBG, so that the side BD also equals the side GD. I.6
Again, since the angle EGF is half of a right angle, and the angle at F is right, for it equals the I.34
opposite angle, the angle at C, the remaining angle FEG is half of a right angle. Therefore the I.32
angle EGF equals the angle FEG, so that the side GF also equals the side EF. I.6
Now, since the square on EC equals the square on CA, the sum of the squares on EC and CA is
double the square on CA. But the square on EA equals the sum of the squares on EC and CA, I.47
therefore the square on EA is double the square on AC.
Again, since FG equals EF, the square on FG also equals the square on FE. Therefore the sum
of the squares on GF and FE is double the square on EF. But the square on EG equals the sum I.47
of the squares on GF and FE, therefore the square on EG is double the square on EF.
And EF equals CD, therefore the square on EG is double the square on CD. But the square on
EA was also proved to be double the square on AC, therefore the sum of the squares on AE and I.34
EG is double the sum of the squares on AC and CD.
And the square on AG equals the sum of the squares on AE and EG, therefore the square on
AG is double the sum of the squares on AC and CD. But the sum of the squares on AD and DG I.47
equals the square on AG, therefore the sum of the squares on AD and DG is double the sum of
the squares on AC and CD.
And DG equals DB, therefore the sum of the squares on AD and DB is double the sum of the
squares on AC and CD.
Therefore if a straight line is bisected, and a straight line is added to it in a straight line, then the
square on the whole with the added straight line and the square on the added straight line both
together are double the sum of the square on the half and the square described on the straight line
made up of the half and the added straight line as on one straight line.
Q.E.D.
The proof is almost the same as the previous proposition. The entire situation is the same, except the
point D is moved past the end B of the line AB.
As before draw CE at right angles to AB and equal to half of it. Finish the diagram. That results in
isosceles right triangles, ACE, ECB, and EFG, as well as two other right triangles, AEG and ADG. Then,
This proposition can be interpreted to derive the same algebraic identities as the previous proposition.
For example, when y is identified with CD and z is identified with CB, then the proposition states that
The diagonal of a square is incommensurable with the side of a square, as noted elsewhere in the
Elements. Another way of saying that is that the ratio of the diagonal to a side, what is generally known
as the square root of 2, is not the ratio of two whole numbers, in other words 2 is an irrational number.
Nonetheless, finding a close rational approximation for 2 is an important goal. For practical reasons,
close aprroximations are useful, and for theoretical reasons, close approximations are interesting.
Some time before Euclid, Plato refers in his Republic the "rational diameter" of 5. A square of side
length s = 5 has a diameter (diagonal) of d = 50, which is irrational. But 49 is close to 50, so 7 is close
to this irrational length. There are a number of side lengths s whose squares s2 differ from a square
number d2 by only 1. The first few are displayed in the table below. The column labelled "angle" gives
the angle opposite the diameter d. By the law of cosines, the cosine of the angle equals 1 – d2/(2s2).
Note how the squares of the diagonals alternately are less and greater by 1 than twice the squares of the
sides, and the angle correspondingly alternates between acute and obtuse, but quickly approaches a right
angle. Likewise, the ratio d/s is alternately less and greater then the irrational number 2.
The pattern of the lengths of the sides and diagonals is evident. The next side s' is the sum of the current
side and diagonal, while the next diagonal d' is the sum of twice the current side and diagonal:
s' = s + d
d' = 2s + d.
A pattern requires a verification, and this proposition supplies that. What needs to be verified is that if
2s2 differs from d2 by exactly 1, then so does 2s'2 differ from d'2 by exactly 1.
Let CD equal s, and let BC equal d. Then s' = AC = CB, and d' = AD. By this proposition
Thus, if 2s2 differs from d2 by 1, then 2s'2 also differs from d'2 by 1. More precisely, if one difference is
+1, then the other difference is –1.
Select topic
Book II introduction
Proposition 11
To cut a given straight line so that the rectangle contained by the whole and one of the
segments equals the square on the remaining segment.
Let AB be the given straight line.
It is required to cut AB so that the rectangle contained by the whole and one of the segments
equals the square on the remaining segment.
But EF equals EB, therefore the rectangle CF by FA together with the square on AE equals the
square on EB.
But the sum of the squares on BA and AE equals the square on EB, for the angle at A is right,
therefore the rectangle CF by FA together with the square on AE equals the sum of the squares I.47
on BA and AE.
Subtract the square on AE from each. Therefore the remaining rectangle CF by FA equals the
square on AB.
Now the rectangle CF by FA is FK, for AF equals FG, and the square on AB is AD, therefore
FK equals AD.
And HD is the rectangle AB by BH, for AB equals BD, and FH is the square on AH, therefore
the rectangle AB by BH equals the square on HA.
Therefore the given straight line AB has been cut at H so that the rectangle AB by BH equals the
square on HA.
Q.E.F.
This construction cuts a line into two parts to solve the equation a (a – x) = x2 geometrically.
This construction is used in the proof of IV.10, which is later used to construct a regular pentagon. It
accomplishes the same thing as the construction of proposition VI.30, which cuts a line into extreme and
mean ratio, defined in VI.Def.3, and that construction is used later in XIII.17.
The difference between this proposition and VI.30 is a matter of terminology. Propositions dealing with
ratios of lines are postponed until Book VI, but any ratio concerning lines can be converted into a
statement about areas of rectangles. Proposition VI.16 states that the line A is to the line B as the line C
is to the line D is equivalent to the statement that the rectangle A by D equals the rectangle B by C.
The construction of this proposition cuts a line into two parts A and B so that the rectangle A + B by A
equals the square B by B. The construction in VI.30 cuts a line so that A + B : B = B : A, which by VI.16,
or by its special case VI.17, is the same thing.
Construction steps
This is the first of several propositions in the Elements that treats these concepts. At this point, ratios
have not been introduced, so Euclid describes it in basic terms, that a given straight line is cut so that
"the rectangle contained by the whole and one of the segments equals the square on the remaining
segment."
Select topic
Book II introduction
© 1996 http://aleph0.clarku.edu/~djoyce/java/elements/elements.html
D.E.Joyce
Dept. Math. & Comp. Sci.
Clark University
Proposition 12
In obtuse-angled triangles the square on the side opposite the obtuse angle is greater than the
sum of the squares on the sides containing the obtuse angle by twice the rectangle contained by
one of the sides about the obtuse angle, namely that on which the perpendicular falls, and the
straight line cut off outside by the perpendicular towards the obtuse angle.
Let ABC be an obtuse-angled triangle having the angle BAC obtuse, and draw BD from the point I.12
B perpendicular to CA produced.
I say that the square on BC is greater than the sum of the squares on BA and AC by twice the
rectangle CA by AD.
Since the straight line CD has been cut at random at the point
A, the square on DC equals the sum of the squares on CA and II.4
AD and twice the rectangle CA by AD.
But the square on CB equals the sum of the squares on CD and DB, for the angle at D is right,
and the square on AB equals the sum of the squares on AD and DB, therefore the square on CB
equals the sum of the squares on CA and AB plus twice the rectangle CA by AD, so that the I.47
square on CB is greater than the sum of the squares on CA and AB by twice the rectangle CA by
AD.
Therefore in obtuse-angled triangles the square on the side opposite the obtuse angle is greater than
the sum of the squares on the sides containing the obtuse angle by twice the rectangle contained by
one of the sides about the obtuse angle, namely that on which the perpendicular falls, and the straight
line cut off outside by the perpendicular towards the obtuse angle.
Q.E.D.
This proposition for obtuse angles, together with the next one for acute triangles, complement the
Pythagorean theorem, Proposition I.47, for right triangles. Prop.I.47 says that if triangle ABC has a right
angle at A, then
a2 = b2 + c2
where a, b, and c are the sides opposite the angles A, B, and C, respectively.
a2 = b2 + c2 - 2ch
This conclusion is very close to the law of cosines for oblique triangles.
a2 = b2 + c2 – 2bc cos A,
since the height h equals b cos A. Trigonometry was developed some time after the Elements was
written, and the negative numbers needed here (for the cosine of an obtuse angle) were not accepted
until long after most of trigonometry was developed. Nonetheless, this proposition and the next may be
considered geometric versions of the law of cosines.
Neither this nor the next is used in the rest of the Elements.
Select topic
Book II introduction
© 1996 http://aleph0.clarku.edu/~djoyce/java/elements/elements.html
D.E.Joyce
Dept. Math. & Comp. Sci.
Clark University
Proposition 13
In acute-angled triangles the square on the side opposite the acute angle is less than the sum of
the squares on the sides containing the acute angle by twice the rectangle contained by one of
the sides about the acute angle, namely that on which the perpendicular falls, and the straight
line cut off within by the perpendicular towards the acute angle.
Let ABC be a triangle having the angle at B acute, and draw AD from the point A perpendicular I.12
to BC.
I say that the square on AC is less than the sum of the squares on CB and BA by twice the
rectangle CB by BD.
Since the straight line CB has been cut at random at D, the sum
of the squares on CB and BD equals twice the rectangle CB by II.7
BD plus the square on DC.
But the square on AB equals the sum of the squares on BD and DA, for the angle at D is right,
and the square on AC equals the sum of the squares on AD and DC, therefore the sum of the
squares on CB and BA equals the square on AC plus twice the rectangle CB by BD, so that the I.47
square on AC alone is less than the sum of the squares on CB and BA by twice the rectangle CB
by BD.
Therefore in acute-angled triangles the square on the side opposite the acute angle is less than the
sum of the squares on the sides containing the acute angle by twice the rectangle contained by one of
the sides about the acute angle, namely that on which the perpendicular falls, and the straight line cut
off within by the perpendicular towards the acute angle.
Q.E.D.
Select topic
Book II introduction
© 1996 http://aleph0.clarku.edu/~djoyce/java/elements/elements.html
D.E.Joyce
Dept. Math. & Comp. Sci.
Clark University
Proposition 14
Then, if BE equals ED, then that which was proposed is done, for a square BD has been
constructed equal to the rectilinear figure A.
Describe the semicircle BHF with center G and radius one of the straight lines GB or GF. I.Def.18
Produce DE to H, and join GH.
Then, since the straight line BF has been cut into equal segments at G and into unequal
segments at E, the rectangle BE by EF together with the square on EG equals the square on II.5
GF.
But GF equals GH, therefore the rectangle BE by EF together with the square on GE equals
the square on GH.
But the sum of the squares on HE and EG equals the square on GH, therefore the rectangle I.47
BE by EF together with the square on GE equals the sum of the squares on HE and EG.
Subtract the square on GE from each. Therefore the remaining rectangle BE by EF equals
the square on EH.
But the rectangle BE by EF is BD, for EF equals ED, therefore the parallelogram BD equals
the square on HE.
Therefore the rectilinear figure A also equals the square which can be described on EH.
Therefore a square, namely that which can be described on EH, has been constructed equal
to the given rectilinear figure A.
Q.E.F.
The construction of a square equal to a given is short as described in the proof. The verification that this
construction works is also short with the help of Proposition II.5 and Proposition I.47, the Pythagorean
theorem. First, Prop. II.5 allows us to convert the rectangle, BE by ED, into the difference of two
squares, GF2 – GE2. Note that GF equals GH, the hypotenuse of a right triangle GHE. Using I.47 we
can replace the difference of two squares, GH2 – GE2, by the single square, EH2. Thus, the original
rectangle equals the squre EH2.
This proposition finishes the for quadrature of rectilinear figures. The narrow meaning of the word
"quadrature" is to find a square with the same area of a given figure, also called "squaring" the figure. In
a broader sense, "quadrature" means finding the area of a given figure.
Proposition I.45 on application of areas of rectilinear figures allows us to replace the figure under
question with a rectangle of the same area. Now, the semicircle construction in this proposition finds
what is called the "mean proportional" between the sides of the rectangle. If the sides of the rectangle are
denoted a and b, then the mean proportional x between them satisfies the proportion a:x = x:b, and that's
equivalent to an equality of areas ab = x2, that is to say, the square on this mean proportional has the
same area as the rectangle. Thus, any rectilinear figure can be squared.
This result is an end in itself. It is not used in the rest of the Elements.
There is another proof of this proposition that is based on similar triangles. Referring to the figure in the
proposition, draw lines BH and BF, and you'll see three similar right triangles: BFH, BHE, and HGE.
From their similarity it follows that BE:EH = EH:EF. That says EH is the required mean proportional.
Proportions aren't developed until Book V, and similar triangles aren't mentioned until Book VI. So in
order to complete the theory of quadrature of rectilinear figures early in the Elements, Euclid chose a
different proof that doesn't depend on similar triangles. Note that this same result appears in the garb of
proportions in Proposition VI.13. Also in Book VI, Proposition VI.17 shows that the square on the mean
proportional equals the rectangle on the two straight lines.
What about circles and other shapes? The general theory of circles is treated in Book III, but there are no
propositions about the areas of circles until book XII. Proposition XII.2 says the areas of circles are
proportional to the squares on their diameters. That allows the area of two circles to be compared, but it
doesn't answer the question "what's the area of this circle?" in the same way that this proposition does
for rectilinear figures. That would require finding a square equal to a given circle.
This problem of "quadrature of the circle" was one of three famous problems that goes back at least to
the time of Anaxagoras, about 150 years before Euclid. It is equivalent to constructing a line segment of
length pi (relative to a unit length). This problem was solved by ancient Greek geometers but not by
means of the Euclidean tools of straightedge and compass; higher curves were required. In fact, by the
time of Pappus it was believed that the circle could not be squared using only straightedge, compass,
and, furthermore, couldn't be squared even with the help of the conic sections (parabola, hyperbola, and
ellipse). But the ancient Greeks had no mathematical proof that it could not be squared.
That the circle could not be squared with Euclidean tools was not shown until 1882 when Lindemann
proved that pi is a transcendental number.
Book II
introduction
© 1996,2002 http://aleph0.clarku.edu/~djoyce/java/elements/elements.html
D.E.Joyce
Dept. Math. & Comp. Sci.
Clark University
Table of contents
● Definitions (11)
● Propositions (37)
Definitions
Definition 1.
Equal circles are those whose diameters are equal, or whose radii are equal.
Definition 2.
A straight line is said to touch a circle which, meeting the circle and being produced, does not cut
the circle.
Definition 3.
Circles are said to touch one another which meet one another but do not cut one another.
Definition 4.
Straight lines in a circle are said to be equally distant from the center when the perpendiculars
Definition 5.
And that straight line is said to be at a greater distance on which the greater perpendicular falls.
Definition 6.
A segment of a circle is the figure contained by a straight line and a circumference of a circle.
Definition 7.
An angle of a segment is that contained by a straight line and a circumference of a circle.
Definition 8.
An angle in a segment is the angle which, when a point is taken on the circumference of the
segment and straight lines are joined from it to the ends of the straight line which is the base of
the segment, is contained by the straight lines so joined.
Definition 9.
And, when the straight lines containing the angle cut off a circumference, the angle is said to
stand upon that circumference.
Definition 10.
A sector of a circle is the figure which, when an angle is constructed at the center of the circle, is
contained by the straight lines containing the angle and the circumference cut off by them.
Definition 11.
Similar segments of circles are those which admit equal angles, or in which the angles equal one
another.
Propositions
Proposition 1.
To find the center of a given circle.
Corollary. If in a circle a straight line cuts a straight line into two equal parts and at right angles,
then the center of the circle lies on the cutting straight line.
Proposition 2.
If two points are taken at random on the circumference of a circle, then the straight line joining
the points falls within the circle.
Proposition 3.
If a straight line passing through the center of a circle bisects a straight line not passing through
the center, then it also cuts it at right angles; and if it cuts it at right angles, then it also bisects it.
Proposition 4.
If in a circle two straight lines which do not pass through the center cut one another, then they do
not bisect one another.
Proposition 5.
If two circles cut one another, then they do not have the same center.
Proposition 6.
If two circles touch one another, then they do not have the same center.
Proposition 7.
If on the diameter of a circle a point is taken which is not the center of the circle, and from the
point straight lines fall upon the circle, then that is greatest on which passes through the center,
the remainder of the same diameter is least, and of the rest the nearer to the straight line through
the center is always greater than the more remote; and only two equal straight lines fall from the
point on the circle, one on each side of the least straight line.
Proposition 8.
If a point is taken outside a circle and from the point straight lines are drawn through to the circle,
one of which is through the center and the others are drawn at random, then, of the straight lines
which fall on the concave circumference, that through the center is greatest, while of the rest the
nearer to that through the center is always greater than the more remote, but, of the straight lines
falling on the convex circumference, that between the point and the diameter is least, while of the
rest the nearer to the least is always less than the more remote; and only two equal straight lines
fall on the circle from the point, one on each side of the least.
Proposition 9.
If a point is taken within a circle, and more than two equal straight lines fall from the point on the
circle, then the point taken is the center of the circle.
Proposition 10.
A circle does not cut a circle at more than two points.
Proposition 11.
If two circles touch one another internally, and their centers are taken, then the straight line
joining their centers, being produced, falls on the point of contact of the circles.
Proposition 12.
If two circles touch one another externally, then the straight line joining their centers passes
through the point of contact.
Proposition 13.
A circle does not touch another circle at more than one point whether it touches it internally or
externally..
Proposition 14.
Equal straight lines in a circle are equally distant from the center, and those which are equally
distant from the center equal one another.
Proposition 15.
Of straight lines in a circle the diameter is greatest, and of the rest the nearer to the center is
always greater than the more remote.
Proposition 16.
The straight line drawn at right angles to the diameter of a circle from its end will fall outside the
circle, and into the space between the straight line and the circumference another straight line
cannot be interposed, further the angle of the semicircle is greater, and the remaining angle less,
than any acute rectilinear angle.
Corollary. From this it is manifest that the straight line drawn at right angles to the diameter of a
circle from its end touches the circle.
Proposition 17.
From a given point to draw a straight line touching a given circle.
Proposition 18.
If a straight line touches a circle, and a straight line is joined from the center to the point of
contact, the straight line so joined will be perpendicular to the tangent.
Proposition 19.
If a straight line touches a circle, and from the point of contact a straight line is drawn at right
angles to the tangent, the center of the circle will be on the straight line so drawn.
Proposition 20.
In a circle the angle at the center is double the angle at the circumference when the angles have
the same circumference as base.
Proposition 21.
Proposition 22.
The sum of the opposite angles of quadrilaterals in circles equals two right angles.
Proposition 23.
On the same straight line there cannot be constructed two similar and unequal segments of circles
on the same side.
Proposition 24.
Similar segments of circles on equal straight lines equal one another.
Proposition 25.
Given a segment of a circle, to describe the complete circle of which it is a segment.
Proposition 26.
In equal circles equal angles stand on equal circumferences whether they stand at the centers or at
the circumferences.
Proposition 27.
In equal circles angles standing on equal circumferences equal one another whether they stand at
the centers or at the circumferences.
Proposition 28.
In equal circles equal straight lines cut off equal circumferences, the greater circumference equals
the greater and the less equals the less.
Proposition 29.
In equal circles straight lines that cut off equal circumferences are equal.
Proposition 30.
To bisect a given circumference.
Proposition 31.
In a circle the angle in the semicircle is right, that in a greater segment less than a right angle, and
that in a less segment greater than a right angle; further the angle of the greater segment is greater
than a right angle, and the angle of the less segment is less than a right angle.
Proposition 32.
If a straight line touches a circle, and from the point of contact there is drawn across, in the circle,
a straight line cutting the circle, then the angles which it makes with the tangent equal the angles
in the alternate segments of the circle.
Proposition 33.
On a given straight line to describe a segment of a circle admitting an angle equal to a given
rectilinear angle.
Proposition 34.
From a given circle to cut off a segment admitting an angle equal to a given rectilinear angle.
Proposition 35.
If in a circle two straight lines cut one another, then the rectangle contained by the segments of
the one equals the rectangle contained by the segments of the other.
Proposition 36.
If a point is taken outside a circle and two straight lines fall from it on the circle, and if one of
them cuts the circle and the other touches it, then the rectangle contained by the whole of the
straight line which cuts the circle and the straight line intercepted on it outside between the point
and the convex circumference equals the square on the tangent.
Proposition 37.
If a point is taken outside a circle and from the point there fall on the circle two straight lines, if
one of them cuts the circle, and the other falls on it, and if further the rectangle contained by the
whole of the straight line which cuts the circle and the straight line intercepted on it outside
between the point and the convex circumference equals the square on the straight line which falls
on the circle, then the straight line which falls on it touches the circle.
© 1996
D.E.Joyce
Clark University
Definition 1
Equal circles are those whose diameters are equal, or whose radii are equal.
This should not be a definition but a postulate or a theorem. The subject of the area of circles is
developed in Proposition XII.2 and this definition is not used there. Instead the concept of equality, or
rather, inequality is the same as it is in the rest of the Elements. For instance, if one figure is the
contained in another, then the first is less than the other. Thus, there is a prior defintion for the equality
of two figures.
Two circles are illustrated, namely circle BCD and circle FGH. The center of circle BCD is A, while the
center of circle FGH is E. They are equal by Euclid's definition since their diameters BC and FG are
equal, or since their radii AB and EF are equal.
© 1996, 1997
D.E.Joyce
Clark University
Definition 10
A sector of a circle is the figure which, when an angle is constructed at the center of the circle, is
contained by the straight lines containing the angle and the circumference cut off by them.
© 1996
D.E.Joyce
Clark University
Definitions 2 and 3
Def. 2. A straight line is said to touch a circle which, meeting the circle and being produced, does
not cut the circle.
Def. 3. Circles are said to touch one another which meet one another but do not cut one another.
Consider circle BCD in the figure. The line EF touches this circle at the point C. Another expression for
the same thing is that EF is tangent to the circle at C.
Two circles can touch each other either internally or externally. Circle HKL touches circle BCD
externally, while circle DMN touches circle BCD internally.
© 1996
D.E.Joyce
Clark University
Definitions 4 and 5
Def. 4. Straight lines in a circle are said to be equally distant from the center when the
perpendiculars drawn to them from the center are equal.
Def. 5. And that straight line is said to be at a greater distance on which the greater perpendicular
falls.
These definitions could have been broadened to distances from a line to a point, but Euclid's needs are
for this situation.
© 1996
D.E.Joyce
Clark University
Definitions 6 through 9
Def. 6. A segment of a circle is the figure contained by a straight line and a circumference of a
circle.
Def. 7. An angle of a segment is that contained by a straight line and a circumference of a circle.
Def. 8. An angle in a segment is the angle which, when a point is taken on the circumference of
the segment and straight lines are joined from it to the ends of the straight line which is the base
of the segment, is contained by the straight lines so joined.
Def. 9. And, when the straight lines containing the angle cut off a circumference, the angle is said
to stand upon that circumference.
A line in a circle, such as the line BC, divides the circle into two
segments, the small blue segment BDC, and the large yellow
segment BEC.
An example of an angle in a segment is the angle BFC in the yellow segment BEC. This angle BFC
stands upon the circumference (arc) BDC. Angles in segments are rectilinear, and they are important. In
proposition III.21, Euclid proves that all the angles in a given segment are equal.
© 1996
D.E.Joyce
Clark University
Proposition 1
Then, since AD equals DB, and DG is common, the two sides AD and DG equal the two I.
sides BD and DG respectively. And the base GA equals the base GB, for they are radii, Def.15
therefore the angle ADG equals the angle GDB. I.8
But, when a straight line standing on a straight line makes the adjacent angles equal to one I.Def.10
another, each of the equal angles is right, therefore the angle GDB is right.
But the angle FDB is also right, therefore the angle FDB equals the angle GDB, the greater
equals the less, which is impossible. Therefore G is not the center of the circle ABC.
Corollary
From this it is manifest that if in a circle a straight line cuts a straight line into two equal parts and at
right angles, then the center of the circle lies on the cutting straight line.
Since the definition of a circle, I.Def.15, includes the existence of a center, Euclid is justified in taking a
point G as the center.
In this proof G is shown to lie on the perpendicular bisector of the line AB. He leaves to the reader to
show that G actually is the point F on the perpendicular bisector, but that's clear since only the midpoint
F is equidistant from the two points C and E on the circle. From that observation it also follows that the
center of a circle is unique, although the uniqueness can easily be proved in other ways.
As Todhunter remarked, Euclid implicitly assumes that the perpendicular bisector of AB actually
intersects the circle in points C and E.
About half the proofs in Book III and several of those in Book IV begin with taking the center of a
circle, but in plane geometry, it isn't necessary to invoke this proposition III.1 since the only way that
circles can occur is if they are constructed around a center to begin with. Even in solid geometry, the
center of a circle is usually known so that III.1 isn't necessary. Indeed, that is the case whenever the
center is needed in Euclid's books on solid geometry (see XI.23, XIII.9 through XIII.13, and XIII.16).
Sections of spheres cut by planes are also circles as are certain sections of cylinders and cones, but in
these cases too, the centers can easily be found without recourse to III.1. Thus, III.1 redundant, although
it is an interesting construction.
Select topic
Book III introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 1
Then, since AD equals DB, and DG is common, the two sides AD and DG equal the two I.
sides BD and DG respectively. And the base GA equals the base GB, for they are radii, Def.15
therefore the angle ADG equals the angle GDB. I.8
But, when a straight line standing on a straight line makes the adjacent angles equal to one I.Def.10
another, each of the equal angles is right, therefore the angle GDB is right.
But the angle FDB is also right, therefore the angle FDB equals the angle GDB, the greater
equals the less, which is impossible. Therefore G is not the center of the circle ABC.
Corollary
From this it is manifest that if in a circle a straight line cuts a straight line into two equal parts and at
right angles, then the center of the circle lies on the cutting straight line.
Since the definition of a circle, I.Def.15, includes the existence of a center, Euclid is justified in taking a
point G as the center.
In this proof G is shown to lie on the perpendicular bisector of the line AB. He leaves to the reader to
show that G actually is the point F on the perpendicular bisector, but that's clear since only the midpoint
F is equidistant from the two points C and E on the circle. From that observation it also follows that the
center of a circle is unique, although the uniqueness can easily be proved in other ways.
As Todhunter remarked, Euclid implicitly assumes that the perpendicular bisector of AB actually
intersects the circle in points C and E.
About half the proofs in Book III and several of those in Book IV begin with taking the center of a
circle, but in plane geometry, it isn't necessary to invoke this proposition III.1 since the only way that
circles can occur is if they are constructed around a center to begin with. Even in solid geometry, the
center of a circle is usually known so that III.1 isn't necessary. Indeed, that is the case whenever the
center is needed in Euclid's books on solid geometry (see XI.23, XIII.9 through XIII.13, and XIII.16).
Sections of spheres cut by planes are also circles as are certain sections of cylinders and cones, but in
these cases too, the centers can easily be found without recourse to III.1. Thus, III.1 redundant, although
it is an interesting construction.
Select topic
Book III introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 2
If two points are taken at random on the circumference of a circle, then the straight line joining
the points falls within the circle.
Let ABC be a circle, and let two points A and B be taken at random on its circumference.
I say that the straight line joined from A to B falls within the circle.
I.
Then, since DA equals DB, the angle DAE also equals Def.15
the angle DBE. I.5
And the angle DAE equals the angle DBE, therefore the angle DEB is greater than the angle
DBE. And the side opposite the greater angle is greater, therefore DB is greater than DE. But I.19
DB equals DF, therefore DF is greater than DE, the less greater than the greater, which is I.Def.15
impossible.
Therefore the straight line joined from A to B does not fall outside the circle.
Similarly we can prove that neither does it fall on the circumference itself, therefore it falls
within.
Therefore if two points are taken at random on the circumference of a circle, then the straight line
joining the points falls within the circle.
Q.E.D.
The figure for this proposition is rather strange, but that is necessary since it refers to a hypothetical
situation which is shown to be impossible. In this figure AEB is supposed to be a straight line that lies on
outside the circle. There are other impossible figures in later propositions in this Book.
That Euclid even has this proposition is remarkable. Of course, it should be included, but there are
equally obvious statements (but difficult to prove) left out in earlier books. For instance, that the two
circles constructed in a plane on a line AB intersect is not proved although it is used in I.1. This indicates
that more care has been given to the foundations for this book than for the previous books.
Euclid leaves to the reader to prove that AB cannot lie on the circumference, and that is not particularly
difficult to prove.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 3
If a straight line passing through the center of a circle bisects a straight line not passing
through the center, then it also cuts it at right angles; and if it cuts it at right angles, then it
also bisects it.
Let a straight line CD passing through the center of a circle ABC bisect a straight line AB not
passing through the center at the point F.
Take the center E of the circle ABC, and join EA and EB. III.1
Therefore CD, which passes through the center and bisects AB which does not pass through
the center, also cuts it at right angles.
I say that it also bisects it, that is, that AF equals FB.
For, with the same construction, since EA equals EB, the angle EAF also equals the angle I.5
EBF.
But the right angle AFE equals the right angle BFE, therefore EAF and EBF are two
triangles having two angles equal to two angles and one side equal to one side, namely EF, I.26
which is common to them, and opposite one of the equal angles. Therefore they also have
the remaining sides equal to the remaining sides
Therefore if a straight line passing through the center of a circle bisects a straight line not passing
through the center, then it also cuts it at right angles; and if it cuts it at right angles, then it also
bisects it.
Q.E.D.
This proposition is used in the next one, a few others in Book III, and XII.16.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 4
If in a circle two straight lines which do not pass through the center cut one another, then they
do not bisect one another.
Let ABCD be a circle, and in it let the two straight lines AC and BD, which do not pass through
the center, cut one another at E.
For, if so, let them bisect one another, so that AE equals EC, and BE III.1
equals ED. Take the center F of the circle ABCD. Join FE.
Again, since a straight line FE bisects a straight line BD, it also cuts it at right angles. Therefore III.3
the angle FEB is right.
But the angle FEA was also proved right, therefore the angle FEA equals the angle FEB, the less
equals the greater, which is impossible.
Therefore if in a circle two straight lines which do not pass through the center cut one another, then
they do not bisect one another.
Q.E.D.
This proposition is not used in the rest of the Elements. The contrapositive of this statement is more
positive: if two straight lines in a circle bisect each other, then they meet at the center.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 5
If two circles cut one another, then they do not have the same center.
Let the circles ABC and CDG cut one another at the points B and C.
But EC was proved equal to EF also, therefore EF also equals EG, the less equals the greater
which is impossible.
Therefore the point E is not the center of the circles ABC and CDG.
Therefore if two circles cut one another, then they do not have the same center.
Q.E.D.
Note that no use was made in the proof of the point B. That means the proof actually shows that if two
circles meet, then they do not have the same center, and that covers not only this proposition but the
next, too, where the two touch each other.
This proposition is used in III.10 which states that circles cannot intersect at more than two points.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 6
If two circles touch one another, then they do not have the same center.
Let the two circles ABC and CDE touch one another at the
point C.
Then, since the point F is the center of the circle ABC, FC equals FB. Again, since the point I.Def.15
F is the center of the circle CDE, FC equals FE.
But FC was proved equal to FB, therefore FE also equals FB, the less equals the greater,
which is impossible.
Therefore if two circles touch one another, then they do not have the same center.
Q.E.D.
As mentioned before, this proposition is almost the same as the previous. Both could be included in one
statement: circles that meet don't have the same center, or the contrapositive: concentric circles don't
meet.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 7
If on the diameter of a circle a point is taken which is not the center of the circle, and from the
point straight lines fall upon the circle, then that is greatest on which passes through the
center, the remainder of the same diameter is least, and of the rest the nearer to the straight
line through the center is always greater than the more remote; and only two equal straight
lines fall from the point on the circle, one on each side of the least straight line.
Let ABCD be a circle, and let AD be a diameter of it. Let F be a point F on AD which is not the
center of the circle. Let E be the center of the circle. Let straight lines FB, FC, and FG fall upon
the circle ABCD from F.
I say that FA is greatest, FD is least, and of the rest FB is greater than FC, and FC greater than
FG.
Then, since in any triangle the sum of any two sides is greater than I.20
the remaining one, the sum of EB and EF is greater than BF.
Again, since BE equals CE, and FE is common, the two sides BE and
EF equal the two sides CE and EF. But the angle BEF is also greater I.24
than the angle CEF, therefore the base BF is greater than the base CF.
Again, since the sum of GF and FE is greater than EG, and EG equals ED, the sum of GF and I.20
FE is greater than ED.
Subtract EF from each. Therefore the remainder GF is greater than the remainder FD.
Therefore FA is greatest, FD is least, FB is greater than FC, and FC greater than FG.
I say also that from the point F only two equal straight lines fall on the circle ABCD, one on
each side of the least FD.
Construct the angle FEH equal to the angle GEF on the straight line EF and at the point E on I.23
it. Join FH.
I say again that another straight line equal to FG does not fall on
the circle from the point F.
Then, since FK equals FG, and FH equals FG, FK also equals FH, the nearer to the straight
Above
line through the center being thus equal to the more remote, which is impossible.
Therefore another straight line equal to GF does not fall from the point F upon the circle.
Therefore only one straight line so falls.
Therefore if on the diameter of a circle a point is taken which is not the center of the circle, and from
the point straight lines fall upon the circle, then that is greatest on which passes through the center,
the remainder of the same diameter is least, and of the rest the nearer to the straight line through the
center is always greater than the more remote; and only two equal straight lines fall from the point on
the circle, one on each side of the least straight line.
Q.E.D.
The statement of this proposition is daunting. It concerns the distances from a point F inside a circle to
the points on the circumference. The point F is assumed not to be the center. If a diameter AD is passed
through F, then one of the points A is the point on the circumference furthest from F and the other D is
the closest. As a point travels the circumference from A to D it gets closer to F. The final part of the
statement is that if G is one point on the circumference, then there is exactly one other point H on the
circumference the same distance from F (assuming, of course, that G is neither A nor D).
Note
There is some ambiguity in the statement of this proposition. It is not clear what the phrase "the nearer to
the straight line through the center" means. It may well refer to the angle, so that FB is considered nearer
to FA than FC since the angle BFA is less than the angle CFA. If so, there there is a missing detail in the
proof, as De Morgan pointed out. It is declared that the angle BEF is greater than the angle CEF, but that
hasn't been proved. DeMorgan and others have described various ways to fill this logical gap.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 8
If a point is taken outside a circle and from the point straight lines are drawn through to the
circle, one of which is through the center and the others are drawn at random, then, of the
straight lines which fall on the concave circumference, that through the center is greatest,
while of the rest the nearer to that through the center is always greater than the more remote,
but, of the straight lines falling on the convex circumference, that between the point and the
diameter is least, while of the rest the nearer to the least is always less than the more remote;
and only two equal straight lines fall on the circle from the point, one on each side of the least.
Let ABC be a circle, and let a point D be taken outside ABC. Let straight lines DA, DE, DF, and
DC be drawn through from D, and let DA be drawn through the center.
I say that, of the straight lines falling on the concave circumference AEFC, the straight line DA
through the center is greatest, while DE is greater than DF, and DF greater than DC. But, of the
straight lines falling on the convex circumference HLKG, the straight line DG between the point
and the diameter AG is least, and the nearer to the least DG is always less than the more remote,
namely DK is less than DL, and DL is less than DH.
Take the center M of the circle ABC. Join ME, MF, MC, MK, ML, and III.1
MH.
But the sum of EM and MD is greater than ED, therefore AD is also I.20
greater than ED.
Similarly we can prove that FD is greater than CD. Therefore DA is greatest, while DE is
greater than DF, and DF is greater than DC.
Next, since the sum of MK and KD is greater than MD, and MG equals MK, therefore the I.20
remainder KD is greater than the remainder GD, so that GD is less than KD.
And, since on MD, one of the sides of the triangle MLD, two straight lines MK and KD are
constructed meeting within the triangle, therefore the sum of MK and KD is less than the sum of I.21
ML and LD. And MK equals ML, therefore the remainder DK is less than the remainder DL.
Similarly we can prove that DL is also less than DH. Therefore DG is least, while DK is less
than DL, and DL is less than DH.
I say also that only two equal straight lines will fall from the point D on the circle, one on
each side of the least DG.
Construct the angle DMB equal to the angle KMD on the straight line I.23
MD and at the point M on it. Join DB.
I say that no other straight line equal to the straight line DK falls on
the circle from the point D.
For, if possible, let a straight line so fall, and let it be DN. Then, since
DK equals DN, and DK equals DB, DB also equals DN, that is, the
Above
nearer to the least DG equal to the more remote, which was proved
impossible.
Therefore no more than two equal straight lines fall on the circle ABC from the point D, one
on each side of DG the least.
Therefore if a point is taken outside a circle and from the point straight lines are drawn through to the
circle, one of which is through the center and the others are drawn at random, then, of the straight
lines which fall on the concave circumference, that through the center is greatest, while of the rest the
nearer to that through the center is always greater than the more remote, but, of the straight lines
falling on the convex circumference, that between the point and the diameter is least, while of the rest
the nearer to the least is always less than the more remote; and only two equal straight lines fall on
the circle from the point, one on each side of the least.
Q.E.D.
This proposition has a statement even more complicated than the previous proposition. This one deals
with the distances from a point D outside a circle to the points on the circumference. If the diameter AG
extended passes through D, then one of its endpoints G is the point on the circumference closest to D
and the other A is furthest. As a point moves along the circumference from A to D it gets closer to D.
Euclid considers two parts of the circumference, the convex part is the near part exposed to the point D,
while the concave part is the part on the far side of the circle. The final part of the statement is that if K
is one point on the circumference, then there is exactly one other point B on the circumference the same
distance from D (assuming, of course, that K is neither G nor A).
Note
There is a logical gap in the proof of this proposition similar to that in the previous proposition. Again,
various ways have been proposed to fill it.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 9
If a point is taken within a circle, and more than two equal straight lines fall from the point on
the circle, then the point taken is the center of the circle.
Then, since AE equals EB, and ED is common, the two sides AE and ED equal the two sides
BE and ED, and the base DA equals the base DB, therefore the angle AED equals the angle I.8
BED.
Therefore the angles AED and BED are each right. Therefore GK cuts AB into two equal
parts and at right angles.
And since, if in a circle a straight line cuts a straight line into two equal parts and at right
angles, the center of the circle is on the cutting straight line, therefore the center of the circle III.1,Cor
is on GK.
For the same reason the center of the circle ABC is also on HL.
And the straight lines GK and HL have no other point common but the point D, therefore the
point D is the center of the circle ABC.
Therefore if a point is taken within a circle, and more than two equal straight lines fall from the point
on the circle, then the point taken is the center of the circle.
Q.E.D.
The statement of this proposition is already covered by the last part of Proposition III.7 which says for a
point in a circle that is not the center at most two points lie on the circumference at any distance from
that point.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 10
But it was also proved to lie on AC, and the straight lines AC and NO meet at no point
except at P, therefore the point P is the center of the circle ABC.
Similarly we can prove that P is also the center of the circle DEF, therefore the two circles III.5
ABC and DEF which cut one another have the same center P, which is impossible.
Therefore a circle does not cut a circle at more than two points.
Q.E.D.
The figure is another impossible figure. Both curves are supposed to be circumferences of circles, but, of
course, they cannot both be drawn as circles since the situation is proved not to occur. Although Euclid
names four points where the circles meet, only three, B, G, and H, are used in the proof.
The proof actually shows that the two circles cannot meet in more than two points, where "meet" could
be either cut or touch.
Heath remarks that the lines bisecting BG and BH have not been shown to meet. In fact, they have, since
the center of the circle ABC has been shown to be on both.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 11
If two circles touch one another internally, and their centers are taken, then the straight line
joining their centers, being produced, falls on the point of contact of the circles.
Let the two circles ABC and ADE touch one another internally at the point A, and let the centers III.1
F and G of the circles ABC and ADE be taken.
I say that the straight line joined from G to F and produced falls on A.
Therefore if two circles touch one another internally, and their centers are taken, then the straight
line joining their centers, being produced, falls on the point of contact of the circles.
Q.E.D.
In order to carry through the proof, in particular so that FA = FH, the circle ABC needs to be the larger
circle.
Various conclusions in the proof are based on the figure rather than rigorous deductive reasoning.
Camerer and others have suggested ways of filling the gaps.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 12
If two circles touch one another externally, then the straight line joining their centers passes
through the point of contact.
Let the two circles ABC and ADE touch one another externally at the point A. Take the center F III.1
of ABC, and the center G of ADE.
I say that the straight line joined from F to G passes through the point of contact at A.
Therefore the straight line joined from F to G does not fail to pass through the point of contact
at A, therefore it passes through it.
Therefore if two circles touch one another externally, then the straight line joining their centers
passes through the point of contact.
Q.E.D.
This proposition was certainly added to the Elements after Euclid, perhaps by Heron or a later
commentator.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 13
A circle does not touch another circle at more than one point whether it touches it internally or
externally.
For, if possible, let the circle ABDC touch the circle EBFD, first internally, at more points
than one, namely D and B.
Take the center G of the circle ABDC and the center H of EBFD. III.1
Therefore a circle does not touch a circle internally at more points than one.
Therefore a circle does not touch a circle externally at more points than one.
Therefore a circle does not touch another circle at more than one point whether it touches it
internally or externally.
Q.E.D.
In the second impossible figure there are three curves connecting A to C. The two circles are not
supposed to cut each other, but just to touch each other at the two points A and C, and the straight line
AC should lie between the two circles and not within either one.
There are logical flaws in this proof similar to those in the last two proofs.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 14
Equal straight lines in a circle are equally distant from the center, and those which are equally
distant from the center equal one another.
Let AB and CD be equal straight lines in a circle ABDC.
Take the center E of the circle ABDC. Draw EF and EG from E III.1
perpendicular to AB and CD, and join AE and EC. I.12
For the same reason CD is also double CG. But AB equals CD, therefore AF also equals CG.
Also, since AE equals EC, the square on AE also equals the square on EC. But the sum of
the squares on AF and EF equals the square on AE, for the angle at F is right, and the sum of
the squares on EG and GC equals the square on EC, for the angle at G is right. Therefore the I.47
sum of the squares on AF and FE equals the sum of the squares on CG and GE, of which the
square on AF equals the square on CG, for AF equals CG. Therefore the remaining square
on FE equals the square on EG. Therefore EF equals EG.
But straight lines in a circle are said to be equally distant from the center when the
perpendiculars drawn to them from the center are equal. Therefore AB and CD are equally III.Def.4
distant from the center.
Next, let the straight lines AB and CD be equally distant from the center, that is, let EF equal
EG.
For, with the same construction, we can prove, as before, that AB is double AF, and CD
double CG. And, since AE equals CE, the square on AE equals the square on CE. But the I.47
sum of the squares on EF and FA equals the square on AE, and the sum of the squares on
EG and GC equals the square on CE.
Therefore the sum of the squares on EF and FA equals the sum of the squares on EG and
GC, of which the square on EF equals the square on EG, for EF equals EG. Therefore the
remaining square on AF equals the square on CG. Therefore AF equals CG. And AB is
double AF, and CD double CG, therefore AB equals CD.
Therefore equal straight lines in a circle are equally distant from the center, and those which are
equally distant from the center equal one another.
Q.E.D.
Note how Euclid has proved twice in the course of this proof the side-side-right angle congruence
theorem. See the note after I.26 about congruence theorems for triangles.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 15
Of straight lines in a circle the diameter is greatest, and of the rest the nearer to the center is
always greater than the more remote.
Let ABCD be a circle, AD its diameter, and E its center. Let BC be nearer to the center AD,
and FG more remote.
Again, since AE equals EM, and ED equals EN, AD equals the sum of ME and EN.
But the sum of ME and EN is greater than MN, and MN equals BC, therefore AD is greater I.20
than BC.
And, since the two sides ME and EN equal the two sides FE and EG, and the angle MEN I.24
greater than the angle FEG, therefore the base MN is greater than the base FG.
Therefore of straight lines in a circle the diameter is greatest, and of the rest the nearer to the center
is always greater than the more remote.
Q.E.D.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 16
The straight line drawn at right angles to the diameter of a circle from its end will fall outside
the circle, and into the space between the straight line and the circumference another straight
line cannot be interposed, further the angle of the semicircle is greater, and the remaining
angle less, than any acute rectilinear angle.
Let ABC be a circle about D as center and AB as diameter.
I say that the straight line drawn from A at right angles to AB from its end will fall outside the
circle.
Since DA equals DC, the angle DAC also equals the I.5
angle ACD.
But the angle DAC is right, therefore the angle ACD is also right. Thus, in the triangle ACD, I.17
the two angles DAC and ACD equal two right angles, which is impossible.
Therefore the straight line drawn from the point A at right angles to BA will not fall within the
circle.
Similarly we can prove that neither will it fall on the circumference, therefore it will fall
outside.
I say next that into the space between the straight line AE and the circumference CHA another
straight line cannot be interposed.
For, if possible, let another straight line be so interposed, as FA. Draw DG from the point D I.12
perpendicular to FA.
Then, since the angle AGD is right, and the angle DAG is less than a right angle, AD is greater I.17
than DG. I.19
But DA equals DH, therefore DH is greater than DG, the less greater than the greater, which is
impossible.
Therefore another straight line cannot be interposed into the space between the straight line
and the circumference.
I say further that the angle of the semicircle contained by the straight line BA and the
circumference CHA is greater than any acute rectilinear angle, and the remaining angle
contained by the circumference CHA and the straight line AE is less than any acute rectilinear
angle.
For, if there is any rectilinear angle greater than the angle contained by the straight line BA
and the circumference CHA, and any rectilinear angle less than the angle contained by the
circumference CHA and the straight line AE, then into the space between the circumference
and the straight line AE a straight line will be interposed such as will make an angle contained
by straight lines which is greater than the angle contained by the straight line BA and the
circumference CHA, and another angle contained by straight lines which is less than the angle
contained by the circumference CHA and the straight line AE.
But such a straight line cannot be interposed, therefore there will not be any acute angle
contained by straight lines which is greater than the angle contained by the straight line BA
Above
and the circumference CHA, nor yet any acute angle contained by straight lines which is less
than the angle contained by the circumference CHA and the straight line AE.
Therefore the straight line drawn at right angles to the diameter of a circle from its end will fall
outside the circle, and into the space between the straight line and the circumference another straight
line cannot be interposed, further the angle of the semicircle is greater, and the remaining angle less,
than any acute rectilinear angle.
Q.E.D.
Corollary
From this it is manifest that the straight line drawn at right angles to the diameter of a circle from its
end touches the circle.
This proposition is used in the proof of proposition IV.4 and two others in Book IV. The corollary is
used in III.33, III.37, a few propositions in Book IV, and XII.16.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 17
I say that AB has been drawn from the point A touching the
circle BCD.
For, since E is the center of the circles BCD and AFG, EA equals EF, and ED equals EB.
Therefore the two sides AE and EB equal the two sides FE and ED, and they contain a
common angle, the angle at E, therefore the base DF equals the base AB, and the triangle I.4
DEF equals the triangle BEA, and the remaining angles to the remaining angles, therefore
the angle EDF equals the angle EBA.
But the angle EDF is right, therefore the angle EBA is also right.
Now EB is a radius, and the straight line drawn at right angles to the diameter of a circle, III.16,Cor
from its end, touches the circle, therefore AB touches the circle BCD.
Therefore from the given point A the straight line AB has been drawn touching the circle
BCD.
Q.E.F.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 18
If a straight line touches a circle, and a straight line is joined from the center to the point of
contact, the straight line so joined will be perpendicular to the tangent.
For let a straight line DE touch the circle ABC at the point C. Take the center F of the circle III.1
ABC, and join FC from F to C.
Then, since the angle FGC is right, the angle FCG is acute, and
I.17
the side opposite the greater angle is greater, therefore FC is I.19
greater than FG.
Similarly we can prove that neither is any other straight line except FC. Therefore FC is
perpendicular to DE.
Therefore if a straight line touches a circle, and a straight line is joined from the center to the point of
contact, the straight line so joined will be perpendicular to the tangent.
Q.E.D.
This proposition is used in a few propositions in Books III and IV beginning with III.36.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 19
If a straight line touches a circle, and from the point of contact a straight line is drawn at right
angles to the tangent, the center of the circle will be on the straight line so drawn.
For let a straight line DE touch the circle ABC at the point C. Draw CA from C at right angles I.11
to DE.
For suppose it is not, but, if possible, let F be the center, and join
CF.
Since a straight line DE touches the circle ABC, and FC has been
joined from the center to the point of contact, FC is perpendicular III.18
to DE. Therefore the angle FCE is right.
But the angle ACE is also right, therefore the angle FCE equals the
angle ACE, the less equals the greater, which is impossible.
Similarly we can prove that neither is any other point except a point on AC.
Therefore if a straight line touches a circle, and from the point of contact a straight line is drawn at
right angles to the tangent, the center of the circle will be on the straight line so drawn.
Q.E.D.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 20
In a circle the angle at the center is double the angle at the circumference when the angles
have the same circumference as base.
Let ABC be a circle, let the angle BEC be an angle at its center, and the angle BAC an angle at
the circumference, and let them have the same circumference BC as base.
Then, since EA equals EB, the angle EAB also equals the angle
EBA. Therefore the sum of the angles the angles EAB and EBA I.5
is double the angle EAB.
But the angle BEF equals the sum of the angles EAB and EBA, I.32
therefore the angle BEF, is also double the angle EAB.
For the same reason the angle FEC is also double the angle EAC.
Therefore the whole angle BEC is double the whole angle BAC.
Again let another straight line be inflected, and let there be another angle BDC. Join DE and
produced it to G.
Similarly then we can prove that the angle GEC is double the angle EDC, of which the angle
GEB is double the angle EDB. Therefore the remaining angle BEC is double the angle BDC.
Therefore in a circle the angle at the center is double the angle at the circumference when the angles
have the same circumference as base.
Q.E.D.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 21
I say that the angles BAD and BED equal one another.
Take the center F of the circle ABCD, and join BF and FD. III.1
Now, since the angle BFD is at the center, and the angle BAD
at the circumference, and they have the same circumference III.20
BCD as base, therefore the angle BFD is double the angle BAD.
For the same reason the angle BFD is also double the angle
BED. Therefore the angle BAD equals the angle BED.
Therefore in a circle the angles in the same segment equal one another.
Q.E.D.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 22
The sum of the opposite angles of quadrilaterals in circles equals two right angles.
Let ABCD be a circle, and let ABCD be a quadrilateral in it.
I say that the sum of the opposite angles equals two right angles.
Then, since in any triangle the sum of the three angles equals two
right angles, the sum of the three angles CAB, ABC, and BCA of the I.32
triangle ABC equals two right angles.
But the angle CAB equals the angle BDC, for they are in the same
segment BADC, and the angle ACB equals the angle ADB, for they III.21
are in the same segment ADCB, therefore the whole angle ADC
equals the sum of the angles BAC and ACB.
Add the angle ABC to each. Therefore the sum of the angles ABC, BAC, and ACB equals the
sum of the angles ABC and ADC. But the sum of the angles ABC, BAC, and ACB equals two
right angles, therefore the sum of the angles ABC and ADC also equal two right angles.
Similarly we can prove that the sum of the angles BAD and DCB also equals two right angles.
Therefore the sum of the opposite angles of quadrilaterals in circles equals two right angles.
Q.E.D.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 23
On the same straight line there cannot be constructed two similar and unequal segments of
circles on the same side.
For, if possible, on the same straight line AB let two
similar and unequal segments of circles ACB and ADB
be constructed on the same side. Draw ACD through,
and join CB and DB.
Then, since the segment ACB is similar to the segment ADB, and similar segments of III.
circles are those which admit equal angles, the angle ACB equals the angle ADB, the Def.11
exterior to the interior, which is impossible. I.16
Therefore on the same straight line there cannot be constructed two similar and unequal segments of
circles on the same side.
Q.E.D.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 24
For, if the segment AEB is superposed on CFD, and if the point A is placed on C and the
straight line AB on CD, then the point B coincides with the point D, because AB equals CD,
and, AB coinciding with CD, the segment AEB also coincides with CFD.
For, if the straight line AB coincides with CD but the segment AEB does not coincide with
III.23
CFD, then it either falls within it, or outside it, or it falls awry, as CGD, and a circle cuts a III.10
circle at more points than two, which is impossible.
Therefore, if the straight line AB is superposed on CD, then the segment AEB does not fail to C.N.4
coincide with CFD also, therefore it coincides with it and equals it.
Therefore similar segments of circles on equal straight lines equal one another.
Q.E.D.
The proof here uses the method of superposition which was also used for I.4 and I.8.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 25
It is required to describe the complete circle belonging to the segment ABC, that is, of which it
is a segment.
I.10
Bisect AC at D, draw DB from the point D at right angles to AC, and join AB. I.11
The angle ABD is then greater than, equal to, or less than the angle BAD.
First let it be greater. Construct the angle BAE on the straight line BA,
and at the point A on it, equal to the angle ABD. Draw DB through to E, I.23
and join EC.
Then, since the angle ABE equals the angle BAE, the straight line EB I.6
also equals EA.
And, since AD equals DC, and DE is common, the two sides AD and DE equal the two sides CD
and DE respectively, and the angle ADE equals the angle CDE, for each is right, therefore the I.4
base AE equals the base CE.
But AE was proved equal to BE, therefore be also equals CE. Therefore the three straight lines
AE, EB, and EC equal one another.
Therefore the circle drawn with center E and radius one of the straight lines AE, EB, or EC also III.9
passes through the remaining points and has been completed.
Therefore, given a segment of a circle, the complete circle has been described.
And it is manifest that the segment ABC is less than a semicircle, because the center E happens
to be outside it.
Similarly, even if the angle ABD equals the angle BAD and AD being
equal to each of the two BD and DC, the three straight lines DA, DB,
and DC will equal one another, D will be the center of the completed
circle, and ABC will clearly be a semicircle.
But, if the angle ABD is less than the angle BAD, and if we construct, on
the straight line BA and at the point A on it, an angle equal to the angle I.23
ABD, the center will fall on DB within the segment ABC, and the
segment ABC will clearly be greater than a semicircle.
Therefore, given a segment of a circle, the complete circle has been described.
Q.E.F.
The construction in this proposition is not used in the rest of the Elements.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 26
In equal circles equal angles stand on equal circumferences whether they stand at the centers
or at the circumferences.
Let ABC and DEF be equal circles, and in them let there be equal angles, namely at the
centers the angles BGC and EHF, and at the circumferences the angles BAC and EDF.
Now, since the circles ABC and DEF are equal, the radii are equal.
Thus the two straight lines BG and GC equal the two straight lines EH and HF, and the I.4
angle at G equals the angle at H, therefore the base BC equals the base EF.
And, since the angle at A equals the angle at D, the segment BAC is similar to the segment III.Def.11
EDF, and they are upon equal straight lines.
But similar segments of circles on equal straight lines equal one another, therefore the
segment BAC equals EDF. But the whole circle ABC also equals the whole circle DEF, III.24
therefore the remaining circumference BKC equals the circumference ELF.
Therefore in equal circles equal angles stand on equal circumferences whether they stand at the
centers or at the circumferences.
Q.E.D.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 27
In equal circles angles standing on equal circumferences equal one another whether they stand
at the centers or at the circumferences.
For in equal circles ABC and DEF, on equal circumferences BC and EF, let the angles BGC
and EHF stand at the centers G and H, and the angles BAC and EDF at the circumferences.
I say that the angle BGC equals the angle EHF, and the angle BAC equals the angle EDF.
For, if the angle BGC does not equal the angle EHF, one of them is greater. Let the angle BGC
be greater. Construct the angle BGK equal to the angle EHF on the straight line BG and at the I.23
point G on it.
Now equal angles stand on equal circumferences when they are at the centers, therefore the I.26
circumference BK equals the circumference EF.
But EF equals BC, therefore BK also equals BC, the less equals the greater, which is
impossible.
Therefore the angle BGC is not unequal to the angle EHF, therefore it equals it.
And the angle at A is half of the angle BGC, and the angle at D half of the angle EHF, therefore III.20
the angle at A also equals the angle at D.
Therefore in equal circles angles standing on equal circumferences equal one another whether they
stand at the centers or at the circumferences.
Q.E.D.
This proposition is used in a few propositions in Books III, IV, VI, and XII starting with III.29.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 28
In equal circles equal straight lines cut off equal circumferences, the greater circumference
equals the greater and the less equals the less.
Let ABC and DEF be equal circles, and in the circles let AB and DE be equal straight lines
cutting off ACB and DFE as greater circumferences and AGB and DHE as lesser.
Now, since the circles are equal, the radii are also equal, therefore the two sides AK and KB
equal the two sides DL and LE, and the base AB equals the base DE, therefore the angle AKB I.8
equals the angle DLE.
But equal angles stand on equal circumferences when they are at the centers, therefore the III.26
circumference AGB equals DHE.
And the whole circle ABC also equals the whole circle DEF, therefore the remaining
circumference ACB also equals the remaining circumference DFE.
Therefore in equal circles equal straight lines cut off equal circumferences, the greater circumference
equals the greater and the less equals the less.
Q.E.D.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 29
In equal circles straight lines that cut off equal circumferences are equal.
Take the centers K and L of the circles. Join BK, KC, EL, and LF. III.1
Now, since the circumference BGC equals the circumference EHF, the angle BKC also equals III.27
the angle ELF.
And, since the circles ABC and DEF are equal, the radii are also equal, therefore the two sides
BK and KC equal the two sides EL and LF, and they contain equal angles, therefore the base I.4
BC equals the base EF.
Therefore in equal circles straight lines that cut off equal circumferences are equal.
Q.E.D.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 30
Then, since AC equals CB, and CD is common, the two sides AC and CD equal the two sides
BC and CD, and the angle ACD equals the angle BCD, for each is right, therefore the base AD I.4
equals the base DB.
But equal straight lines cut off equal circumferences, the greater equal to the greater, and the
less to the less, and each of the circumferences AD and DB is less than a semicircle, therefore III.28
the circumference AD equals the circumference DB.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 31
In a circle the angle in the semicircle is right, that in a greater segment less than a right angle,
and that in a less segment greater than a right angle; further the angle of the greater segment
is greater than a right angle, and the angle of the less segment is less than a right angle.
Let ABCD be a circle, let BC be its diameter, and E its center. Join BA, AC, AD, and DC.
I say that the angle BAC in the semicircle BAC is right, the angle ABC in the segment ABC
greater than the semicircle is less than a right angle, and the angle ADC in the segment ADC
less than the semicircle is greater than a right angle.
Then, since BE equals EA, the angle ABE also equals the angle
BAE. Again, since CE equals EA, the angle ACE also equals the I.5
angle CAE. Therefore the whole angle BAC equals the sum of the
two angles ABC and ACB.
But the angle FAC exterior to the triangle ABC also equals the sum of the two angles ABC and
ACB. Therefore the angle BAC also equals the angle FAC. Therefore each is right. Therefore I.32
the angle BAC in the semicircle BAC is right.
Next, since in the triangle ABC the sum of the two angles ABC and BAC is less than two right
angles, and the angle BAC is a right angle, the angle ABC is less than a right angle. And it is I.17
the angle in the segment ABC greater than the semicircle.
Next, since ABCD is a quadrilateral in a circle, and the sum of the opposite angles of
quadrilaterals in circles equals two right angles, while the angle ABC is less than a right angle, III.22
therefore the remaining angle ADC is greater than a right angle. And it is the angle in the
segment ADC less than the semicircle.
I say further that the angle of the greater segment, namely that contained by the circumference
ABC and the straight line AC, is greater than a right angle, and the angle of the less segment,
namely that contained by the circumference ADC and the straight line AC, is less than a right
angle.
This is at once manifest. For, since the angle contained by the straight lines BA and AC is right,
the angle contained by the circumference ABC and the straight line AC is greater than a right
angle.
Again, since the angle contained by the straight lines AC and AF is right, the angle contained
by the straight line CA and the circumference ADC is less than a right angle.
Therefore in a circle the angle in the semicircle is right, that in a greater segment less than a right
angle, and that in a less segment greater than a right angle; further the angle of the greater segment
is greater than a right angle, and the angle of the less segment is less than a right angle.
Q.E.D.
This proposition is used in III.32 and in each of the rest of the geometry books, namely, Books IV, VI,
XI, XII, XIII. It is also used in Book X.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 32
If a straight line touches a circle, and from the point of contact there is drawn across, in the
circle, a straight line cutting the circle, then the angles which it makes with the tangent equal
the angles in the alternate segments of the circle.
For let a straight line EF touch the circle ABCD at the point B, and from the point B let there be
drawn across, in the circle ABCD, a straight line BD cutting it.
I say that the angles which BD makes with the tangent EF equal the angles in the alternate
segments of the circle, that is, that the angle FBD equals the angle constructed in the segment
BAD, and the angle EBD equals the angle constructed in the segment DCB.
Therefore BA is a diameter of the circle ABCD. Therefore the angle ADB, being an angle in a III.31
semicircle, is right.
Therefore the sum of the remaining angles BAD and ABD equals one right angle. I.32
But the angle ABF is also right, therefore the angle ABF equals the sum of the angles BAD and
ABD.
Subtract the angle ABD from each. Therefore the remaining angle DBF equals the angle BAD
in the alternate segment of the circle.
Next, since ABCD is a quadrilateral in a circle, the sum of its opposite angles equals two right III.22
angles.
But the sum of the angles DBF and DBE also equals two right angles, therefore the sum of the
angles DBF and DBE equals the sum of the angles BAD and BCD, of which the angle BAD was
proved equal to the angle DBF, therefore the remaining angle DBE equals the angle DCB in the
alternate segment DCB of the circle.
Therefore if a straight line touches a circle, and from the point of contact there is drawn across, in the
circle, a straight line cutting the circle, then the angles which it makes with the tangent equal the
angles in the alternate segments of the circle.
Q.E.D.
This proposition is used in the next two propositions and a couple of the propositions in Book IV.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 33
On a given straight line to describe a segment of a circle admitting an angle equal to a given
rectilinear angle.
Let AB be the given straight line, and the angle at C the given rectilinear angle.
Then, since AF equals FB, and FG is common, the two sides AF and FG equal the two
sides BF and FG, and the angle AFG equals the angle BFG, therefore the base AG equals I.4
the base BG.
Therefore the circle described with center G and radius GA passes through B also.
Now, since AD is drawn from A, the end of the diameter AE, at right angles to AE, III.16,Cor.
therefore AD touches the circle ABE.
Since then a straight line AD touches the circle ABE, and from the point of contact at A a
straight line AB has been drawn across in the circle ABE, the angle DAB equals the angle III.32
AEB in the alternate segment of the circle.
But the angle DAB equals the angle at C, therefore the angle at C also equals the angle
AEB.
Therefore on the given straight line AB the segment AEB of a circle has been described
admitting the angle AEB equal to the given angle, the angle at C.
Next let the angle at C be right, and let it be again required to describe on AB a segment of
a circle admitting an angle equal to the right angle at C.
And the angle BAD equals the angle in the segment AEB, for the latter too is itself a right III.31
angle, being an angle in a semicircle.
But the angle BAD also equals the angle at C, therefore the angle AEB also equals the
angle at C.
Therefore again the segment AEB of a circle has been described on AB admitting an angle
equal to the angle at C.
Then, since AF again equals FB, and FG is common, the two sides AF and FG equal the
two sides BF and FG, and the angle AFG equals the angle BFG, therefore the base AG I.4
equals the base BG.
Therefore the circle described with center G and radius GA also passes through B. Let it so
pass, as AEB.
Now, since AD is drawn at right angles to the diameter AE from its end, AD touches the III.16 Cor.
circle AEB.
And AB has been drawn across from the point of contact at A, therefore the angle BAD III.32
equals the angle constructed in the alternate segment AHB of the circle.
Therefore the angle in the segment AHB also equals the angle at C.
Therefore on the given straight line AB the segment AHB of a circle has been described
admitting an angle equal to the angle at C.
Q.E.F.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 34
From a given circle to cut off a segment admitting an angle equal to a given rectilinear angle.
Let ABC be the given circle, and the angle at D the given rectilinear angle.
It is required to cut off from the circle ABC a segment admitting an angle equal to the given
rectilinear angle, the angle at D.
But the angle FBC equals the angle at D, therefore the angle in the segment BAC equals the
angle at D.
Therefore from the given circle ABC the segment BAC has been cut off admitting an angle
equal to the given rectilinear angle, the angle at D.
Q.E.F.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 35
If in a circle two straight lines cut one another, then the rectangle contained by the segments of
the one equals the rectangle contained by the segments of the other.
For in the circle ABCD let the two straight lines AC and BD cut one another at the point E.
Since, then, the straight line AC has been cut into equal parts at G and into unequal parts at E, II.5
the rectangle AE by EC together with the square on EG equals the square on GC.
Add the square on GF. Therefore the rectangle AE by EC plus the sum of the squares on GE
and GF equals the sum of the squares on CG and GF.
But the square on FE equals the sum of the squares on EG and GF, and the square on FC
equals the sum of the squares on CG and GF. Therefore the rectangle AE by EC plus the square I.47
on FE equals the square on FC.
And FC equals FB, therefore the rectangle AE by EC plus the square on EF equals the square
on FB.
For the same reason, also, the rectangle DE by EB plus the square on FE equals the square on
FB.
But the rectangle AE by EC plus the square on FE was also proved equal to the square on FB,
therefore the rectangle AE by EC plus the square on FE equals the rectangle DE by EB plus the
square on FE.
Subtract the square on FE from each. Therefore the remaining rectangle AE by EC equals the
rectangle DE by EB.
Therefore if in a circle two straight lines cut one another, then the rectangle contained by the
segments of the one equals the rectangle contained by the segments of the other.
Q.E.D.
By means of proposition VI.16, the statement, "the rectangle AE by EC equals the rectangle DE by EB,"
may be converted into one about ratios, namely, "AE : EB = DE : EC."
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 36
If a point is taken outside a circle and two straight lines fall from it on the circle, and if one of
them cuts the circle and the other touches it, then the rectangle contained by the whole of the
straight line which cuts the circle and the straight line intercepted on it outside between the
point and the convex circumference equals the square on the tangent.
Let a point D be taken outside the circle ABC, and from D let the two straight lines DCA and
DB fall on the circle ABC. Let DCA cut the circle ABC, and let BD touch it.
Then DCA is either through the center or not through the center.
First let it be through the center, and let F be the center of the III.18
circle ABC. Join FB. Therefore the angle FBD is right.
And the sum of the squares on FB and BD equals the square on FD, therefore the rectangle AD I.47
by DC plus the square on FB equals the sum of the squares on FB and BD.
Subtract the square on FB from each. Therefore the remaining rectangle AD by DC equals the
square on the tangent DB.
Again, let DCA not be through the center of the circle ABC.
Take the center E, and draw EF from E perpendicular to AC. III.1
Join EB, EC, and ED.
Now, since the straight line AC has been bisected at the point F, and CD is added to it, the II.6
rectangle AD by DC plus the square on FC equals the square on FD.
Add the square on FE to each. Therefore the rectangle AD by DC plus the sum of the squares
on CF and FE equals the sum of the squares on FD and FE.
But the square on EC equals the sum of the squares on CF and FE, for the angle EFC is right,
and the square on ED equals the sum of the squares on DF and FE, therefore the rectangle AD I.47
by DC plus the square on EC equals the square on ED.
And EC equals EB, therefore the rectangle AD by DC plus the square on EB equals the square
on ED.
But the sum of the squares on EB and BD equals the square on ED, for the angle EBD is right,
therefore the rectangle AD by DC plus the square on EB equals the sum of the squares on EB I.47
and BD.
Subtract the square on EB from each. Therefore the remaining rectangle AD by DC equals the
square on DB.
Therefore if a point is taken outside a circle and two straight lines fall from it on the circle, and if one
of them cuts the circle and the other touches it, then the rectangle contained by the whole of the
straight line which cuts the circle and the straight line intercepted on it outside between the point and
the convex circumference equals the square on the tangent.
Q.E.D.
Select topic
Book III introduction
© 1996
D.E.Joyce
Clark University
Proposition 37
If a point is taken outside a circle and from the point there fall on the circle two straight lines,
if one of them cuts the circle, and the other falls on it, and if further the rectangle contained by
the whole of the straight line which cuts the circle and the straight line intercepted on it outside
between the point and the convex circumference equals the square on the straight line which
falls on the circle, then the straight line which falls on it touches the circle.
Let a point D be taken outside the circle ABC, from D let the two straight
lines DCA and DB fall on the circle ACB, let DCA cut the circle and DB
fall on it, and let the rectangle AD by DC equal the square on DB.
And FE equals FB, therefore the two sides DE and EF equal the two
sides DB and BF, and FD is the common base of the triangles, therefore I.8
the angle DEF equals the angle DBF.
But the angle DEF is right, therefore the angle DBF is also right.
Similarly this can be proved to be the case even if the center is on AC.
Therefore if a point is taken outside a circle and from the point there fall on the circle two straight
lines, if one of them cuts the circle, and the other falls on it, and if further the rectangle contained by
the whole of the straight line which cuts the circle and the straight line intercepted on it outside
between the point and the convex circumference equals the square on the straight line which falls on
the circle, then the straight line which falls on it touches the circle.
Q.E.D.
© 1996
D.E.Joyce
Clark University
Table of contents
● Definitions (7)
● Propositions (16)
● Guide to Book IV
● Logical structure of Book IV
Definitions
Definition 1.
A rectilinear figure is said to be inscribed in a rectilinear figure when the respective angles of the
inscribed figure lie on the respective sides of that in which it is inscribed.
Definition 2.
Similarly a figure is said to be circumscribed about a figure when the respective sides of the
circumscribed figure pass through the respective angles of that about which it is circumscribed.
Definition 3.
A rectilinear figure is said to be inscribed in a circle when each angle of the inscribed figure lies
on the circumference of the circle.
Definition 4.
A rectilinear figure is said to be circumscribed about a circle when each side of the
circumscribed figure touches the circumference of the circle.
Definition 5.
Similarly a circle is said to be inscribed in a figure when the circumference of the circle touches
each side of the figure in which it is inscribed.
Definition 6.
A circle is said to be circumscribed about a figure when the circumference of the circle passes
through each angle of the figure about which it is circumscribed.
Definition 7.
A straight line is said to be fitted into a circle when its ends are on the circumference of the
circle.
Propositions
Proposition 1.
To fit into a given circle a straight line equal to a given straight line which is not greater than the
diameter of the circle.
Proposition 2.
To inscribe in a given circle a triangle equiangular with a given triangle.
Proposition 3.
To circumscribe about a given circle a triangle equiangular with a given triangle.
Proposition 4.
To inscribe a circle in a given triangle.
Proposition 5.
To circumscribe a circle about a given triangle.
Corollary. When the center of the circle falls within the triangle, the triangle is acute-angled;
when the center falls on a side, the triangle is right-angled; and when the center of the circle falls
outside the triangle, the triangle is obtuse-angled.
Proposition 6.
To inscribe a square in a given circle.
Proposition 7.
To circumscribe a square about a given circle.
Proposition 8.
To inscribe a circle in a given square.
Proposition 9.
To circumscribe a circle about a given square.
Proposition 10.
To construct an isosceles triangle having each of the angles at the base double the remaining one.
Proposition 11.
To inscribe an equilateral and equiangular pentagon in a given circle.
Proposition 12.
To circumscribe an equilateral and equiangular pentagon about a given circle.
Proposition 13.
To inscribe a circle in a given equilateral and equiangular pentagon.
Proposition 14.
To circumscribe a circle about a given equilateral and equiangular pentagon.
Proposition 15.
To inscribe an equilateral and equiangular hexagon in a given circle.
Corollary. The side of the hexagon equals the radius of the circle.
And, in like manner as in the case of the pentagon, if through the points of division on the circle
we draw tangents to the circle, there will be circumscribed about the circle an equilateral and
equiangular hexagon in conformity with what was explained in the case of the pentagon.
And further by means similar to those explained in the case of the pentagon we can both inscribe
a circle in a given hexagon and circumscribe one about it.
Proposition 16.
Corollary. And, in like manner as in the case of the pentagon, if through the points of division on
the circle we draw tangents to the circle, there will be circumscribed about the circle a fifteen-
angled figure which is equilateral and equiangular.
And further, by proofs similar to those in the case of the pentagon, we can both inscribe a circle
in the given fifteen-angled figure and circumscribe one about it.
Guide to Book IV
All but two of the propositions in this book are constructions to inscribe or circumscribe figures.
Inscribe Inscribe
Circumscribe figure Circumscribe circle
Figure figure in circle in
about circle about figure
circle figure
Regular
IV.11 IV.12 IV.13 IV.14
pentagon
Regular
IV.15 IV.15,Cor IV.15,Cor IV.15,Cor
hexagon
Regular 15-
IV.16 IV.16,Cor IV.16,Cor IV.16,Cor
gon
There are only two other propositions. Proposition IV.1 is a basic construction to fit a line in a circle,
and proposition IV.10 constructs a particular triangle needed in the construction of a regular pentagon.
The proofs of the propositions in Book IV rely heavily on the propositions in Books I and III. Only one
proposition from Book II is used and that is the construction in II.11 used in proposition IV.10 to
construct a particular triangle needed in the construction of a regular pentagon.
Most of the propositions of Book IV are logically independent of each other. There is a short chain of
1, 5 10
2, 10 11
11 12
1, 2, 11 16
© 1996
D.E.Joyce
Clark University
Definitions
Def. 1. A rectilinear figure is said to be inscribed in a rectilinear figure when the respective
angles of the inscribed figure lie on the respective sides of that in which it is inscribed.
Def. 2. Similarly a figure is said to be circumscribed about a figure when the respective sides of
the circumscribed figure pass through the respective angles of that about which it is
circumscribed.
Def. 3. A rectilinear figure is said to be inscribed in a circle when each angle of the inscribed
figure lies on the circumference of the circle.
Def. 4. A rectilinear figure is said to be circumscribed about a circle when each side of the
circumscribed figure touches the circumference of the circle.
Def. 5. Similarly a circle is said to be inscribed in a figure when the circumference of the circle
touches each side of the figure in which it is inscribed.
Def. 6. A circle is said to be circumscribed about a figure when the circumference of the circle
passes through each angle of the figure about which it is circumscribed.
Def. 7. A straight line is said to be fitted into a circle when its ends are on the circumference of
the circle.
The first figure shows a smaller quadrilateral inscribed in a larger quadrilateral, and the larger one is
circumscribed about the smaller one. The second figure shows a quadrilateral inscribed in a circle, and
the circle is circumscribed about the quadrilateral. The third figure shows a circle inscribed in a
quadrilateral, and the quadrilateral is circumscribed about the circle. Note also that in the second figure,
each side of the quadrilateral is fitted into the circle.
© 1996
D.E.Joyce
Clark University
Proposition 1
To fit a straight line into a given circle equal to a given straight line which is not greater than
the diameter of the circle.
Let ABC be the given circle, and D the given straight line not greater than the diameter of
the circle.
But, if BC is greater than D, make CE equal to D, describe the circle EAF with center C and I.3
radius CE, and join CA.
Then, since the point C is the center of the circle EAF, CA equals CE.
Therefore CA has been fitted into the given circle ABC equal to the given straight line D. IV.Def.7
Q.E.F.
The hypothesis that the line to be fitted into the circle is no longer than the diameter of the circle is
certainly necessary, but Euclid did not show it was sufficient. That is sufficient to conclude the two
circles actually meet at a point A is never demonstrated. This logical gap has appeared before in the
Elements, for instance in Propositions I.1 and I.22.
This proposition is used in the proofs of IV.10, IV.16, and occasionally in Books X, XI, and XII.
Book IV introduction
© 1996
D.E.Joyce
Clark University
Proposition 2
It is required to inscribe a triangle equiangular with the triangle DEF in the circle ABC.
Then, since a straight line AH touches the circle ABC, and from the point of contact at A the
straight line AC is drawn across in the circle, therefore the angle HAC equals the angle ABC III.32
in the alternate segment of the circle.
But the angle HAC equals the angle DEF, therefore the angle ABC also equals the angle
DEF.
For the same reason the angle ACB also equals the angle DFE, therefore the remaining I.32
angle BAC also equals the remaining angle EDF.
Therefore a triangle equiangular with the given triangle has been inscribed in the given IV.Def.2
circle.
Q.E.F.
Book IV introduction
Proposition 3
It is required to circumscribe a triangle equiangular with the triangle DEF about the circle
ABC.
Now, since LM, MN, and NL touch the circle ABC at the points A, B, and C, and KA, KB,
and KC have been joined from the center K to the points A, B, and C, therefore the angles at III.18
the points A, B, and C are right.
And, since the four angles of the quadrilateral AMBK equal four right angles, inasmuch as
AMBK is in fact divisible into two triangles, and the angles KAM and KBM are right,
therefore the sum of the remaining angles AKB and AMB equals two right angles.
But the sum of the angles DEG and DEF also equals two right angles, therefore the sum of
the angles AKB and AMB equals the sum of the angles DEG and DEF, of which the angle I.13
AKB equals the angle DEG, therefore the remaining angle AMB equals the remaining angle
DEF.
Similarly it can be proved that the angle LNB also equals the angle DFE, therefore the I.32
remaining angle MLN equals the angle EDF.
Therefore the triangle LMN is equiangular with the triangle DEF, and it has been IV.Def.4
circumscribed about the circle ABC.
Therefore a triangle equiangular with the given triangle has been circumscribed about a
given circle.
Q.E.F.
This proposition is not used elsewhere in the Elements, but is included as a mate to the previous
proposition in which a triangle is inscribed inside rather than circumscribed outside a given circle.
Book IV introduction
© 1996
D.E.Joyce
Clark University
Proposition 4
Bisect the angles ABC and ACB by the straight lines I.9
BD and CD, and let these meet one another at the
point D. Draw DE, DF, and DG from D perpendicular
I.12
to the straight lines AB, BC, and CA.
Now, since the angle ABD equals the angle CBD, and the right angle BED also equals the
right angle BFD, EBD and FBD are two triangles having two angles equal to two angles and
one side equal to one side, namely that opposite one of the equal angles, which is BD I.26
common to the triangles, therefore they will also have the remaining sides equal to the
remaining sides, therefore DE equals DF.
Therefore the three straight lines DE, DF, and DG equal one another. Therefore the circle
described with center D and radius one of the straight lines DE, DF, or DG also passes
through the remaining points and touches the straight lines AB, BC, and CA, because the
angles at the points E, F, and G are right.
For, if it cuts them, the straight line drawn at right angles to the diameter of the circle from
its end will be found to fall within the circle, which was proved absurd, therefore the circle III.16
described with center D and radius one of the straight lines DE, DF, or DG does not cut the
straight lines AB, BC, and CA Therefore it touches them, and is the circle inscribed in the IV.Def.5
triangle ABC.
Therefore the circle EFG has been inscribed in the given triangle ABC.
Q.E.F.
It is easy to supply the missing argument that the angle bisectors BD and CD do meet.
This circle inscribed in a triangle has come to be known as the incircle of the triangle, its center the
incenter of the triangle, and its radius the inradius of the triangle.
The incircle is a circle tangent to the three lines
AB, BC, and AC. If these three lines are
extended, then there are three other circles also
tangent to them, but outside the triangle. They
are called the excircles.
Heron's formula
Heron of Alexandria (first century C.E.) was an important Greek mathematician who wrote, among
other things, a commentary on the Elements which is lost now but was known to Proclus and an-Nairizi.
In Heron's Metrica, which was rediscovered in 1896, there appears a proof of what is called Heron's
formula. It states that the area of a triangle is the square root of s(s-a)(s-b)(s-c) where a = BC, b = AC,
and c = AB, the sides of the triangle, and s is the semiperimeter (a + b + c)/2. Archimedes may have
known this formula, but but we don't have his proof. Heath gives Heron's complete proof, but here we'll
just look at the first part that involves the incircle.
Let D be the incenter of the triangle ABC, and let DE, DF, and
DG be perpendicular lines drawn to the sides as in Euclid's proof.
These three lines are radii of the incircle, and therefore have
length r, the inradius. The triangle ABD has base AB and height
r, so its area is r AB/2. Likewise, the area of triangle BCD is r
BC/2, and the area of triangle CAD is r CA/2. Adding these
together we find the area of triangle ABC is r (AB + BC + CA)/2.
Therefore we have
Area(ABC) = rs
From the other excircles we get two more equations. We then have
Book IV introduction
© 1996
D.E.Joyce
Clark University
Proposition 5
Bisect the straight lines AB and AC at the points D and E. Draw DF and EF from the points
I.10
D and E at right angles to AB and AC. They will then meet within the triangle ABC, or on I.11
the straight line BC, or outside BC.
Therefore the circle described with center F and radius one of the straight lines FA, FB, or
FC also passes through the remaining points, and the circle is circumscribed about the IV.Def.6
triangle ABC.
Similarly we can prove that CF also equals AF, so that BF also equals FC. Therefore the
circle described with center F and radius one of the straight lines FA, FB, or FC also passes IV.Def.6
through the remaining points, and is circumscribed about the triangle ABC.
[Corollary]
And it is manifest that when the center of the circle falls within the triangle, the angle BAC,
being in a segment greater than the semicircle, is less than a right angle, when the center falls
on the straight line BC, the angle BAC, being in a semicircle, is right, and when the center of III.31
the circle falls outside the triangle, the angle BAC, being in a segment less than the semicircle,
is greater than a right angle.
As noted by Simson and others, Euclid does not justify the intersection of the perpendicular bisectors
The note following the proposition is not actually called a corollary in the Greek text. It is just a remark
following the proposition.
Circumcircles
This circle drawn about a triangle is called, naturally enough, the circumcircle of the triangle, its center
the circumcenter of the triangle, and its radius the circumradius. Much has been discovered about the
theory of incircles and circumcircles since Euclid.
This relation is easy to derive from the figure. Angle AFC is twice
the angle at B [III.20], but it is also twice angle AFE since the
triangles AFE and CFE are congruent. Therefore angle AFE equals
the angle at B. Then the sine of B can be found in the right triangle
AFE as the ratio of the side AE opposite angle AFE to the
hypotenuse AF. Since AE is half of AC, it follows that sin B = AC/
(2R) which yields one of the three equations for the law of sines.
There is also an equation relating the circumradius R, the inradius r, and the three exradii rA, rB, and rC:
4R = rA + rB + rC – r,
and a number of other interesting results about circumcircles, incircles, and other constructions based on
an arbitrary triangle.
Book IV introduction
© 1996
D.E.Joyce
Clark University
Proposition 6
Draw two diameters AC and BD of the circle ABCD at right angles to one another, and join AB, III.1
BC, CD, and DA. I.11
Then, since BE equals ED, for E is the center, and EA is common I.4
and at right angles, therefore the base AB equals the base AD.
For the same reason each of the straight lines BC and CD also
equals each of the straight lines AB and AD. Therefore the
quadrilateral ABCD is equilateral.
For, since the straight line BD is a diameter of the circle ABCD, therefore BAD is a semicircle, III.31
therefore the angle BAD is right.
For the same reason each of the angles ABC, BCD, and CDA is also right. Therefore the
quadrilateral ABCD is right-angled.
But it was also proved equilateral, therefore it is a square, and it has been inscribed in the circle
ABCD.
Therefore the square ABCD has been inscribed in the given circle.
Q.E.F.
This construction is used in a few propositions of Book XII, the first being XII.2.
Book IV introduction
© 1996
D.E.Joyce
Clark University
Proposition 7
For the same reason the angles at the points B, C, and D are
also right.
Now, since the angle AEB is right, and the angle EBG is also right, therefore GH is parallel I.28
to AC.
For the same reason AC is also parallel to FK, so that GH is also parallel to FK. I.30
Similarly we can prove that each of the straight lines GF and HK is parallel to BED.
Therefore GK, GC, AK, FB, and BK are parallelograms, therefore GF equals HK, and GH I.34
equals FK.
And, since AC equals BD, and AC also equals each of the straight lines GH and FK, and
BD equals each of the straight lines GF and HK, therefore the quadrilateral FGHK is I.34
equilateral.
For, since GBEA is a parallelogram, and the angle AEB is right, therefore the angle AGB is I.34
also right.
Similarly we can prove that the angles at H, K, and F are also right.
But it was also proved equilateral, therefore it is a square, and it has been circumscribed
about the circle ABCD.
Book IV introduction
© 1996
D.E.Joyce
Clark University
Proposition 8
Bisect the straight lines AD and AB at the points E and F respectively. Draw EH through E
I.10
parallel to either AB or CD, and draw FK through F parallel to either AD or BC. Therefore each I.31
of the figures AK, KB, AH, HD, AG, GC, BG, and GD is a parallelogram, and their opposite I.34
sides are evidently equal.
Therefore the circle described with center G and radius one of the straight lines GE, GF, GH,
or GK also passes through the remaining points.
And it touches the straight lines AB, BC, CD, and DA, because the angles at E, F, H, and K are
right.
For, if the circle cuts AB, BC, CD, or DA, the straight line drawn at right angles to the diameter
of the circle from its end will fall within the circle, which was proved absurd. Therefore the III.16
circle described with center G and radius one of the straight lines GE, GF, GH, or GK does not
cut the straight lines AB, BC, CD, and DA.
Therefore it touches them, and has been inscribed in the square ABCD.
Q.E.F.
This is a straightforward proposition, one of four in the sequence IV.6 through IV.9 about circles and
squares.
Book IV introduction
© 1996
D.E.Joyce
Clark University
Proposition 9
Similarly we can prove that each of the angles ABC, BCD, and
CDA is bisected by the straight lines AC and DB.
Now, since the angle DAB equals the angle ABC, and the angle EAB is half of the angle DAB, and
the angle EBA half of the angle ABC, therefore the angle EAB also equals the angle EBA, so that I.6
the side EA also equals EB.
Similarly we can prove that each of the straight lines EA and EB equals each of the straight lines
EC and ED.
Therefore the four straight lines EA, EB, EC, and ED equal one another.
Therefore the circle described with center E and radius one of the straight lines EA, EB, EC, or
ED also passes through the remaining points, and it is circumscribed about the square ABCD.
Q.E.F.
This is a straightforward proposition, one of four in the sequence IV.6 through IV.9 about circles and
squares.
Book IV introduction
© 1996
D.E.Joyce
Clark University
Proposition 10
To construct an isosceles triangle having each of the angles at the base double the remaining
one.
Set out any straight line AB, and cut it at the point C so that the rectangle AB by BC equals the
square on CA. Describe the circle BDE with center A and radius AB. Fit in the circle BDE the II.11
straight line BD equal to the straight line AC which is not greater than the diameter of the circle IV.1
BDE.
Join AD and DC, and circumscribe the circle ACD about the triangle ACD. IV.5
Since, then, BD touches it, and DC is drawn across from the point of contact at D, therefore the III.32
angle BDC equals the angle DAC in the alternate segment of the circle.
Since, then, the angle BDC equals the angle DAC, add the angle CDA to each, therefore the
whole angle BDA equals the sum of the two angles CDA and DAC.
But the exterior angle BCD equals the sum of the angles CDA and DAC, therefore the angle I.32
BDA also equals the angle BCD.
But the angle BDA equals the angle CBD, since the side AD also equals AB, so that the angle I.5
DBA also equals the angle BCD.
Therefore the three angles BDA, DBA, and BCD equal one another.
And, since the angle DBC equals the angle BCD, the side BD also equals the side DC. I.6
But BD equals CA by hypothesis, therefore CA also equals CD, so that the angle CDA also I.5
equals the angle DAC. Therefore the sum of the angles CDA and DAC is double the angle DAC.
And the angle BCD equals the sum of the angles CDA and DAC, therefore the angle BCD is
also double the angle CAD.
But the angle BCD equals each of the angles BDA and DBA, therefore each of the angles BDA
and DBA is also double the angle DAB.
Therefore the isosceles triangle ABD has been constructed having each of the angles at the base
DB double the remaining one.
Q.E.F.
The goal of the proposition is to construct a 36°-72°-72° isosceles triangle ABD. It's actually constructed
on a given side AB. The base will equal the larger part of AB when AB is cut at a point C so that
AB BC = AC2. The constuction for that cut was given in proposition II.11. The difficulty of the proof is
showing that this construction results in the desired triangle. Cutting AB at that point is also called
cutting the line "in extreme and mean ratio," see VI.Def.3 for the definition of "extreme and mean ratio,"
and see proposition VI.30 for details.
At this point Euclid has shown that one of the two angles at D,
namely angle BDC, equals the angle A. When he shows that the
other, namely angle CDA also equals angle A, then since the
triangle ABD is isosceles, he will have shown each of the base
angles of triangle ABD is twice the vertex angle A, and the proof
will be complete.
The rest is relatively easy. First, the small triangle BCD is isosceles, a fact that can be seen from the
following equation about angles:
Therefore, the sides CD and BD are equal, but from the original construction, BD = CA. Hence, the
triangle ADC is also isosceles, so the two angles CDA and A are equal, as needed.
Comments
Euclid could have split the statement and the proof of this proposition into two. The first part would state
that if an isosceles triangle has its base equal to a segment of its side so that square on the base equals
the rectangle contained by the side and the remaining segment of the side, then each base angle of the
triangle is twice the vertex angle. Most of the proof of this proposition IV.10 is actually a proof of this
first part. The other part would be the constuction.
There is a converse of this proposition, one the Euclid did not state. Namely, if an isosceles triangle has
each base angle equal to twice the vertex angle, then the base is equal to a segment of its side so that
square on the base equals the rectangle contained by the side and the remaining segment of the side. In
other words, 36°-72°-72° isosceles triangles are characterized by this property.
The triangle ABD constructed in this proposition is one of ten sectors of a regular decagon (10-gon).
Thus, it is one short step from this proposition to the construction of a regular decagon inscribed in a
circle. If alternate vertices of a regular decagon are connected, then a regular pentagon is formed which
is inscribed in the circle. It is unclear why Euclid did not use such a construction rather than the one he
chose in the next proposition
It was probably Euclid who made a concerted effort to include as many propositions that he could in the
first four books that did not rely on proportions. The theory of similar triangles is not broached until
Book VI which depends on the theory of proportion in Book V. The clever proof that Euclid gave to this
proposition does not depend on similar triangles, and so it could be placed here in Book IV. There is,
however, a simpler proof that does depend on similar triangles.
As Euclid does, begin by cutting a straight line AB at the point C so that the rectangle AB by BC equals
the square on CA (II.11). Otherwise said, the straight line AB has been cut in extreme and mean ratio at
C so that the proportion AB:AC = AC:BC holds. (See VI.Def.3, VI.17, and VI.30.) Next, construct an
isosceles triangle with one side AB, a second side AD equal to side AB, and the base equal BD equal to
AC (I.22). Then we have the proportion AD:BD = BD:BC. Therefore, two triangles ADB and DBC have
one angle equal to one angle (angle D of the first triangle equals angle B of the second) and the sides
about the equal angles proportional. Therefore, by VI.6, the triangles are equiangular. It easily follows
that both triangles have their base angles each equal to twice their vertex angles.
This construction was designed to be used in the next proposition which inscribes a regular pentagon in
a circle.
Book IV introduction
© 1996, 2002
D.E.Joyce
Clark University
Proposition 11
Set out the isosceles triangle FGH having each of the angles at G and H double the angle at F. IV.10
Inscribe in the circle ABCDE the triangle ACD equiangular with the triangle FGH, so that the
angles CAD, ACD, and CDA equal the angles at F, G, and H respectively. Therefore each of
IV.2
the angles ACD and CDA is also double the angle CAD.
But equal angles stand on equal circumferences, therefore the five circumferences AB, BC, III.26
CD, DE, and EA equal one another.
But straight lines that cut off equal circumferences are equal, therefore the five straight lines III.29
AB, BC, CD, DE, and EA equal one another. Therefore the pentagon ABCDE is equilateral.
For, since the circumference AB equals the circumference DE, add BCD to each, therefore the
whole circumference ABCD equals the whole circumference EDCB.
And the angle AED stands on the circumference ABCD, and the angle BAE on the III.27
circumference EDCB, therefore the angle BAE also equals the angle AED.
For the same reason each of the angles ABC, BCD, and CDE also equals each of the angles
BAE and AED, therefore the pentagon ABCDE is equiangular.
But it was also proved equilateral, therefore an equilateral and equiangular pentagon has been
inscribed in the given circle.
Q.E.F.
Richmond's construction
The easiest way to verify that Richmond's construction works is by means of trigonometry.
It is evident that there are many lines parallel to the base CD of the triangle, namely BE and LG, as well
as innumerable ones in the smaller pentagrams. That means that there will be many 36°-72°-72°
triangles besides the large one ACD. The next smaller one is ALG, then FKH, then many smaller ones.
Also, each of these various sized 36°-72°-72° triangles is congruent many others in the diagram. For
instance, triangles ALG and EAK are congruent.
There are also a series of obtuse 36°-36°-108° isosceles triangles of varying sizes.
All these parallel lines and similar triangles yield numerous relationships among the various diagonals
and sides of the pentagons. Some of these relationships are additive equations:
d1= s1+ d2
s1= d2+ s2
d2= s2+ d3
s2= d3+ s3
and so forth.
Other relationships are based on the property of 36°-72°-72° triangles used in their construction in
IV.10, namely that the square of the base of such a triangle equals the product of a side and the
difference between the side and the base. In terms of the diagonals and sides of the pentagons, this gives
the equations:
d1d2= s12
s1s2= d22
d2d3= s22
s2s3= d32
and so forth.
After ratios are proportions are developed in Book V and Book VI, we can add the following continued
proportion to the list of relationships:
See the Guide to proposition X.2 which shows that diagonal and side of a regular pentagon are
incommensurable. In more modern terms we would say that their ratio, which is called the "golden
ratio," is an irrational number.
This construction is used in the next proposition to circumscribe a regular pentagon around a circle and
later in IV.16 to construct a regular 15-gon. It is also used in XIII.16 for the construction of a regular
icosahedron (a 20-sided polyhedron each of whose faces is an equilateral triangle). Surprizingly, it is not
used in XIII.17 for construct a regular dodecahedron (a 12-sided polyhedron each of whose faces is a
regular pentagon); the regular pentagons needed for it are constructed in space directly without the help
of this proposition.
Book IV introduction
Proposition 12
Let A, B, C, D, and E be conceived to be the angular points of the inscribed pentagon, so IV.11
that the circumferences AB, BC, CD, DE, and EA are equal. Draw GH, HK, KL, LM, and III.16,
MG through A, B, C, D, and E touching the circle. Take the center F of the circle ABCDE, Cor
and join FB, FK, FC, FL, and FD. III.1
For the same reason the angles at the points B and D are also
right.
And, since the angle FCK is right, therefore the square on I.47
FK equals the sum of the squares on FC and CK.
For the same reason the square on FK also equals the sum of the squares on FB and BK, so
that the sum of the squares on FC and CK equals the sum of the squares on FB and BK, of I.47
which the square on FC equals the square on FB, therefore the remaining square on CK
equals the square on BK.
And, since FB equals FC, and FK is common, the two sides BF and FK equal the two sides
CF and FK, and the base BK equals the base CK, therefore the angle BFK equals the angle I.8
KFC, and the angle BKF equals the angle FKC. Therefore the angle BFC is double the
angle KFC, and the angle BKC double the angle FKC.
For the same reason the angle CFD is also double the angle CFL, and the angle DLC
double the angle FLC.
Now, since the circumference BC equals CD, the angle BFC also equals the angle CFD. III.27
And the angle BFC is double the angle KFC, and the angle DFC double the angle LFC,
therefore the angle KFC also equals the angle LFC.
But the angle FCK also equals the angle FCL, therefore FKC and FLC are two triangles
having two angles equal to two angles and one side equal to one side, namely FC which is
common to them, therefore they will also have the remaining sides equal to the remaining I.26
sides, and the remaining angle to the remaining angle, therefore the straight line KC equals
CL, and the angle FKC equals the angle FLC.
For the same reason it can be proved that HK is also double BK.
Similarly each of the straight lines HG, GM, and ML can also be proved equal to each of
the straight lines HK and KL, therefore the pentagon GHKLM is equilateral.
For, since the angle FKC equals the angle FLC, and the angle HKL was proved double the
angle FKC, and the angle KLM double the angle FLC, therefore the angle HKL also equals
the angle KLM.
Similarly each of the angles KHG, HGM, and GML can also be proved equal to each of the
angles HKL and KLM. Therefore the five angles GHK, HKL, KLM, LMG, and MGH equal
one another.
And it was also proved equilateral, and it has been circumscribed about the circle ABCDE.
Q.E.F.
This construction depends on the last. First, inscribe a regular pentagon in the circle, then take tangents
to the circle at the five vertices of the inscribed pentagon. The result will be a circumscribed pentagon.
This method generally works to create a regular circumscribed n-gon given a regular inscribed n-gon.
Conversely, if you have a regular circumscribed n-gon, then you can connect the points of tangency in
sequence to get a regular inscribed n-gon.
Book IV introduction
© 1996
D.E.Joyce
Clark University
Proposition 13
Bisect the angles BCD and CDE by the straight lines CF and DF respectively. Join the straight I.9
lines FB, FA, and FE from the point F at which the straight lines CF and DF meet one another.
And, since the angle CDE is double the angle CDF, and
the angle CDE equals the angle ABC, while the angle
CDF equals the angle CBF, therefore the angle CBA is
also double the angle CBF. Therefore the angle ABF
equals the angle FBC. Therefore the angle ABC is
bisected by the straight line BF.
Similarly it can be proved that the angles BAE and AED are also bisected by the straight lines
FA and FE respectively.
Now draw FG, FH, FK, FL, and FM from the point F perpendicular to the straight lines AB, I.12
BC, CD, DE, and EA.
Then, since the angle HCF equals the angle KCF, and the right angle FHC also equals the
angle FKC, FHC and FKC are two triangles having two angles equal to two angles and one
side equal to one side, namely FC which is common to them and opposite one of the equal I.26
angles, therefore they also have the remaining sides equal to the remaining sides. Therefore the
perpendicular FH equals the perpendicular FK.
Similarly it can be proved that each of the straight lines FL, FM, and FG also equals each of
the straight lines FH and FK, therefore the five straight lines FG, FH, FK, FL, and FM equal
one another.
Therefore the circle described with center F and radius one of the straight lines FG, FH, FK,
FL, or FM also passes through the remaining points, and it touches the straight lines AB, BC,
CD, DE, and EA, because the angles at the points G, H, K, L, and M are right.
For, if it does not touch them. but cuts them, it will result that the straight line drawn at right
angles to the diameter of the circle from its end falls within the circle, which was proved III.16
absurd.
Therefore the circle described with center F and radius one of the straight lines FG, FH, FK,
FL, or FM does not cut the straight lines AB, BC, CD, DE, and EA. Therefore it touches them.
Therefore a circle has been inscribed in the given equilateral and equiangular pentagon.
Q.E.F.
The method given here to inscribe a circle in a regular pentagon works in general to inscribe a circle in a
regular n-gon. Simply bisect two of the angles of the n-gon to find the center of the circle. Then draw a
perpendicular to one of the sides. The foot of the perpendicular gives a point on the circumference of the
circle.
Book IV introduction
© 1996
D.E.Joyce
Clark University
Proposition 14
Bisect the angles BCD and CDE by the straight lines CF and DF respectively. Join the straight I.9
lines FB, FA, and FE from the point F at which the straight lines meet to the points B, A, and E.
Now, since the angle BCD equals the angle CDE, and the angle
FCD is half of the angle BCD, and the angle CDF half of the angle I.6
CDE, therefore the angle FCD also equals the angle CDF, so that
the side FC also equals the side FD.
Similarly it can be proved that each of the straight lines FB, FA, and FE also equals each of the
straight lines FC and FD. Therefore the five straight lines FA, FB, FC, FD, and FE equal one
another.
Therefore the circle described with center F and radius one of the straight lines FA, FB, FC, FD,
or FE also passes through the remaining points, and is circumscribed.
Therefore a circle has been circumscribed about the given equilateral and equiangular pentagon.
Q.E.F.
Book IV introduction
© 1996
D.E.Joyce
Clark University
Proposition 15
Draw the diameter AD of the circle ABCDEF. Take the center G of the circle. Describe the
circle EGCH with center D and radius DG. Join EG and CG and carry them through to the III.1
points B and F. Join AB, BC, CD, DE, EF, and FA.
For, since the point G is the center of the circle ABCDEF, GE equals
GD.
Again, since the point D is the center of the circle GCH, DE equals
DG.
And the sum of the three angles of the triangle equals two right I.32
angles, therefore the angle EGD is one-third of two right angles.
Similarly, the angle DGC can also be proved to be one third of two
right angles.
And, since the straight line CG standing on EB makes the sum of the adjacent angles EGC and
CGB equal to two right angles, therefore the remaining angle CGB is also one-third of two I.13
right angles.
Therefore the angles EGD, DGC, and CGB equal one another, so that the angles vertical to I.15
them, the angles BGA, AGF, and FGE, are equal.
Therefore the six angles EGD, DGC, CGB, BGA, AGF, and FGE equal one another.
But equal angles stand on equal circumferences, therefore the six circumferences AB, BC, CD, III.26
DE, EF, and FA equal one another.
And straight lines that cut off equal circumferences are equal, therefore the six straight lines III.29
equal one another. Therefore the hexagon ABCDEF is equilateral.
For, since the circumference FA equals the circumference ED, add the circumference ABCD to
each, therefore the whole FABCD equals the whole EDCBA. And the angle FED stands on the III.27
circumference FABCD, and the angle AFE on the circumference EDCBA, therefore the angle
AFE equals the angle DEF.
Similarly it can be proved that the remaining angles of the hexagon ABCDEF are also severally
equal to each of the angles AFE and FED, therefore the hexagon ABCDEF is equiangular.
But it was also proved equilateral, and it has been inscribed in the circle ABCDEF.
Therefore an equilateral and equiangular hexagon has been inscribed in the given circle.
Q.E.F.
Corollary
From this it is manifest that the side of the hexagon equals the radius of the circle.
And, in like manner as in the case of the pentagon, if through the points of division on the circle we draw
tangents to the circle, there will be circumscribed about the circle an equilateral and equiangular
hexagon in conformity with what was explained in the case of the pentagon.
And further by means similar to those explained in the case of the pentagon we can both inscribe a
circle in a given hexagon and circumscribe one about it.
The corollary is used in several propositions in Book XIII starting with XIII.9.
Book IV introduction
© 1996
D.E.Joyce
Clark University
Proposition 16
It is required to inscribe in the circle ABCD a fifteen-angled figure which shall be both
equilateral and equiangular.
If therefore we join BE and EC and continually fit into the circle ABCD straight lines equal to IV.1
them, a fifteen-angled figure which is both equilateral and equiangular will be inscribed in it.
Q.E.F.
Corollary
And, in like manner as in the case of the pentagon, if through the points of division on the circle we draw
tangents to the circle, there will be circumscribed about the circle a fifteen-angled figure which is
equilateral and equiangular.
And further, by proofs similar to those in the case of the pentagon, we can both inscribe a circle in the
given fifteen-angled figure and circumscribe one about it.
The arc AC is 1/3 of the circle, since A and B are two of the three equally spaced vertices of a regular
triangle. Likewise, the arc AC is 1/5 of the circle, since A and C are two adjacent points of a regular
pentagon. Therefore, the difference of these two arcs, AC – AB, which is the arc BC is 1/3 –1/5 of the
circle, that is 2/15 of the circle. Since E bisects that arc BC, therefore BE and EC are each 1/15 of the
circle. The rest of the regular 15-gon can then easily be constructed.
Now, by the end of Book IV, Euclid has described how to construct many regular polygons. The regular
3-gon, known as the equilateral triangle, was constructed in I.1, while the regular 4-gon, known as the
square, was constructed in I.46. In book IV, regular 5-gons and regular 6-gons have been constructed.
An application of III.30 (which was used in this proposition) can double the number of sides of a regular
polygon, and therefore regular polygons with 8, 10, 12, 16, 20, 24, etc., sides can be constructed. This
proposition shows how to use a regular m-gon and a regular n-gon to produce a regular mn-gon,
provided that m and n are relatively prime numbers. That produced a 15-gon, and from that we can
produce regular polygons with 30, 60, 120, etc., sides. Thus, a regular n-gon can be constructed if the
only prime numbers that divide n are 2, 3, and 5, where 2 can be a repeated factor, but 3 and 5 are not
repeated.
But are there any others? What about regular polygons with 7, 9, 11, 13, 17, 18, 19, etc., sides? Euclid
said nothing about them, but the ancient Greek mathematicians expected that they couldn't be
constructed with only the Euclidean tools of straightedge and compass. There were constructions
involving conic sections (hyperbolas, parabolas, ellipses) to trisect an angle. With such a construction a
9-gon can be made. But methods involving conic sections go beyond Euclidean tools. With the help of
non-algebraic curves, like Archimedes' spiral, an angle can be divided into any number of equal parts,
and with the aid of those curves any n-gon can be constructed. But, again, they go beyond Euclidean
tools.
The problem of constructing other regular polygons with Euclidean tools remained just that, a problem,
for over 2000 years. Finally, Carl Friedrich Gauss (1777-1855) made progress. He described in his
Disquitiones Arithmeticae, a major work on number theory, how to construct a regular 17-gon with
Euclidean tools. Thus, 17 can be added to 3 and 5 as prime numbers that can divide n, but at most once.
k
Furthermore, he showed that any prime number which is of the form 22 + 1 can be included. Such
prime numbers are called Fermat primes. The known Fermat primes are 3 (which is 220 + 1), 5 (which is
221 + 1), 17 (which is 222 + 1), 257 (which is 223 + 1), and 65537 (which is 224 + 1). Thus, 257 and
65537 can be appended to the list 3, 5, 17. It is not known whether there are any more Fermat primes.
Gauss was convinced that the only constructable n-gons were those where n was only divisible by 2 and
the Fermat primes, where the Fermat primes were not repeated. But he had no proof of that, but in 1837
Wantzel did.
Book IV introduction
© 1996, 2002
D.E.Joyce
Clark University
Table of contents
● Definitions (18)
● Propositions (25)
● Guide to Book V
● Logical structure of Book V
Definitions
Definition 1
A magnitude is a part of a magnitude, the less of the greater, when it measures the greater.
Definition 2
The greater is a multiple of the less when it is measured by the less.
Definition 3
A ratio is a sort of relation in respect of size between two magnitudes of the same kind.
Definition 4
Magnitudes are said to have a ratio to one another which can, when multiplied, exceed one
another.
Definition 5
Magnitudes are said to be in the same ratio, the first to the second and the third to the fourth,
when, if any equimultiples whatever are taken of the first and third, and any equimultiples
whatever of the second and fourth, the former equimultiples alike exceed, are alike equal to, or
alike fall short of, the latter equimultiples respectively taken in corresponding order.
Definition 6
Let magnitudes which have the same ratio be called proportional.
Definition 7
When, of the equimultiples, the multiple of the first magnitude exceeds the multiple of the
second, but the multiple of the third does not exceed the multiple of the fourth, then the first is
said to have a greater ratio to the second than the third has to the fourth.
Definition 8
A proportion in three terms is the least possible.
Definition 9
When three magnitudes are proportional, the first is said to have to the third the duplicate ratio of
that which it has to the second.
Definition 10
When four magnitudes are continuously proportional, the first is said to have to the fourth the
triplicate ratio of that which it has to the second, and so on continually, whatever be the
proportion.
Definition 11
Antecedents are said to correspond to antecedents, and consequents to consequents.
Definition 12
Alternate ratio means taking the antecedent in relation to the antecedent and the consequent in
relation to the consequent.
Definition 13
Inverse ratio means taking the consequent as antecedent in relation to the antecedent as
consequent.
Definition 14
A ratio taken jointly means taking the antecedent together with the consequent as one in relation
to the consequent by itself.
Definition 15
A ratio taken separately means taking the excess by which the antecedent exceeds the consequent
in relation to the consequent by itself.
Definition 16
Conversion of a ratio means taking the antecedent in relation to the excess by which the
antecedent exceeds the consequent.
Definition 17
A ratio ex aequali arises when, there being several magnitudes and another set equal to them in
multitude which taken two and two are in the same proportion, the first is to the last among the
first magnitudes as the first is to the last among the second magnitudes. Or, in other words, it
means taking the extreme terms by virtue of the removal of the intermediate terms.
Definition 18
A perturbed proportion arises when, there being three magnitudes and another set equal to them
in multitude, antecedent is to consequent among the first magnitudes as antecedent is to
consequent among the second magnitudes, while, the consequent is to a third among the first
magnitudes as a third is to the antecedent among the second magnitudes.
Propositions
Proposition 1
If any number of magnitudes are each the same multiple of the same number of other
magnitudes, then the sum is that multiple of the sum.
Proposition 2
If a first magnitude is the same multiple of a second that a third is of a fourth, and a fifth also is
the same multiple of the second that a sixth is of the fourth, then the sum of the first and fifth also
is the same multiple of the second that the sum of the third and sixth is of the fourth.
Proposition 3
If a first magnitude is the same multiple of a second that a third is of a fourth, and if
equimultiples are taken of the first and third, then the magnitudes taken also are equimultiples
respectively, the one of the second and the other of the fourth.
Proposition 4
If a first magnitude has to a second the same ratio as a third to a fourth, then any equimultiples
whatever of the first and third also have the same ratio to any equimultiples whatever of the
second and fourth respectively, taken in corresponding order.
Proposition 5
If a magnitude is the same multiple of a magnitude that a subtracted part is of a subtracted part,
then the remainder also is the same multiple of the remainder that the whole is of the whole.
Proposition 6
If two magnitudes are equimultiples of two magnitudes, and any magnitudes subtracted from
them are equimultiples of the same, then the remainders either equal the same or are
equimultiples of them.
Proposition 7
Equal magnitudes have to the same the same ratio; and the same has to equal magnitudes the
same ratio.
Corollary If any magnitudes are proportional, then they are also proportional inversely.
Proposition 8
Of unequal magnitudes, the greater has to the same a greater ratio than the less has; and the same
has to the less a greater ratio than it has to the greater.
Proposition 9
Magnitudes which have the same ratio to the same equal one another; and magnitudes to which
the same has the same ratio are equal.
Proposition 10
Of magnitudes which have a ratio to the same, that which has a greater ratio is greater; and that to
which the same has a greater ratio is less.
Proposition 11
Ratios which are the same with the same ratio are also the same with one another.
Proposition 12
If any number of magnitudes are proportional, then one of the antecedents is to one of the
consequents as the sum of the antecedents is to the sum of the consequents.
Proposition 13
If a first magnitude has to a second the same ratio as a third to a fourth, and the third has to the
fourth a greater ratio than a fifth has to a sixth, then the first also has to the second a greater ratio
Proposition 14
If a first magnitude has to a second the same ratio as a third has to a fourth, and the first is greater
than the third, then the second is also greater than the fourth; if equal, equal; and if less, less.
Proposition 15
Parts have the same ratio as their equimultiples.
Proposition 16
If four magnitudes are proportional, then they are also proportional alternately.
Proposition 17
If magnitudes are proportional taken jointly, then they are also proportional taken separately.
Proposition 18
If magnitudes are proportional taken separately, then they are also proportional taken jointly.
Proposition 19
If a whole is to a whole as a part subtracted is to a part subtracted, then the remainder is also to
the remainder as the whole is to the whole.
Corollary. If magnitudes are proportional taken jointly, then they are also proportional in
conversion.
Proposition 20
If there are three magnitudes, and others equal to them in multitude, which taken two and two are
in the same ratio, and if ex aequali the first is greater than the third, then the fourth is also greater
than the sixth; if equal, equal, and; if less, less.
Proposition 21
If there are three magnitudes, and others equal to them in multitude, which taken two and two
together are in the same ratio, and the proportion of them is perturbed, then, if ex aequali the first
magnitude is greater than the third, then the fourth is also greater than the sixth; if equal, equal;
and if less, less.
Proposition 22
If there are any number of magnitudes whatever, and others equal to them in multitude, which
taken two and two together are in the same ratio, then they are also in the same ratio ex aequali.
Proposition 23
If there are three magnitudes, and others equal to them in multitude, which taken two and two
together are in the same ratio, and the proportion of them be perturbed, then they are also in the
same ratio ex aequali.
Proposition 24
If a first magnitude has to a second the same ratio as a third has to a fourth, and also a fifth has to
the second the same ratio as a sixth to the fourth, then the sum of the first and fifth has to the
second the same ratio as the sum of the third and sixth has to the fourth.
Proposition 25
If four magnitudes are proportional, then the sum of the greatest and the least is greater than the
sum of the remaining two.
for Book V
Book V covers the abstract theory of ratio and proportion. A ratio is an indication of the relative size of
two magnitudes. The propositions in the following book, Book VI, are all geometric and depend on
ratios, so the theory of ratios needs to be developed first. To get a better understanding of what ratios are
in geometry, consider the first proposition VI.1. It states that triangles of the same height are
proportional to their bases, that is to say, one triangle is to another as one base is to the other. (A
proportion is simply an equality of two ratios.) A simple example is when one base is twice the other,
therefore the triangle on that base is also twice the triangle on the other base. This ratio of 2:1 is fairly
easy to comprehend. Indeed, any ratio equal to a ratio of two numbers is easy to comprehend. Given a
proportion that says a ratio of lines equals a ratio of numbers, for instance, A:B = 8:5, we have two
interpretations. One is that there is a shorter line CA = 8C while B = 5C. This interpretation is the
definition of proportion that appears in Book VII. A second interpretation is that 5 A = 8 B. Either
interpretation will do if one of the ratios is a ratio of numbers, and if A:B equals a ratio of numbers that A
and B are commensurable, that is, both are measured by a common measure.
Many straight lines, however, are not commensurable. If A is the side of a square and B its diagonal,
then A and B are not commensurable; the ratio A:B is not the ratio of numbers. This fact seems to have
been discovered by the Pythagoreans, perhaps Hippasus of Metapontum, some time before 400 B.C.E., a
hundred years before Euclid's Elements.
The difficulty is one of foundations: what is an adequate definition of proportion that includes the
incommensurable case? The solution is that in V.Def.5. That definition, and the whole theory of ratio
and proportion in Book V, are attributed to Eudoxus of Cnidus (died. ca. 355 B.C.E.)
The first group of propositions, 1, 2, 3, 5, and 6 only mention multitudes of magnitudes, not ratios. They
each either state, or depend strongly on, a distributivity or an associativity. In the following identities, m
and n refer to numbers (that is, multitudes) while letters near the end of the alphabet refer to magnitudes.
(m + n)x = mx + nx.
m(nx) = (mn)x.
m(x - y) = mx - my.
(m - n)x = mx - nx.
The rest of the propositions develop the theory of ratios and proportions starting with basic properties
and progressively becoming more advanced.
V.4. If w:x = y:z, then for any numbers m and n, mw:mx = ny:nz.
V.7. Substitution of equals in ratios. If x = y, then x:z = y:z and z:x = z:y.
V.8. If x < y, then x:z < y:z but z:x > z:y.
V.9. (A converse to V.7.) If x:z = y:z, then x = y. Also, if z:x = z:y, then x = y.
V.10. (A converse to V.8.) If x:z < y:z, then x < y. But if z:x < z:y, then x > y
V.11. Transitivity of equal ratios. If u:v = w:x and w:x = y:z, then u:v = y:z.
V.12. If x1:y1 = x2:y2 = ... = xn:yn, then each of these ratios also equals the ratio
(x1 + x2 + ... + xn) : (y1 + y2 + ... + yn).
V.13. Substitution of equal ratios in inequalities of ratios. If u:v = w:x and w:x > y:z, then u:v > y:z.
V.17. Proportional taken jointly implies proportional taken separately. If (w + x):x = (y + z):z, then w:
x = y:z.
V.18. Proportional taken separately implies proportional taken jointly. (A converse to V.17.) If w:x = y:z,
then (w + x):x = (y + z):z.
V.20 is just a preliminary proposition to V.22, and V.21 is just a preliminary proposition to V.23.
V.22. Ratios ex aequali. If x1:x2 = y1:y2, x2:x3 = y2:y3, ... , and x :x =y :y , then x1:xn = y1:yn.
n-1 n n-1 n
V.23. Perturbed ratios ex aequali. If u:v = y:z and v:w = x:y, then u:w = x:z.
V.25. If w:x = y:z and w is the greatest of the four magnitudes while z is the least, then w + z > x + y.
© 1996
D.E.Joyce
Clark University
Definition 3
A ratio is a sort of relation in respect of size between two magnitudes of the same kind.
A convenient notation for a ratio of two magnitudes A and B of the same kind is A:B.
No mixed ratios
All of Euclid's ratios are pure ratios of two magnitudes of the same kind, in other words, there are no
mixed ratios in the Elements. A familiar example of a mixed ratio is velocity, the ratio of a distance to a
time, measured in units such as kilometers/hour.
That isn't to say that ratios of different kinds of magnitudes aren't equated. In fact, that's one of the more
important aspects of ratios. For example, the fundamental proposition of Book VI, proposition VI.1, says
that given two triangles of the same height, the ratio of the triangles A:B is the same as the ratio of their
heights Ah:Bh. That says that the ratio of two plane figures equals the ratio of two lines.
Now, a common operation on proportions (equalities of ratios) is that of alternation (see V.Def.12 and
V.16) which in its general form says that if A:B = C:D, then A:C = B:D. In the Elements alternation only
applies when all four quantities are of the same kind. But if alternation is applied to the proportion of
VI.1, then we get A:Ah = B:Bh, the equality of two mixed ratios, ratios of plane figures to lines. This
step, and the acceptance of mixed ratios, which seems to us like a small thing, was not taken until
centuries after Euclid.
A ratio is a pair of magnitudes of the same kind considered as a pair, but soon identified with other
ratios. Definition 3 promises that ratios have sizes, that is, given two ratios A:B and C:D, either the first
ratio is greater, equal, or less than the second ratio. That promise begins to be fulfilled in Definitions V.
Def.5 and V.Def.7 where equality and order of ratios defined. Note that equality and order are defined
for ratios, but they were assumed for numbers and magnitudes.
Since equality and order are defined, their expected properties are proved in propositions, or at least
some of the properties. For example, proposition V.11 states that two ratios that are the same as a third
are the same as each other, a statement analogous to C.N.1 for magnitudes.
Equivalence relations
(Equivalence relations were mentioned before in the guide for the Common Notions. It was mentioned
there that equality of magnitudes of the same kinds is an equivalence relation.)
The process used for defining ratios of magnitudes was something new for Eudoxus and Euclid, but that
process is now commonplace in mathematics to construct new kinds of things. The process starts with
entities x, y, z, etc., that are well understood, such as pairs of magnitudes of the same kind. Then a
relation E on these entities is found which is intended to be equality for them. For ratios, that is given in
V.Def.5. Right now, let xEy denote that x is related to y by the relation E. Next, it is verified that the
relation E is an equivalence relation, that is, a reflexive, symmetric, and transitive relation.
A relation E is reflexive if for any x it is the case that xEx, that is, anything is related to itself by E. A
relation is E is symmetric if whenever xEy, then yEx. And it is transitive if whenever xEy and yEz, then
xEz.
Once E is known to be an equivalence relation, new entities are conceived which are named by the old
entities x, y, z, etc., but the new entities are taken to be equal, x = y, when their names are equivalent
under the relation E, that is, xEy. Proportion as an equivalence relation is discussed in the Guide to
definition V.Def.5.
There are several operations on ratios and proportions defined soon. For instance, V.Def.9 defines
duplicate ratios, under certain assumptions, which may be thought of as the squares of ratios. See also
definitions V.Def.12 through V.18. But ratios are neither numbers nor magnitudes, and the usual
operations of addition, subtraction, multiplication, and division that apply to numbers don't apply to
ratios.
Numbers can be added and subtracted, and so can magnitudes of the same kind, but ratios cannot. Take
for example a ratio A:B of plane figures and a ratio C:D of angles. What could be meant by their sum (A:
B) + (C:D)? One obvious approach is to treat ratios as quotients. That suggests A/B + C/D = (AD + BC)/
BD, but a product of a plane figure and an angle, such as AD, has no meaning, so the obvious approach
has obvious difficulties.
Multiplication and division are not automatic for ratios. Ratios A:B and B:C are compounded to form A:
C, which may be thought of as the product of the two ratios, and the duplicate ratio mentioned above is a
special case of a compound ratio. But the compound of two ratios A:B and C:D depends on the middle
terms B and C being the same. The proof of proposition V.18 assumes that fourth proportionals exist, a
property unjustified by any postulate, but if fourth proportionals do exist, then the ratio C:D is equal to
some ratio B:E, and then the compound of A:B and C:D is the compound of A:B and B:E, and that
compound is A:E. Thus, multiplication is an operation when fourth proportionals exist. Division is also
an operation when fourth proportionals exist since D:C may be thought of as the reciprocal of C:D.
Several kinds of ratios appear in the Elements. There are ratios of numbers, ratios of lines constructable
in plane geometry, ratios of rectilinear angles, ratios of plane figures constructable in plane geometry,
and ratios of solids.
Numeric ratios, that is, ratios of numbers, are treated in the books on number theory, Books VII through
VIII. In modern terminology these numeric ratios are called "positive rational numbers." Numeric ratios
and proportions have a separate, simpler definition in VII.Def.20. That definition is compatible with the
definitions here in Book V, but that compatibility is not demonstrated in the Elements.
The problem with numeric ratios is that there are not enough of them. That is ratios of magnitudes are
not always equal to ratios of numbers.
The illustration to the right shows a square with side A and diameter B.
The ratio B to A does exist according to the next definition V.Def.4 since
some multiple of each is greater than the other. In modern terms this ratio
would be identified with the square root of 2 and is known to be an
irrational number, that is, it is not equal to a numeric ratio. It is,
nonetheless, a ratio in Euclid's terminology. The ratio B:A is a ratio of
lines, but it is not a ratio of numbers.
Since this and other ratios of lines are not ratios of numbers, a more general definition of ratio is
required. That more general definition is the one given here and continuing through V.Def.6.
In modern terminology, the numeric ratios are positive rational numbers. The field of all rational
numbers including 0 and the negative rational numbers is commonly denoted Q. The ratios of lines
constructable in plane geometry form the field extended from Q by closure under square roots. A
convenient notation for that field is Q . It is a much larger field, but does not include all real numbers.
For instance, the cube root of 2, needed for doubling a cube, the sine of 20°, needed for trisecting angles,
and pi, needed for squaring the circle, all are missing from Q .
The conic sections are part of solid geometry but they are not treated in the Elements. Cones are
discussed in Book XII, but their sections (intersections with planes) which include ellipses, parabolas,
and hyperbolas are not even defined in the Elements. Euclid's work on the Conics was superceded by
Apollonius' and no longer exists. Intersections of conics lead to lines of new lengths that can be used to
solve problems such as doubling a cube and trisecting an angle, but they don't help in squaring the circle.
Thus, there are more ratios of lines constructable in solid geometry than ratios of lines constructable in
plane geometry.
The ratios of rectilinear figures form the same field Q as the ratios of lines. This follows from the
theory of application of areas developed in Book I, see proposition I.44. But there are other plane figures
besides rectilinear ones: circles. The ratio of a circle to the square on its radius is pi. Thus, pi is a ratio of
plane figures even though it is not a ratio of lines.
Throughout Book V the only ratios that are considered are those with two terms in accordance with V.
Def.3, but in Book VII ratios of three or more terms are used in proposition VI.33. In that proposition, a
ratio A:B:C of three numbers is considered, and a certain proportion A:B:C = E:F:G is shown. These
multiterm ratios and proportions are probably left over from an earlier time. In any case, the multiterm
proportion may be interpreted as an abbreviation for two proportions A:B = E:F and B:C = F:G. Then it
follows ex aequali that A:C = E:G.
Book V introduction
© 1997
D.E.Joyce
Clark University
Definition 4
Magnitudes are said to have a ratio to one another which can, when multiplied, exceed one
another.
This definition limits the existence of ratios to comparable magnitudes of the same kind where
comparable means each, when multiplied, can exceed the other. The ratio doesn't exist when one
magnitude is so small or the other so large that no multiple of the one can exceed the other. This
definition excludes the ratio of a finite straight line to an infinite straight line and the ratio of an
infinitesimal straight line, should any exist, to a finite straight line.
The result on horn angles in proposition III.16 excludes ratios between horn angles and rectilinear
angles. That proposition states that a horn angle is less than any rectilinear angle, hence no multiple of a
horn angle is greater than a rectilinear angle. The situation of horn angles is much worse than that,
however, since horn angles of different sizes aren't even comparable.
This definition is used repeatedly as a axiom for magnitudes rather than a definition. It is frequently
invoked in this book, starting with proposition V.8 but also required for more fundamental properties,
and elsewhere, such as the important proposition X.1. In the proofs of these propositions one magnitude
is less than another, and it is asserted that some multiple of the smaller is greater than the larger. Euclid
implicitly assumes that the magnitudes he discusses, except horn angles, are all comparable. Straight
lines, rectilinear angles, plane figures, and solids are all comparable to any other of the same type.
This principle of comparability should be explicit in order to justify the principle of comparability for
magnitudes of these kinds. One solution is to make it a postulate that straight lines are comparable. From
that postulate comparability of each of the other kinds of magnitudes could be proved.
Several of the propositions, stated and unstated, depend on this principle. Without it, some are simply
false for kinds of magnitude that have infinitesimals. If x and y are two magnitudes of the same kind,
then x is infinitesimal with respect to y, or y is infinite with respect to x, if no multiple of x is greater than
y. For example, horn angles are infinitesimal with respect to rectilinear angles. Although this definition
excludes ratios between horn angles and rectilinear angles, it allows a ratio between a rectilinear angle B
and the sum of a horn angle A and the rectilinear angle, and, according to the next three definitions, the
two ratios B:(A + B) and B:B so not satisfy the law of trichotomy, that is, they aren't the same ratio but
neither is greater than the other either. Examples involving infinitesimals can be useful to show which
propositions require treating this definition as an axiom.
Book V introduction
© 1997
D.E.Joyce
Clark University
Definition 7
When, of the equimultiples, the multiple of the first magnitude exceeds the multiple of the
second, but the multiple of the third does not exceed the multiple of the fourth, then the first is
said to have a greater ratio to the second than the third has to the fourth.
Definition 5 explained when two ratios were equal, namely, w:x = y:z when
there are numbers n and m such that nw > mx but ny is not greater than mz.
When defining greater and lesser, there are a number of properties that should be verified. These are
various transitivities and the law of trichotomy, some of the same properities of greater and lesser that
magnitudes have. (See the Guide for the Common Notions.)
If u:v < w:x, and w:x = y:z, then u:v < y:z.
If u:v = w:x, and w:x < y:z, then u:v < y:z.
If u:v < w:x, and w:x < y:z, then u:v < y:z.
Euclid only has the first property, which is is proposition V.13. Its proof depends only on the definitions.
The second is so much like it that it isn't mentioned but it is used in the same way as the third. The third
one is quite easy to prove.
Stated for ratios, the law of trichotomy says that for any two ratios w:x and y:z, exactly one of the
following three cases holds: w:x < y:z, or w:x = y:z, orw:x > y:z.
Euclid missed the law of trichotomy for magnitudes in his list of Common Notions, and he missed it for
ratios, too. The side of the law which says at most one of the three cases can occur is first used in
proposition V.9, while the side which says at least one occurs is first used in V.10.
The side of the law which says at most one of the three cases can occur is fairly easy to prove.
From the definitions themselves it is clear that w:x > y:z contradicts w:x = y:z. The first says there are n
and m such that nw > mx but ny is not greater than mz, while the second concludes from nw > mx that
ny > mz yielding a contradiction based on the law of trichotomy for magnitudes. Similarly w:x < y:z
contradicts w:x = y:z.
Once transitivity has been shown, it can be shown that w:x > y:z contradicts w:x < y:z, for then w:x > w:x,
and that contradicts w:x = w:x. (There are also proofs that don't depend on transitivity.)
The other side of the law of trichotomy, the one that says at least one of the three cases holds, is a bit
harder to prove, and it depends on treating V.Def.4 as an axiom of comparability. In fact, it is false
without it. First, a proof using V.Def.4 as an axiom, then a counterexample to show that's necessary.
A proof of trichotomy. Let w:x and y:z be any two ratios. We need to show that one of the three cases
holds. We'll assume the ratios aren't the same and show one of them is greater than the other. When
they're not the same, then there are numbers m and n such that
We have three cases to consider, and two of them are easy. In one case, nw > mx but not ny > mz, so for
that case w:x > y:z. In another case, nw < mx but not ny < mz, so for that case w:x < y:z.
Consider now the last case: nw = mx but ny does not equal mz. Then one of ny and mz is greater, say
ny > mz. Now using V.Def.4 as an axiom, there is some some number k such that k(ny - mz) > z. Since k
(ny - mz) = kny - kmz, therefore kny > kmz + z, that is, kny > (km + 1)z. But knw = kmx, and
kmx < (km + 1)x. Therefore, (kn)w < (km + 1)x. But (kn)y > (km + 1)z. Therefore w:x < y:z. Q.E.D.
Now, the ratios x:x and x:(x + y) are not equal since 2x = 2x but 2x < 2(x + y). Next, x:(x + y) is not
greater than x:x, since nx > m(x + y) implies nx > mx. Finally, x:x is not greater than x:(x + y), for if
nx > mx, then n > m, so nx is not less than mx > x, and since y is an infinitesimal with respect to x,
x > my, therefore nx - mx > my, that is, nx > m(x + y).
In summary, since we have a proof of trichotomy that uses V.Def.4 as an axiom of comparability, and a
counterexample of trichotomy that violates the axiom of comparability, we can conclude that any proof
trichotomy requires the axiom of comparability.
Book V introduction
© 1997
D.E.Joyce
Clark University
Definitions 1 and 2
Def. 1. A magnitude is a part of a magnitude, the less of the greater, when it measures the greater.
Def. 2. The greater is a multiple of the less when it is measured by the less.
The two magnitudes mentioned in each definition are of the same kind. Following Euclid, they are
illustrated here as lines, but they could both be planar figures, or solids, or angles, or any other kind of
magnitude so long as they are of the same kind.
© 1997
D.E.Joyce
Clark University
http://aleph0.clarku.edu/~djoyce/java/elements/bookV/defV1.html5/26/2008 3:50:39 PM
Euclid's Elements, Book V, Definitions 5 and 6
Definitions 5 and 6
Def. 5. Magnitudes are said to be in the same ratio, the first to the second and the third to the
fourth, when, if any equimultiples whatever are taken of the first and third, and any equimultiples
whatever of the second and fourth, the former equimultiples alike exceed, are alike equal to, or
alike fall short of, the latter equimultiples respectively taken in corresponding order.
Def. 6. Let magnitudes which have the same ratio be called proportional.
Definition 5 defines two ratios w:x and y:z to be the same, written w:x = y:z, when for all numbers n and
m it is the case that if nw is greater, equal, or less than mx, then ny is greater, equal, or less than mz,
respectively, that is,
Note that whenever the symbol >=< is used there are three parallel statements being made.
The four magnitudes do not all have to be of the same kind, but the first pair w and x need to be of the
one kind, and the second pair y and z of one kind, either the same kind as that of w and x or a different
kind. Perhaps the best illustration of these definitions comes from proposition VI.1 in which Euclid first
uses them to construct a proportion.
The goal in this proposition is to show that the lines are proportional to the triangles. More precisely, the
line BC is to the line CD as the triangle ABC is to the triangle ACD, that is, the ratio BC:CD of lines is
the same as the ratio ABC:ACD of triangles. Even though the ratios derive from different kinds of
magnitudes, they are to be compared and shown equal.
According to Definition 5, in order to show the ratios are the same, Euclid takes any one multiple of BC
and ABC (which he illustrates by taking three times each), and any one multiple of CD and ACD (which
he also illustrates by taking three times each). Then he proceeds to show that the former equimultiples,
namely HC and CL, alike exceed, are alike equal to, or alike fall short of, the latter equimultiples,
namely, AHC and ACL.
Symbolically, in order to prove BC:CD = ABC:ACD, Euclid proves for any numbers n and m that the
line n BC is greater, equal, or less than the line m CD when the triangle n ABC is greater, equal, or less
than the triangle m ACD. We will abbreviate this condition symbolically as
Note that in order to check this condition, it is only necessary to compare lines to lines and planar
figures to planar figures. To see how Euclid does this, refer to VI.1.
As it sometimes happens, a ratio of two magnitudes A:B is the same as a ratio of numbers m:n. Take for
instance the case when A is a line that is twice a line U while B is a line that is three times the line U.
Then, we could show that the ratio of magnitudes A:B is the same as the numrical ratio 2:3. Such ratios
are studied in detail in Book X. That book begins by defining in X.Def.1. what it means for two
quantities to be "commensurable." For instance, the two lines A and B are commensurable since there is
a unit U that measures both. Later in Book X (propositions X.5 and X.6) it is explicitly shown that two
magnitudes are commensurable if and only if their ratio is a numeric ratio.
Using modern concepts and notations, we can more easily see what the general definition of equality of
two magnitudes means. If we treat ratios as real numbers, the a proportion such as the one described
above, BC:CD = ABC:ACD, means that the ratio BC:CD compares to all numerical ratios (that is,
rational numbers) m/n the same way that ABC:ACD does. Another way of saying this is that equality of
two real numbers is determined by their relation to all rational numbers. This is often expressed by
saying that the set of rational numbers is dense in the set of real numbers.
Of course, Euclid did not have what modern mathematicians call real numbers. Indeed, there is an
ontological difference between real numbers and Euclid's ratios. Some real numbers are not ratios of the
magnitudes of any kind mentioned in the Elements.
Equivalence relations were defined in the Guide for V.Def.3. Three things need to be checked to see if
proportion is an equivalence relation: reflexivity, symmetry, and transitivity.
First, reflexivity. Is it the case for any pair of magnitudes of the same type A and B that A and B are in
the same ratio as A and B? That means for any numbers m and n,
That is trivial.
Second, symmetry. Is it the case that if A and B are in the same ratio as C and D, then C and D are in the
same ratio as A and B? The first says
This can be shown using the law of trichotomy for magnitudes. (Suppose nC > mD. If nA is not greater
than mB, then it is less or equal, but then nC is less or equal to mD, contradicting nC > mD. etc.) Euclid
missed symmetry, but he uses it very frequently.
Third, transitivity. Euclid states this explicitly in proposition V.11. The proof relies only on the
definition.
When A and B are in the same ratio as C and D, then the four magnitudes are said to be proportional, or
in proportion, according to definition 6. Is that the same as saying the ratios A:B and C:D are equal?
A more fundamental question is "do ratios exist?" Are they some kind of mathematical object like
numbers and magnitudes? The Elements do not require it. Instead, proportion is a relation held between
one pair of magnitudes and another pair of magnitudes. Yet it is very easy to read Book V as though
ratios are mathematical objects of some abstract variety. And it's easy to read "A and B have the same
ratio as C and D" as saying that the ratio A:B is the same ratio as C:D.
Not every relation allows that reading, but equivalence relations do, and proportion is an equivalence
relation.
The philosophical questions "do ratios exist?" and "is a proportion equality of ratios?" can be converted
to the question "why do equivalence relations create entities?" or a little more conservatively, "why do
equivalence relations allow us to think and act as if the entities exist?"
It is hard to imagine that Euclid did not think of ratios as things and proportions as equalities, especially
since the next definition defines when one ratio is larger than another. Perhaps he did but continued to
write noncommittally.
Book V introduction
© 1997, 2002
D.E.Joyce
Clark University
Definitions 8 through 10
Def. 8. A proportion in three terms is the least possible.
Def. 9. When three magnitudes are proportional, the first is said to have to the third the duplicate
ratio of that which it has to the second.
Def. 10. When four magnitudes are continuously proportional, the first is said to have to the
fourth the triplicate ratio of that which it has to the second, and so on continually, whatever be
the proportion.
In the illustration A, B, and C form three terms for the proportion A:B = B:C,
therefore the ratio A:C is the duplicate ratio of A:B. For a numerical
example, 9:4 is the duplicate ratio of 3:2.
The illustration also shows a continued proportion of four magnitudes, A, B, C, and D, since A:B = B:
C = C:D. Also, A:D is the triplicate ratio of A:B. For a numerical example, 27:8 is the triplicate ratio of
3:2.
Book V introduction
© 1997
D.E.Joyce
Clark University
Definitions 11 through 13
Def. 11. Antecedents are said to correspond to antecedents, and consequents to consequents.
Def. 12. Alternate ratio means taking the antecedent in relation to the antecedent and the
consequent in relation to the consequent.
Def. 13. Inverse ratio means taking the consequent as antecedent in relation to the antecedent as
consequent.
The figure illustrates the proportion A:B = C:D. Thus, A and C are
corresponding terms since they're the antecedents. Also, B and D
are corresponding terms since they're the consequents.
Note that for alternate ratios to exist, all four magnitudes must be
of the same kind. The alternate ratios in this proportion are A:C
and B:D. Euclid proves these are the same ratio in proposition
V.16. Thereafter, given one proportion A:B = C:D, he concludes
alternately the alternate proportion A:C = B:D.
The ratio inverse to A:B is B:A. It is evident from the definition V.Def.5 that A:B = C:D and B:A = D:C
reduce to the same conditions on A, B, C, and D. Therefore, if two ratios are the same, then their two
inverse ratios are also the same. For some reason, this statement is misplaced as the corollary after
proposition V.7.
Several of the propositions are stated using the antecedent terms but they apply as well for the
consequent terms by inversion. For example, proposition V.24 says that if u:v = w:x and y:v = z:x, then
(u + y):v = (w + z):x. But the statement using consequents is valid, too: if v:u = x:w and v:y = x:z, then v:
(u + y) = x(w + z).
The symmetry of the antecedent and consequent terms of a ratio a:b is not, however, one of perfect
parallelism. They're opposite in regard to order. Proposition V.8 shows that the ratio is a:b increasing in
a since if a increases then the ratio increases, but the ratio is decreasing in b since the if b increases then
the ratio decreases. But that's still a kind of symmetry.
Book V introduction
© 1997
D.E.Joyce
Clark University
Definitions 14 through 16
Def. 14. A ratio taken jointly means taking the antecedent together with the consequent as one in
relation to the consequent by itself.
Def. 15. A ratio taken separately means taking the excess by which the antecedent exceeds the
consequent in relation to the consequent by itself.
Def. 16. Conversion of a ratio means taking the antecedent in relation to the excess by which the
antecedent exceeds the consequent.
Taking jointly the ratio u:v yields the ratio (u + v):v. Taking separately the ratio (u + v):v returns the
ratio u:v. Taking the ratio (u + v):v in conversion yields the ratio (u + v):u.
These conversions are only important when the ratios are in proportions.
The following three proportions are shown to be
equivalent in propositions V.17 and V.18.
1. (u + v):v = (x + y):y.
2. (u + v):u = (x + y):x.
3. u:v = x:y.
Proposition V.17 and V.18 show proportions 1 and 3 are equivalent. That means proportion 2 and the
inverse of 3, v:u = y:x, are also equivalent. And of course, 3 and its inverse are equivalent, so all three
proportions are equivalent.
Furthermore, when all the magnitudes are of the same kind, then the alternate proportions are also
equivalent by V.16 making six equivalent statements.
4. (u + v):(x + y) = v:y.
5. (u + v):(x + y) = u:x
6. u:x = v:y
Proposition V.19 goes on to say that 4 implies 5, and its corollary says 1 implies 2.
Heath translates "taken jointly," "taken separately," and "in conversion" by the Latin words
componendo, separando, and convertendo, respectively.
Book V introduction
© 1997
D.E.Joyce
Clark University
Definitions 17 and 18
Def. 17. A ratio ex aequali arises when, there being several magnitudes and another set equal to
them in multitude which taken two and two are in the same proportion, the first is to the last
among the first magnitudes as the first is to the last among the second magnitudes. Or, in other
words, it means taking the extreme terms by virtue of the removal of the intermediate terms.
Def. 18. A perturbed proportion arises when, there being three magnitudes and another set equal
to them in multitude, antecedent is to consequent among the first magnitudes as antecedent is to
consequent among the second magnitudes, while, the consequent is to a third among the first
magnitudes as a third is to the antecedent among the second magnitudes.
16 Select topic
Book V introduction
© 1997
D.E.Joyce
Clark University
Proposition 1
If any number of magnitudes are each the same multiple of the same number of other
magnitudes, then the sum is that multiple of the sum.
Let any number of magnitudes AB and CD each be the same multiple of magnitudes E and F V.Def.2
respectively.
I say that the sum of AB and CD is the same multiple of the sum of E and F that AB is of E.
Since AB is the same multiple of E that CD is of F, therefore there are as many magnitudes
in AB equal to E as there are in CD equal to F.
Now, since AG equals E, and CH equals F, therefore the sum of AG and CH equals the sum
of E and F.
For the same reason GB equals E, and the sum of GB and HD equals the sum of E and F.
Therefore, there are as many magnitudes in AB equal to E as there are in the sum of AB and
CD equal to the sum of E and F. Therefore, the sum of AB and CD is the same multiple of
the sum of E and F that AB is of E.
Therefore, if any number of magnitudes are each the same multiple of the same number of other
magnitudes, then the sum is that multiple of the sum.
Q.E.D.
In modern terminology, this proposition states that multiplication by numbers distributes over addition
of magnitudes, that is,
Here, the m is a number, and all the xi's are magnitudes of the same kind.
Euclid always displays his magnitudes as lines, but they could be magnitudes of other kinds, like plane
regions, for instance. In this proposition, all the magnitudes are of the same kind.
Euclid's proof is only for the simplest nontrivial case. He takes the number n of magnitudes to be 2, and
the multiple m also to be 2, so he proves that if x1 = m y1 and x2 = m y2, then x1 + x2 = m (y1 + y2).
Throughout Book V, Euclid proves the general numerical case by a particular case. The numbers he
chooses are usually 2 and 3.
Proposition V.1 is used in the proofs of four other propositions, namely, V.5, V.8, V.12, and V.17.
18 Select topic
Book V introduction
© 1996
D.E.Joyce
Clark University
Proposition 2
If a first magnitude is the same multiple of a second that a third is of a fourth, and a fifth also
is the same multiple of the second that a sixth is of the fourth, then the sum of the first and fifth
also is the same multiple of the second that the sum of the third and sixth is of the fourth.
Let a first magnitude AB be the same multiple of a second C that a third DE is of a
fourth F, and let a fifth BG be the same multiple of the second C that a sixth EH is V.Def.2
of the fourth F.
I say that the sum AG of the first and fifth is the same multiple of the second, C,
that the sum DH of the third and sixth is of the fourth, F.
For the same reason there are as many magnitudes in BG equal to C as there are in
EH equal to F. Therefore, there are as many magnitudes in the whole AG equal to C
as there are in the whole DH equal to F.
Therefore the sum AG of the first and fifth is the same multiple of the second, C,
that the sum DH of the third and sixth is of the fourth, F.
Therefore, if a first magnitude is the same multiple of a second that a third is of a fourth, and a fifth
also is the same multiple of the second that a sixth is of the fourth, then the sum of the first and fifth
also is the same multiple of the second that the sum of the third and sixth is of the fourth.
Q.E.D.
This proposition simply states that sums of equimultiples are equimultiples, that is, if mc and mf are
equimultiples of c and f, and nc and nf are also equimultiples of c and f, then the sums mc + nc and
mf + nf are also equimultiples of c and f. The proof depends on a form of distributivity, namely, that
multiplication by magnitudes distributes over addition of numbers.
(m + n)c = mc + nc.
Note that the magnitudes don't all have to be of the same kind. Different colors are used in the figures
here to indicate different kinds of magnitudes.
Proposition V.2 is used in the proofs of three other propositions, namely, V.3, V.6, and V.17.
Book V introduction
© 1996
D.E.Joyce
Clark University
Proposition 3
If a first magnitude is the same multiple of a second that a third is of a fourth, and if
equimultiples are taken of the first and third, then the magnitudes taken also are equimultiples
respectively, the one of the second and the other of the fourth.
Let a first magnitude A be the same multiple of a second B that a third C is of a fourth D, and V.Def.2
let equimultiples EF and GH be taken of A and C.
Since EF is the same multiple of A that GH is of C, therefore there are as many magnitudes
as in EF equal to A as there are in GH equal to C.
And, since A is the same multiple of B that C is of D, while EK equals A, and GL equals C,
therefore EK is the same multiple of B that GL is of D.
Since a first magnitude EK is the same multiple of a second B that a third GL is of a fourth
D, and a fifth KF is the same multiple of the second B that a sixth LH is of the fourth D, V.2
therefore the sum EF of the first and fifth is the same multiple of the second B that the sum
GH of the third and sixth is of the fourth D.
Therefore, if a first magnitude is the same multiple of a second that a third is of a fourth, and if
equimultiples are taken of the first and third, then the magnitudes taken also are equimultiples
respectively, the one of the second and the other of the fourth.
Q.E.D.
This proposition says that equimultiples of equimultiples are equimultiples, that is, if w and x are
equimultiples of y and z, and u and v are equimultiples of w and x, then u and v are equimultiples of y
and z. The proof depends on an associativity of multiplication: m (ny) = (mn) y. In Euclid's proof, n is 3
and m is 2.
As in the last proposition, the magnitudes need not all be of the same kind.
Although this proposition is not actually a statement about ratios, it can be interpreted as one. The
hypotheses that A and C are equimultiples of B and D can be interpreted as a proportion A:B = C:D, and
the conclusion that mA and mC are equimultiples of B and D can be interpreted as a proportion mA:
B = mC:D. Under these interpretations this proposition becomes a special case of the next, and it is the
special case that is used to prove the general case in the next proposition.
Book V introduction
© 1996
D.E.Joyce
Clark University
Proposition 4
If a first magnitude has to a second the same ratio as a third to a fourth, then any equimultiples
whatever of the first and third also have the same ratio to any equimultiples whatever of the
second and fourth respectively, taken in corresponding order.
Let a first magnitude A have to a second B the same ratio as a third C to a fourth D, and let
equimultiples E and F be taken of A and C, and G and H other, arbitrary, equimultiples of B
and D.
I say that E is to G as F is to H.
And, since A is to B as C is to D, and equimultiples K and L have been taken of A and C, and
other, arbitrary, equimultiples M and N of B and D, therefore, if K is in excess of M, then L is V.Def.5
in excess of N; if it is equal, equal; and if less, less.
And K and L are equimultiples of E and F, and M and N are other, arbitrary, equimultiples of V.Def.5
G and H, therefore E is to G as F is to H.
Therefore, if a first magnitude has to a second the same ratio as a third to a fourth, then any
equimultiples whatever of the first and third also have the same ratio to any equimultiples whatever of
the second and fourth respectively, taken in corresponding order.
Q.E.D.
Note how Euclid uses the definition to prove that the two ratios pa:qb and pc:qd are the same. (Here, a
and b are magnitudes of one kind, and c and d are magnitudes of another kind, but p and q are numbers.)
We are given a:b = c:d. That means for any numbers m and n that
We have to prove that pa:qb = pc:qd for any numbers p and q. That means, we have to prove that for any
m and n,
Proposition V.4 is used in the proof of one other proposition, namely, V.22.
Book V introduction
© 1996
D.E.Joyce
Clark University
Proposition 5
I say that the remainder EB is also the same multiple of the remainder FD that the whole AB is of
the whole CD.
Therefore AB is the same multiple of each of the magnitudes GF and CD. Therefore GF equals
CD.
Subtract CF from each. Then the remainder GC equals the remainder FD.
And, since AE is the same multiple of CF that EB is of GC, and GC equals DF, therefore AE is
the same multiple of CF that EB is of FD.
But, by hypothesis, AE is the same multiple of CF that AB is of CD, therefore EB is the same
multiple of FD that AB is of CD.
That is, the remainder EB is the same multiple of the remainder FD that the whole AB is of the
whole CD.
Therefore, If a magnitude is the same multiple of a magnitude that a subtracted part is of a subtracted
part, then the remainder also is the same multiple of the remainder that the whole is of the whole.
Q.E.D.
This proposition is analogous to V.1 but involves differences rather than sums. It states that
multiplication by numbers distributes over subtraction of magnitudes: m (x - y) = mx - my.
Note that all the magnitudes in this proposition are of the same kind.
Make CGso that EBis the same multiple of CGthat AEis of CF.
so that, if for example, AE is a third of CF, then CG is to be made a third of EB. Such constructions
cannot be made for all kinds of magnitudes, in particular, angles and arcs.
Alternative proofs that don't require constructions of parts are relatively easy to find. A more interesting
problem of general constructions for magnitudes is discussed in the Guide for proposition V.18.
Book V introduction
© 1996
D.E.Joyce
Clark University
Proposition 6
If two magnitudes are equimultiples of two magnitudes, and any magnitudes subtracted from
them are equimultiples of the same, then the remainders either equal the same or are
equimultiples of them.
Let two magnitudes AB and CD be equimultiples of two magnitudes E and F, and let AG and CH
subtracted from them be equimultiples of the same two E and F.
I say that the remainders GB and HD either equal E and F or are equimultiples of them.
Make CK equal to F.
Since AG is the same multiple of E that CH is of F, while GB equals E, and KC equals F, V.2
therefore AB is the same multiple of E that KH is of F.
Since then each of the magnitudes KH and CD is the same multiple of F, therefore KH equals
CD.
Subtract CH from each. Then the remainder KC equals the remainder HD.
Similarly we can prove that, even if GB is a multiple of E, HD is also the same multiple of F.
Therefore, if two magnitudes are equimultiples of two magnitudes, and any magnitudes subtracted
from them are equimultiples of the same, then the remainders either equal the same or are
equimultiples of them.
Q.E.D.
The proposition states that if ma and mb are equimultiples of a and b, and na and nb are also
equimultiples, then the differences, ma - na and mb - nb are more equimultiples. It's analogous to
proposition V.2 which was for addition.
Its proof depends on a distributivity, namely that multiplication by magnitudes distributes over
subtraction of numbers: (m - n)a = ma - na. Euclid takes 4 as m and 3 as n. He has two cases since since
he doesn't take 1 to be a number.
Book V introduction
© 1996
D.E.Joyce
Clark University
Proposition 7
Equal magnitudes have to the same the same ratio; and the same has to equal magnitudes the
same ratio.
Let A and B be equal magnitudes and C an arbitrary magnitude.
I say that each of the magnitudes A and B has the same ratio to C, and C has the same ratio to
each of the magnitudes A and B.
And D and E are equimultiples of A and B, while F is another, arbitrary, multiple of C, V.Def.5
therefore A is to C as B is to C.
I say next that C also has the same ratio to each of the magnitudes A and B.
With the same construction, we can prove similarly that D equals E, and F is some other
magnitude. If therefore F is in excess of D, it is also in excess of E; if equal, equal; and, if
less, less.
And F is a multiple of C, while D and E are other, arbitrary, equimultiples of A and B, V.Def.5
therefore C is to A as C is to B.
Therefore, equal magnitudes have to the same the same ratio; and the same has to equal magnitudes
the same ratio.
Q.E.D.
Corollary
From this it is manifest that, if any magnitudes are proportional, then they are also proportional
inversely.
This proposition says that if a = b, then a:c = b:c, and c:a = c:b. The proposition is evident. Its converse
is given in V.9.
The corollary is misplaced. There is nothing relevant in the proposition. There's no way it could yield
the corollary since the proposition requires all the magnitudes to be of the same kind and the corollary
doesn't. But the corollary is valid, and it follows easily from definition V.Def.5.
Such a basic property of ratios as this is used frequently when ratios are mentioned. It is used in a few
times in Book V starting with V.10, frequently in Book VI, and ocassionally in later books.
Book V introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 8
Of unequal magnitudes, the greater has to the same a greater ratio than the less has; and the
same has to the less a greater ratio than it has to the greater.
Let AB and C be unequal magnitudes, and let AB be greater, and let D be another, arbitrary,
magnitude.
I say that AB has to D a greater ratio than C has to D, and D has to C a greater ratio than it
has to AB.
Take L double of D and M triple of it, and successive multiples increasing by one, until
what is taken is the first multiple of D that is greater than K. Let it be taken, and let it be N (V.Def.4)
which is quadruple of D and the first multiple of it greater than K.
And, since FG is the same multiple of AE that GH is of EB, therefore FG is the same V.1
multiple of AE that FH is of AB.
And FG is greater than D, therefore the whole FH is greater than the sum of D and M.
But the sum of D and M equals N, inasmuch as M is triple D, and the sum of M and D is
quadruple D, while N is also quadruple D, therefore the sum of M and D equals N.
But FH is greater than the sum of M and D, therefore FH is in excess of N, while K is not in
excess of N.
And FH and K are equimultiples of AB and C, while N is another, arbitrary, multiple of D, V.Def.7
therefore AB has to D a greater ratio than C has to D.
With the same construction, we can prove similarly that N is in excess of K, while N is not
in excess of FH.
And N is a multiple of D, while FH and K are other, arbitrary, equimultiples of AB and C, V.Def.7
therefore D has to C a greater ratio than D has to AB.
Then we can prove similarly that FH and K are equimultiples of AB and C, and, similarly, (V.Def.4)
take N the first multiple of D that is greater than FG, so that FG is again not less than M.
But GH is greater than D, therefore the whole FH is in excess of the sum of D and M, that
is, of N.
Now K is not in excess of N, inasmuch as FG also, which is greater than GH, that is, than
K, is not in excess of N.
And in the same manner, by following the above argument, we complete the demonstration.
Therefore, of unequal magnitudes, the greater has to the same a greater ratio than the less has; and
the same has to the less a greater ratio than it has to the greater.
Q.E.D.
Although the statement of this proposition is easy to comprehend, its proof is difficult. It says that if
x > y, then x:z > y:z but z:x < z:y. Its converse is proposition V.10.
At four points in the proof V.Def.4 is used as an axiom of comparabililty rather than a definition. The
first instance:
Then the less of the magnitudes AEand EB,if multiplied, will eventually be greater than D.
In fact the axiom of comparability is required for this proposition since it is false when infinitesimals are
allowed. When y is infinitesimal with respect to x, then the first statement of the proposition doesn't hold
since x > x - y but it is not the case that x:x > (x - y):x, and the second statement doesn't hold since
x + y > x, but not x:x > x:(x + y).
The proof is slightly more comprehensible when modern algebraic notation is used since that clarifies its
overall structure. Every magnitude in Euclid's proof is represented by a name and illustrated by a line.
With an algebraic notation, we can refer to a magnitude by a formula. For instance, if we let a be AB and
c be C, then we can use a - c for AE, thus reducing the number of variables and easing comprehension.
We can also have variables for numbers, instead of having to choose a specific number as Euclid does
when he takes N to be 4D.
But algebra obscures much, too. Euclid carefully proved distributivity of multiplication by numbers over
addition of magnitudes in V.1, which is used in this proof. We manipulate algebraic expressions almost
automatically. In order to be as correct as Euclid, we should verify the rules of algebra and be aware
when we use them.
With these preliminary qualifications, let's look at a translation of the proof into symbolic algebra.
a = AB
To prove: if a > c, then a:d > c:d, but d:c > d:a. c = C = EB
d=D
m(a - c) = FG
Case 1: Suppose a - c < c. Let m be a number such that m(a - c) > d.
mc = GH = K
Let n be the smallest number such that nd > mc. [What happens when n = 1 to nd = N
Euclid's proof?] (n - 1)d = M
Since mc is not less than (n - 1)d, and m(a - c) > d, therefore, by adding, ma > nd.
ma = FH
But mc is not greater than nd. Therefore a:d > c:d.
Also nd > mc but nd is not greater than ma. Therefore d:c > d:a.
Case 2: Suppose c < a - c. Let Let m be a number such that mc > d. (Same as above)
Since m(a - c) is not less than (n - 1)d, and mc > d, therefore, by adding, ma > nd.
[Euclid says to do the rest in the same manner: Since a - c > c, therefore m(a - c)
> mc. But nd > m(a - c), therefore nd > mc. But ma > nd, therefore a:d > c:d.
Also nd > mc but nd is not greater than ma. Therefore d:c > d:a.]
Proposition V.8 is used a few times in Book V starting with the next proposition.
Book V introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 9
Magnitudes which have the same ratio to the same equal one another; and magnitudes to
which the same has the same ratio are equal.
Let each of the magnitudes A and B have the same ratio to C.
Next, let C have the same ratio to each of the magnitudes A and B.
Otherwise, C would not have the same ratio to each of the magnitudes A and B, but it does, V.8
therefore A equals B.
Therefore, magnitudes which have the same ratio to the same equal one another; and magnitudes to
which the same has the same ratio are equal.
Q.E.D.
Besides the previous proposition, the proof relies on the law of trichotomy for ratios, the part which says
that a:b < a:c and a:b = a:c cannot both occur. Although Euclid didn't prove that, it follows easily from
the definitions in V.Def.5 and V.Def.7.
This proposition relies on using V.Def.4 as an axiom of comparability through its use of the previous
proposition. The axiom is required since the statement of the proposition is false when a is c + y, and b is
c + 2y where y is an infinitesimal with respect to c.
This proposition is used occasionally in Books VI, VII, X, XI, and XII to conclude equality of geometric
magnitudes.
Book V introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 10
Of magnitudes which have a ratio to the same, that which has a greater ratio is greater; and
that to which the same has a greater ratio is less.
Let A have to C a greater ratio than B has to C.
Now A does not equal B, for in that case each of the magnitudes
A and B would have the same ratio to C, but they do not, V.7
therefore A does not equal B.
Nor is A less than B, for in that case A would have to C a less ratio than B has to C, but it does V.8
not, therefore A is not less than B.
Now B does not equal A, for in that case C would have the same ratio to each of the magnitudes V.7
A and B, but it does not, therefore A does not equal B.
Nor is B greater than A, for in that case C would have to B a less ratio than it has to A, but it does V.8
not, therefore B is not greater than A.
But it was proved that it is not equal either, therefore B is less than A.
Therefore, of magnitudes which have a ratio to the same, that which has a greater ratio is greater;
and that to which the same has a greater ratio is less.
Q.E.D.
Part of the law of trichotomy for ratios is used in this proof, the part which says at most one of the three
cases a:c < b:c, a:c = b:c, or a:c > b:c, can occur.
Euclid's proof relies on using V.Def.4 as an axiom of comparability since it uses proposition V.8 and the
law of trichotomy for ratios. But the proposition can also be proved without the axiom.
Suppose a:c > b:c.Then there are numbers mand nsuch that na > mcbut nbis not greater than mc.
Therefore na > nb.Therefore a > b.Thus a:c > b:cimplies a > b.
Book V introduction
© 1996
D.E.Joyce
Clark University
Proposition 11
Ratios which are the same with the same ratio are also the same with one another.
Let A be to B as C is to D, and let C be to D as E is to F.
I say that A is to B as E is to F.
Take equimultiples G, H, and K of A, C, and E, and take other, arbitrary, equimultiples L, M, and N of
B, D, and F.
Then since A is to B as C is to D, and of A and C equimultiples G and H have been taken, and
of B and D other, arbitrary, equimultiples L and M, therefore, if G is in excess of L, H is also V.Def.5
in excess of M; if equal, equal; and if less, less.
But we saw that, if H was in excess of M, G was also in excess of L; if equal, equal; and if
less, less, so that, in addition, if G is in excess of L, K is also in excess of N; if equal, equal; V.Def.5
and if less, less.
And G and K are equimultiples of A and E, while L and N are other, arbitrary, equimultiples V.Def.5
of B and F, therefore A is to B as E is to F.
Therefore, ratios which are the same with the same ratio are also the same with one another.
Q.E.D.
This proposition expresses the transitivity of the relation of being the same when applied to ratios. After
this proposition (and the easily proved properties of reflexivity and symmetry, see the Guide to
definition V.Def.6), the expression "two ratios are the same," or the equivalent expression "two ratios
are equal," is justified. The proof follows directly from the definition. What is remarkable is that
Eudoxus, or Euclid, recognized that this proposition needed to be proved.
The magnitudes may be of three different kinds with A and B of one kind, C and D of a second kind, and
E and F of a third kind.
Book V introduction
© 1996
D.E.Joyce
Clark University
Proposition 12
If any number of magnitudes are proportional, then one of the antecedents is to one of the
consequents as the sum of the antecedents is to the sum of the consequents.
Let any number of magnitudes A, B, C, D, E, and F be proportional, so that A is to B as C is to D, and
as E is to F.
Now G and the sum of G, H, and K are equimultiples of A and the sum of A, C, and E, since,
if any number of magnitudes are each the same multiple the same number of other V.1
magnitudes, then the sum is that multiple of the sum.
For the same reason L and the sum of L, M, and N are also equimultiples of B and the sum of V.Def.5
B, D, and F, therefore A is to B as the sum of A, C, and E is to the sum of B, D, and F.
Therefore, if any number of magnitudes are proportional, then one of the antecedents is to one of the
consequents as the sum of the antecedents is to the sum of the consequents.
Q.E.D.
The general form for this proposition is that if x1:y1 = x2:y2 = ... = xn:yn, then each of these ratios also
equals the ratio (x1 + x2 + ... + xn) : (y1 + y2 + ... + yn).
This proposition is used in V.15 and a few other propositions in books VI, X, and XII.
Book V introduction
© 1996
D.E.Joyce
Clark University
Proposition 13
If a first magnitude has to a second the same ratio as a third to a fourth, and the third has to
the fourth a greater ratio than a fifth has to a sixth, then the first also has to the second a
greater ratio than the fifth to the sixth.
Let a first magnitude A have to a second B the same ratio as a third C has to a fourth D, and let the
third C have to the fourth D a greater ratio than a fifth E has to a sixth F.
I say that the first A also has to the second B a greater ratio than the fifth E to the sixth F.
Since there are some equimultiples of C and E, and of D and F other equimultiples, such that
the multiple of C is in excess of the multiple of D, while the multiple of E is not in excess of
the multiple of F, let them be taken. Let G and H be equimultiples of C and E, and K and L V.Def.7
other, arbitrary, equimultiples of D and F, so that G is in excess of K, but H is not in excess
of L. Whatever multiple G is of C, let M also be that multiple of A, and, whatever multiple K
is of D, let N also be that multiple of B.
But H is not in excess of L, and M and H are equimultiples of A and E, and N and L other, V.Def.7
arbitrary, equimultiples of B and F, therefore A has to B a greater ratio than E has to F.
Therefore, if a first magnitude has to a second the same ratio as a third to a fourth, and the third has
to the fourth a greater ratio than a fifth has to a sixth, then the first also has to the second a greater
ratio than the fifth to the sixth.
Q.E.D.
This proposition states that if two ratios are equal, and one is greater than a third, then so is the other.
That is, if a:b = c:d and c:d > e:f, then a:b > e:f. The magnitudes may be of three different kinds with a
and b of one kind, c and d of a second kind, and e and f of a third kind.
The analogous statement for lesser ratios isn't stated, but, of course, it holds as well. Euclid uses it as
well as this proposition, in V.20.
So does transitivity: if a:b > c:d and c:d > e:f, then a:b > e:f. The proof isn't difficult, but without
symbolic algebra it becomes unwieldly. Euclid would have required 20 lines in his diagram.
This proposition is used in the next one as well as V.20 and V.21.
Book V introduction
© 1996
D.E.Joyce
Clark University
Proposition 14
If a first magnitude has to a second the same ratio as a third has to a fourth, and the first is
greater than the third, then the second is also greater than the fourth; if equal, equal; and if
less, less.
Let a first magnitude A have the same ratio to a second B as a third C has to a fourth D, and let
A be greater than C.
But that to which the same has a greater ratio is less, therefore D is less than B, so that B is V.10
greater than D.
Similarly we can prove that, if A equals C, then B equals D, and, if A is less than C, then B is
less than D.
Therefore, if a first magnitude has to a second the same ratio as a third has to a fourth, and the first is
greater than the third, then the second is also greater than the fourth; if equal, equal; and if less, less.
Q.E.D.
is used. This other form is more general since a and b may be of one kind while c and d can be of a
different kind. (See definition V.Def.12 and proposition V.16 for alternate proportions.) For example, in
proposition VI.25 there are the statements:
...the triangle ABCis to the triangle KGHas the parallelogram BEis to the parallelogram EF.
Therefore, alternately, the triangle ABCis to the parallelogram BEas the triangle KGHis to the
parallelogram EF.But the triangle ABCequals the parallelogram BE,therefore the triangle
KGHalso equals the parallelogram EF.
First, the proportion is converted to its alternate form by V.16. Then, it is claimed that since the first
equals the second, therefore the third equals the fourth. Clearly, V.14 is not being invoked otherwise the
alternate form of the proportion would not be mentioned.
... the rectangle BCby EFequals the rectangle BDby G,therefore CBis to BDas Gis to EF.But CBis
greater than BD,therefore Gis also greater than EF.
Here the first is greater than the second, so the third is greater than the fourth. Proposition VI.16 (if the
rectangle contained by the extremes equals the rectangle contained by the means, then the four straight
lines are proportional) was used to derive the proportion CB:BD = G:EF from the equality of the
rectangles, but it would have been just as easy to conclude CB:G = BD:EF, and then V.14 could be used.
Clearly, this proposition V.14 is not being invoked in either of these propositions, but the alternate form
is used instead. That suggests that the proofs of VI.25 and X.112 were written when V.14 wasn't
available.
is not difficult using the definition V.Def.5. Since a >=< b, therefore 2a >=< 2b. From the proportion a:
b = c:d it follows that 2c >=< 2d. Therefore c >=< d. Q.E.D. The proof is even easier when 1 is
considered to be a number.
Although this alternate form does not rely on using V.Def.4 as an axiom of comparability, the original
form does. The statement of the proposition is false when infinitesimals are allowed. For a particular
example, take the proportion x:(x + y) = x: (x + 2y) or its inverse.
This proposition is used in V.16 and a few other propositions in Books V, VI, X, XII, and XIII.
Book V introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 15
Divide AB into the magnitudes AG, GH, and HB equal to C, and divide
DE into the magnitudes DK, KL, and LE equal to F. Then the number of
java applet or image
the magnitudes AG, GH, and HB equals the number of the magnitudes
DK, KL, and LE.
And, since AG, GH, and HB equal one another, and DK, KL, and LE also equal one another, V.7
therefore AG is to DK as GH is to KL, and as HB is to LE.
Therefore one of the antecedents is to one of the consequents as the sum of the antecedents is to V.12
the sum of the consequents. Therefore AG is to DK as AB is to DE.
This proposition states that if n is any number and c and f any magnitudes of the same kind, then c:f = nc:
nf.
This proposition is used in the next one and a few others in Books V, VI, and XIII.
Book V introduction
© 1996
D.E.Joyce
Clark University
Proposition 16
If four magnitudes are proportional, then they are also proportional alternately.
Let A, B, C, and D be four proportional magnitudes, so that A is to B as C is to D.
Take equimultiples E and F of A and B, and take other, arbitrary, equimultiples G and H of C
and D.
Then, since E is the same multiple of A that F is of B, and parts have the same ratio as their V.15
equimultiples, therefore A is to B as E is to F.
But, if four magnitudes are proportional, and the first is greater than the third, then the V.14
second is also greater than the fourth; if equal, equal; and if less, less.
Now E and F are equimultiples of A and B, and G and H other, arbitrary, equimultiples of C V.Def.5
and D, therefore A is to C as B is to D.
Therefore, if four magnitudes are proportional, then they are also proportional alternately.
Q.E.D.
The four magnitudes A, B, C, and D need to be of the same kind. If A and B are of a different kind than
C and D, then the alternate ratios A:C and B:D would be "mixed." The Greek geometers did not accept
mixed ratios, but modern physicists and engineers routinely use them, as do we all since such a common
measurement as velocity is made out of a ratio of a distance to a time.
This proposition is used in V.19 and a couple others in Book V, and frequently in Books VI, X, XI, and
XII. Occasionally it is used when the magnitudes need not be all of the same kind, as it ought not.
Book V introduction
© 1996
D.E.Joyce
Clark University
Proposition 17
If magnitudes are proportional taken jointly, then they are also proportional taken separately.
Let AB, BE, CD, and DF be magnitudes proportional taken jointly, so that AB is to BE as V.Def.14
CD is to DF.
I say that they are also proportional taken separately, that is, AE is to EB as CF is to DF. V.Def.15
Take equimultiples GH, HK, LM, and MN of AE, EB, CF, and FD, and take other, arbitrary,
equimultiples, KO and NP of EB and FD.
Then, since GH is the same multiple of AE that HK is of EB, therefore GH is the same V.1
multiple of AE that GK is of AB.
But GH is the same multiple of AE that LM is of CF, therefore GK is the same multiple of
AB that LM is of CF.
Again, since LM is the same multiple of CF that MN is of FD, therefore LM is the same V.1
multiple of CF that LN is of CD.
But LM was the same multiple of CF that GK is of AB, therefore GK is the same multiple of
AB that LN is of CD.
Again, since HK is the same multiple of EB that MN is of FD, and KO is also the same
multiple of EB that NP is of FD, therefore the sum HO is also the same multiple of EB that V.2
MP is of FD.
Let GK be in excess of HO. Subtract HK from each. Therefore GH is also in excess of KO.
But we saw that, if GK was in excess of HO, then LN was also in excess of MP, therefore LN
is also in excess of MP, and, if MN is subtracted from each, then LM is also in excess of NP,
so that, if GH is in excess of KO, then LM is also in excess of NP.
Similarly we can prove that, if GH equals KO, then LM also equals NP; and if less, less.
And GH and LM are equimultiples of AE and CF, while KO and NP are other, arbitrary, V.Def.5
equimultiples of EB and FD, therefore AE is to EB as CF is to FD.
Therefore, if magnitudes are proportional taken jointly, then they are also proportional taken
separately.
Q.E.D.
The proposition says that if (w + x):x = (y + z):z, then w:x = y:z. Two of the magnitudes w and x can be
of one kind while the other two y and z are of another kind.
This proposition is used in the next two propositions and a couple in Book X.
Book V introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 18
If magnitudes are proportional taken separately, then they are also proportional taken jointly.
Let AE, EB, CF, and FD be magnitudes proportional taken separately, so that AE is to EB as V.Def.15
CF is to FD.
I say that they are also proportional taken jointly, that is, AB is to BE as CD is to FD. V.Def.14
Then, since AB is to BE as CD is to DG, they are magnitudes proportional taken jointly, so V.17
that they are also proportional taken separately. Therefore AE is to EB as CG is to GD.
But the first CG is greater than the third CF, therefore the second GD is also greater than V.14
the fourth FD.
Similarly we can prove that neither is it in that ratio to a greater, it is therefore in that ratio
to FD itself.
Therefore, if magnitudes are proportional taken separately, then they are also proportional taken
jointly.
Q.E.D.
This proposition is the converse of the last one. It says that if w:x = y:z, then (w + x):x = (y + z):z. As in
the last proposition, two of the magnitudes w and x can be of one kind while the other two y and z are of
another.
If CD:DFdoes not equal AB:BE,then AB:BE = CD:DGwhere DGis some magnitude greater or
less than DF.
Given the other three magnitudes, a fourth proportional DG is assumed. It is not asserted that the fourth
proportional can be constructed; it is only hypothetical. This is the beginning of a proof by contradiction.
This technique of assuming the existence of a fourth proportional to derive a contradiction is also used in
Book XII to prove various proportionalities of areas and volumes, for example, in proposition XII.2
which shows circles are proportional to the squares on their diameters. Eudoxus, who developed the
techniques of both Books V and XII, or Euclid, or both of them, accepted this technique as valid.
The problem is: do fourth proportionals exist? They certainly can't be constructed in all cases. The
problems of doubling a cube, squaring a circle, and trisecting an angle cannot be solved by plane
Euclidean methods, and they all involve inconstructable fourth proportionals. Take doubling a cube for
example. If C is a cube with an edge A, then the inconstructable edge B of a cube with double the
volume of C is the fourth proportional in C:(C+C) = A:B.
Is there a difference between existence and constructibility? Constructibility is a fairly clear concept
since there are postulates for what can be constructed. There are no postulates for things that exist but
aren't constructed, but the existence of a fourth proportional is a good candidate for a such a postulate.
There is a similar situation in modern mathematics with the axiom of choice for set theory. That axiom
says that in certain situations there is at least one set satisfying certain criteria. It does not construct
anything in the usual sense of "construct," and it doesn't even specify a particular set. Although it is
useful in many situations, mathematicians prefer not to use it unless it's necessary.
For this proposition, the assumption of the existence of fourth proportionals is unnecessary.
An alternate proof
Proof: Suppose w:x = y:z. Let n and m be any numbers. Either n < m or not.
Case 1: n < m.
Suppose n(w + x) >=< mx.Subtract nxto get nw >=< (m - n)x.But w:x = y:z,so ny >=< (m - n)z.
Add nzto get n(y + z) >=< mz.
In any case n(w + x) >=< mx implies n(y + z) >=< mz. Therefore (w + x):x = (y + z):z. Q.E.D.
This proposition is used in proposition V.24 and a few in Books VI, X, XII, and XIII.
Book V introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 19
If a whole is to a whole as a part subtracted is to a part subtracted, then the remainder is also
to the remainder as the whole is to the whole.
Let the whole AB be to the whole CD as the part AE subtracted is to the part CF
subtracted.
And, since the magnitudes are proportional taken jointly, they are also proportional
V.17
taken separately, that is, BE is to EA as DF is to CF, and, alternately, BE is to DF as V.16
EA is to FC.
Therefore the remainder EB is also to the remainder FD as the whole AB is to the V.11
whole CD.
Therefore if a whole is to a whole as a part subtracted is to a part subtracted, then the remainder is
also to the remainder as the whole is to the whole.
Q.E.D.
Corollary
V.Def.16
From this it is manifest that, if magnitudes are proportional taken jointly, then they
are also proportional in conversion.
This proposition says that if (u + v):(x + y) equals v:y, then it also equals u:x.
The transformations of proportions taken jointly, taken separately, and in conversion are summarized in
the Guide for V.Def.14.
The magnitudes in this proposition must all be of the same kind, but those in the corollary can be of two
different kinds. Thus, the corollary is out of place. It should probably be after the last proposition since it
follows from the previous two propositions by inversion. As Heiberg and Heath agree, the corollary was
probably interpolated before Theon's time.
This proposition is used in V.25 and a few propositions in Book X. The corollary is used once in each of
Books VI and XIII and fairly often in Book X.
Book V introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 20
If there are three magnitudes, and others equal to them in multitude, which taken two and two
are in the same ratio, and if ex aequali the first is greater than the third, then the fourth is also
greater than the sixth; if equal, equal, and; if less, less.
Let there be three magnitudes A, B, and C, and others D, E, and F equal to
them in multitude, which taken two and two are in the same ratio, so that A is
to B as D is to E, and B is to C as E is to F.
I say that D is also greater than F; if A equals C, equal; and, if less, less.
But, of magnitudes which have a ratio to the same, that which has a greater V.10
ratio is greater, therefore D is greater than F.
Similarly we can prove that, if A equals C, then D also equals F, and if less,
less.
Therefore, if there are three magnitudes, and others equal to them in multitude, which taken two and
two are in the same ratio, and if ex aequali the first is greater than the third, then the fourth is also
greater than the sixth; if equal, equal, and; if less, less.
Q.E.D.
Book V introduction
© 1996
D.E.Joyce
Clark University
Proposition 20
If there are three magnitudes, and others equal to them in multitude, which taken two and two
are in the same ratio, and if ex aequali the first is greater than the third, then the fourth is also
greater than the sixth; if equal, equal, and; if less, less.
Let there be three magnitudes A, B, and C, and others D, E, and F equal to them
in multitude, which taken two and two are in the same ratio, so that A is to B as
D is to E, and B is to C as E is to F.
I say that D is also greater than F; if A equals C, equal; and, if less, less.
But, of magnitudes which have a ratio to the same, that which has a greater ratio V.10
is greater, therefore D is greater than F.
Similarly we can prove that, if A equals C, then D also equals F, and if less, less.
Therefore, if there are three magnitudes, and others equal to them in multitude, which taken two and
two are in the same ratio, and if ex aequali the first is greater than the third, then the fourth is also
greater than the sixth; if equal, equal, and; if less, less.
Q.E.D.
Book V introduction
© 1996
D.E.Joyce
Clark University
Proposition 21
If there are three magnitudes, and others equal to them in multitude, which taken two and two
together are in the same ratio, and the proportion of them is perturbed, then, if ex aequali the
first magnitude is greater than the third, then the fourth is also greater than the sixth; if equal,
equal; and if less, less.
Let there be three magnitudes A, B, and C, and others D, E, and F equal to them
in multitude, which taken two and two are in the same ratio, and let the V.Def.18
proportion of them be perturbed, so that A is to B as E is to F, and B is to C as D
is to E.
I say that D is also greater than F; if A equals C, equal; and if less, less.
But that to which the same has a greater ratio is less, therefore F is less than D, V.10
therefore D is greater than F.
Similarly we can prove that, if A equals C, then D also equals F; and if less, less.
Therefore, if there are three magnitudes, and others equal to them in multitude, which taken two and
two together are in the same ratio, and the proportion of them is perturbed, then, if ex aequali the first
magnitude is greater than the third, then the fourth is also greater than the sixth; if equal, equal; and
if less, less.
Q.E.D.
This proposition is in preparation for V.23. Both this and V.23 rely on treating V.Def.4 as an axiom of
comparability.
Book V introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 22
If there are any number of magnitudes whatever, and others equal to them in multitude, which
taken two and two together are in the same ratio, then they are also in the same ratio ex
aequali.
Let there be any number of magnitudes A, B, and C, and others D, E, and F equal to them in
multitude, which taken two and two together are in the same ratio, so that A is to B as D is
to E, and B is to C as E is to F.
I say that they are also in the same ratio ex aequali, that is, A is to C as D is to F. V.Def.17
Take equimultiples G and H of A and D, and take other, arbitrary, equimultiples K and L of
B and E, and, further, take other, arbitrary, equimultiples M and N of C and F.
Then, since A is to B as D is to E, and of A and D equimultiples G and H have been taken, V.4
and of B and E other, arbitrary, equimultiples K and L, therefore G is to K as H is to L.
Since, then, there are three magnitudes G, K, and M, and others H, L, and N equal to them in
multitude, which taken two and two together are in the same ratio, therefore, ex aequali, if V.20
G is in excess of M, H is also in excess of N; if equal, equal; and if less, less.
And G and H are equimultiples of A and D, and M and N other, arbitrary, equimultiples of C
and F.
Therefore A is to C as D is to F. V.Def.5
Therefore, if there are any number of magnitudes whatever, and others equal to them in multitude,
which taken two and two together are in the same ratio, then they are also in the same ratio ex
aequali.
Q.E.D.
The general statement for this proposition is that for magnitudes x1, x2, ... , and x of one kind, and
n
magnitudes y1, y2, ... , and y of the same or another kind, if x1:x2 = y1:y2, x2:x3 = y2:y3, ... , and x :
n n-1
xn = y :y , then x1:xn = y1:yn.
n-1 n
The proof builds on proposition V.20. Assume A:B = D:E, and B:C = E:F. To show A:C = D:F.
Let n, m, and k be three numbers. By V.4, nA:mB = nD:mE, and mA:kC = mD:kF. By V.20,
This proposition can also be proved directly from the definition Def.V.5 very easily.
The analogous proposition for ratios of numbers is given in proposition VII.14. The proof given there
works for magnitudes as well, but they all have to be of the same kind.
This proposition is used in V.24 and several propositions in Books VI, X, and XII.
Book V introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 23
If there are three magnitudes, and others equal to them in multitude, which taken two and two
together are in the same ratio, and the proportion of them be perturbed, then they are also in
the same ratio ex aequali.
Let there be three magnitudes A, B, and C, and others D, E, and F, equal to them in
multitude, which, taken two and two together, are in the same proportion, and let the V.Def.18
proportion of them be perturbed, so that A is to B as E is to F, and B is to C as D is to E.
I say that A is to C as D is to F.
Then, since G and H are equimultiples of A and B, and parts have the same ratio as their V.15
multiples, therefore A is to B as G is to H.
And, since H and K are equimultiples of B and D, and parts have the same ratio as their V.15
equimultiples, therefore B is to D as H is to K.
Since, then, there are three magnitudes G, H, and L, and others equal to them in multitude
K, M, and N, which taken two and two together are in the same ratio, and the proportion of V.21
them is perturbed, therefore, ex aequali, if G is in excess of L, K is also in excess of N; if
equal, equal; and if less, less.
Therefore A is to C as D is to F. V.Def.5
Therefore, if there are three magnitudes, and others equal to them in multitude, which taken two and
two together are in the same ratio, and the proportion of them be perturbed, then they are also in the
same ratio ex aequali.
Q.E.D.
This proposition says that when a, b, and c are of one kind, and d, e, and f are of the same or another
kind, if a:b = e:f and b:c = d:e, then a:c = d:f.
The proof given here uses proposition V.16 and alternate ratios, and that means it only applies when all
six magnitudes are of the same kind. It is also rather clumsy, since it uses V.15 and V.11 instead of V.4
as the previous proposition V.22 did. It doesn't seem likely that this proof would be written when the
better proof of V.22 could serve as a guide, so it seems likely that V.4 was inserted later and an older
proof of V.22 was cleaned up, but that of V.23 wasn't for some reason such as its relative unimportance.
After all, it is not used in the rest of the Elements.
Let n and m be two arbitrary numbers. By V.15, both A:B = nA:nB, and E:F = mE:mF. Therefore,
by V.11, nA:nB = mE:mF. (The last two sentences would reduce to one with V.4.)
Using alternation V.16 on the other proportion B:C = D:E yields B:D = C:E (but that requires that
all the magnitudes are of the same kind).
For similar reasons nB:nD = B:D = C:E = mC:mE. Therefore, alternately, nB:mC = nD:mE.
Now use V.21 on the two proportions nB:mC = nD:mE and nA:nB = mE:mF, to conclude
Although the last proposition on proportions ex aequali did not depend on treating V.Def.4 as an axiom
of comparability, this proposition on perturbed proportions ex aequali does. For a counterexample
involving infinitesimals, take a = c = e, b = e = (a + y), and f = (a + 2y), where y is infinitesimal with
respect to a.
Book V introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 24
If a first magnitude has to a second the same ratio as a third has to a fourth, and also a fifth
has to the second the same ratio as a sixth to the fourth, then the sum of the first and fifth has
to the second the same ratio as the sum of the third and sixth has to the fourth.
Let a first magnitude AB have to a second C the same ratio as a third DE has to a
fourth F, and let also a fifth BG have to the second C the same ratio as a sixth EH
has to the fourth F.
I say that the sum of the first and fifth, AG, has to the second C the same ratio as
the sum of the third and sixth, DH, has to the fourth F.
Since BG is to C as EH is to F, V.7.Cor
inversely, C is to BG as F is to EH.
Then, since AB is to C as DE is to F,
and C is to BG as F is to EH, V.22
therefore, ex aequali, AB is to BG as
DE is to EH.
And, since the magnitudes are proportional taken separately, they are also V.18
proportional taken jointly, therefore AG is to GB as DH is to HE.
Therefore, if a first magnitude has to a second the same ratio as a third has to a fourth, and also a
fifth has to the second the same ratio as a sixth to the fourth, then the sum of the first and fifth has to
the second the same ratio as the sum of the third and sixth has to the fourth.
Q.E.D.
This proposition says that if u:v = w:x and y:v = z:x, then (u + y):v = (w + z):x.
Although the proposition is stated using the antecedent terms of the proportions, by inversion it applies
to the consequent terms as well.
Book V introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 25
If four magnitudes are proportional, then the sum of the greatest and the least is greater than
the sum of the remaining two.
Let the four magnitudes AB, CD, E, and F be proportional so that AB is to CD as E is to F, and
let AB be the greatest of them and F the least.
I say that the sum of AB and F is greater than the sum of CD and E.
And since the whole AB is to the whole CD as the part AG subtracted is to the part CH
subtracted, therefore the remainder GB is also to the remainder HD as the whole AB is to the V.19
whole CD.
But AB is greater than CD, therefore GB is also greater than HD. (V.14)
And, since AG equals E, and CH equals F, therefore the sum of AG and F equals the sum of
CH and E.
And if, GB and HD being unequal, and GB greater, the sum of AG and F is added to GB, and
the sum of CH and E is added to HD, it follows that the sum of AB and F is greater than the
sum of CD and E.
Therefore, if four magnitudes are proportional, then the sum of the greatest and the least is greater
than the sum of the remaining two.
Q.E.D.
This proposition says that if w:x = y:z and w is the greatest of the four magnitudes while z is the least,
then w + z > x + y. All four magnitudes must be of the same kind.
This proposition is not used in the rest of the Elements but is an end in itself.
A special case of it is when the middle terms are the same: x:y = y:z. In that case y is the mean
proportional, equivalent to the geometric mean for real numbers and described as the square root of the
product xz. The conclusion of the proposition, after dividing by 2, says (x + z)/2 > y. The arithmetic
mean, or average, of two magnitudes is half their sum. Thus, this proposition gives as a corollary
The arithmetic mean of two magnitudes is less than their geometric mean.
Book V introduction
© 1996, 1997
D.E.Joyce
Clark University
Table of contents
● definitions (4)
● propositions (33)
Definitions
Definition 1.
Similar rectilinear figures are such as have their angles severally equal and the sides about the
equal angles proportional.
Definition 2.
Two figures are reciprocally related when the sides about corresponding angles are reciprocally
proportional.
Definition 3.
A straight line is said to have been cut in extreme and mean ratio when, as the whole line is to the
greater segment, so is the greater to the less.
Definition 4.
The height of any figure is the perpendicular drawn from the vertex to the base.
Propositions
Proposition 1.
Triangles and parallelograms which are under the same height are to one another as their bases.
Proposition 2.
If a straight line is drawn parallel to one of the sides of a triangle, then it cuts the sides of the
triangle proportionally; and, if the sides of the triangle are cut proportionally, then the line joining
the points of section is parallel to the remaining side of the triangle.
Proposition 3.
If an angle of a triangle is bisected by a straight line cutting the base, then the segments of the
base have the same ratio as the remaining sides of the triangle; and, if segments of the base have
the same ratio as the remaining sides of the triangle, then the straight line joining the vertex to the
point of section bisects the angle of the triangle.
Proposition 4.
In equiangular triangles the sides about the equal angles are proportional where the
corresponding sides are opposite the equal angles.
Proposition 5.
If two triangles have their sides proportional, then the triangles are equiangular with the equal
angles opposite the corresponding sides.
Proposition 6.
If two triangles have one angle equal to one angle and the sides about the equal angles
proportional, then the triangles are equiangular and have those angles equal opposite the
corresponding sides.
Proposition 7.
If two triangles have one angle equal to one angle, the sides about other angles proportional, and
the remaining angles either both less or both not less than a right angle, then the triangles are
equiangular and have those angles equal the sides about which are proportional.
Proposition 8.
If in a right-angled triangle a perpendicular is drawn from the right angle to the base, then the
triangles adjoining the perpendicular are similar both to the whole and to one another.
Corollary. If in a right-angled triangle a perpendicular is drawn from the right angle to the base,
then the straight line so drawn is a mean proportional between the segments of the base.
Proposition 9.
To cut off a prescribed part from a given straight line.
Proposition 10.
To cut a given uncut straight line similarly to a given cut straight line.
Proposition 11.
To find a third proportional to two given straight lines.
Proposition 12.
To find a fourth proportional to three given straight lines.
Proposition 13.
To find a mean proportional to two given straight lines.
Proposition 14.
In equal and equiangular parallelograms the sides about the equal angles are reciprocally
proportional; and equiangular parallelograms in which the sides about the equal angles are
reciprocally proportional are equal.
Proposition 15.
In equal triangles which have one angle equal to one angle the sides about the equal angles are
reciprocally proportional; and those triangles which have one angle equal to one angle, and in
which the sides about the equal angles are reciprocally proportional, are equal.
Proposition 16.
If four straight lines are proportional, then the rectangle contained by the extremes equals the
rectangle contained by the means; and, if the rectangle contained by the extremes equals the
rectangle contained by the means, then the four straight lines are proportional.
Proposition 17.
If three straight lines are proportional, then the rectangle contained by the extremes equals the
square on the mean; and, if the rectangle contained by the extremes equals the square on the
mean, then the three straight lines are proportional.
Proposition 18.
To describe a rectilinear figure similar and similarly situated to a given rectilinear figure on a
Proposition 19.
Similar triangles are to one another in the duplicate ratio of the corresponding sides.
Corollary. If three straight lines are proportional, then the first is to the third as the figure
described on the first is to that which is similar and similarly described on the second.
Proposition 20.
Similar polygons are divided into similar triangles, and into triangles equal in multitude and in
the same ratio as the wholes, and the polygon has to the polygon a ratio duplicate of that which
the corresponding side has to the corresponding side.
Corollary. Similar rectilinear figures are to one another in the duplicate ratio of the
corresponding sides.
Proposition 21.
Figures which are similar to the same rectilinear figure are also similar to one another.
Proposition 22.
If four straight lines are proportional, then the rectilinear figures similar and similarly described
upon them are also proportional; and, if the rectilinear figures similar and similarly described
upon them are proportional, then the straight lines are themselves also proportional.
Proposition 23.
Equiangular parallelograms have to one another the ratio compounded of the ratios of their sides.
Proposition 24.
In any parallelogram the parallelograms about the diameter are similar both to the whole and to
one another.
Proposition 25.
To construct a figure similar to one given rectilinear figure and equal to another.
Proposition 26.
If from a parallelogram there is taken away a parallelogram similar and similarly situated to the
whole and having a common angle with it, then it is about the same diameter with the whole.
Proposition 27.
Of all the parallelograms applied to the same straight line falling short by parallelogrammic
figures similar and similarly situated to that described on the half of the straight line, that
parallelogram is greatest which is applied to the half of the straight line and is similar to the
difference.
Proposition 28.
To apply a parallelogram equal to a given rectilinear figure to a given straight line but falling
short by a parallelogram similar to a given one; thus the given rectilinear figure must not be
greater than the parallelogram described on the half of the straight line and similar to the given
parallelogram.
Proposition 29.
To apply a parallelogram equal to a given rectilinear figure to a given straight line but exceeding
it by a parallelogram similar to a given one.
Proposition 30.
To cut a given finite straight line in extreme and mean ratio.
Proposition 31.
In right-angled triangles the figure on the side opposite the right angle equals the sum of the
similar and similarly described figures on the sides containing the right angle.
Proposition 32.
If two triangles having two sides proportional to two sides are placed together at one angle so that
their corresponding sides are also parallel, then the remaining sides of the triangles are in a
straight line.
Proposition 33.
Angles in equal circles have the same ratio as the circumferences on which they stand whether
they stand at the centers or at the circumferences.
Proposition VI.1 is the basis for the entire of Book VI except the last proposition VI.33. Only these two
propositions directly use the definition of proportion in Book V. Proposition VI.1 constructs a
proportion between lines and figures while VI.33 constructs a proportion between angles and
circumferences. The intervening propositions use other properties of proportions developed in Book V,
but they do not construct new proportions using the definition of proportion.
Book VI
introduction
© 1996
D.E.Joyce
Clark University
Definition 1
Similar rectilinear figures are such as have their angles severally equal and the sides about the
equal angles proportional.
The words in this definition do not quite express its entire intent. It is apparent from its use that the
notion of similarity assumes a specific correspondence of consecutive vertices and sides. Consider, for
instance, pentagons.
In order for the pentagons ABCDE and FGHKL to be similar, it is required that
1. corresponding angles taken in order are equal, that is, A = F, B = G, C = H, D = K, and E = L, and
2. the sides about their equal angles are proportional in the same order:
EA:AB = LF:FG,
AB:BC = FG:GH,
BC:CD = GH:HK,
CD:EF = HK:KL, and
EA:AB = KL:LF.
It wouldn't be allowed, for instance, if the angles of one figure equalled the angles of the other, but in
some haphazard order. And it wouldn't be allowed for the orders of the terms in the proportions to be
permuted, or inverted, for instance, the second proportion could not be AB:BC = GH:FG.
Propositions VI.4 and VI.5 give two criteria for two triangles to be similar. Proposition VI.4 says that
condition 1 implies similarity, while VI.5 says condition 2 implies similarity. Proposition VI.6 is a side-
angle-side similarity theorem, and VI.7 is a side-side-angle similarity theorem. Many of the other
propositions in this and later books involve similarity in one way or another.
© 1997, 2002
D.E.Joyce
Clark University
Definition 2
Two figures are reciprocally related when the sides about corresponding angles are reciprocally
proportional.
This isn't the actual definition that appears, but an approximation of its intent. A literal translation is
incomplete, and this definition may have been added after Euclid.
Book VI introduction
© 1997
D.E.Joyce
Clark University
Definition 3
A straight line is said to have been cut in extreme and mean ratio when, as the whole line is to the
greater segment, so is the greater to the less.
The line AB is cut in extreme and mean ratio at C since AB:AC = AC:
CB.
A construction to cut a line in this manner first appeared in Book II, proposition II.11. Of course that
was before ratios were defined, and there an equivalent condition was stated in terms of rectangles,
namely, that the square on AC equal the rectangle AB by BC. That construction was later used in Book
IV in order to construct regular pentagons and 15-sided polygons (propositions IV.10 through 12 and
16).
Now that the theory of ratios and proportions has been developed, it is time to define this section as a
ratio, rather than using rectangles. An alternate construction is given in proposition VI.30.
Book VI introduction
© 1997
D.E.Joyce
Clark University
http://aleph0.clarku.edu/~djoyce/java/elements/bookVI/defVI3.html5/26/2008 3:30:14 PM
Euclid's Elements, Book VI, Definition 4
Definition 4
The height of any figure is the perpendicular drawn from the vertex to the base.
If the figure is a triangle, and one side has been declared the base,
then the height is the expected line, the line drawn from the
opposite vertex perpendicular to the base. If the figure is a
parallelogram, and one side has been declared the base, then the
height may be taken to be a perpendicular from either of the two
vertices not on the base.
In the later books on solid geometry, other figures also can have bases and heights such as
parallelepipeds, pyramids, prisms, cones, and cylinders.
Different sides of a figure may be selected as the base depending on the application. In proposition
XI.39 there are two triangular prisms. A triangle is chosen taken to be the base of one, while the base of
the other is a parallelogram. The height of the first is a perpendicular drawn between two triangular
opposite faces, but the height of the other is a perpendicular drawn between the parallelogram taken as
the base and the opposite parallelogram.
Book VI introduction
© 1997
D.E.Joyce
Clark University
Proposition 1
Triangles and parallelograms which are under the same height are to one another as their
bases.
Let ACB and ACD be triangles, and let CE and CF be parallelograms under the same height.
I say that the base CB is to the base CD as the triangle ACB is to the triangle ACD, and as the
parallelogram CE is to the parallelogram CF.
Produce BD in both directions to the points H and L. Make any number of straight lines BG and
GH equal to the base CB, and any number of straight lines DK and KL equal to the base CD. Join I.3
AG, AH, AK, and AL.
Then, since CB, BG, and GH equal one another, the triangles ACB, ABG, and AGH also I.38
equal one another.
Therefore, whatever multiple the base CH is of the base CB, the triangle ACH is also that
multiple of the triangle ACB.
For the same reason, whatever multiple the base CL is of the base CD, the triangle ACL is
also that multiple of the triangle ACD. And, if the base CH equals the base CL, then the I.38
triangle ACH also equals the triangle ACL; if the base CH is in excess of the base CL, the
triangle ACH is also in excess of the triangle ACL; and, if less, less.
Thus, there being four magnitudes, namely two bases CB and CD, and two triangles ACB
and ACD, equimultiples have been taken of the base CB and the triangle ACB, namely the
base CH and the triangle ACH, and other, arbitrary, equimultiples of the base CD and the
triangle ADC, namely the base CL and the triangle ACL, and it has been proved that, if the V.Def.5
base CH is in excess of the base CL, the triangle ACH is also in excess of the triangle ACL; if
equal, equal; and, if less, less. Therefore the base CB is to the base CD as the triangle ACB is
to the triangle ACD.
Next, since the parallelogram CE is double the triangle ACB, and the parallelogram FC is
I.41
double the triangle ACD, and parts have the same ratio as their equimultiples, therefore the V.15
triangle ACB is to the triangle ACD as the parallelogram CE is to the parallelogram FC.
Since, then, it was proved that the base CB is to CD as the triangle ACB is to the triangle
ACD, and the triangle ACB is to the triangle ACD as the parallelogram CE is to the V.11
parallelogram CF, therefore also the base CB is to the base CD as the parallelogram CE is to
the parallelogram FC.
Therefore, triangles and parallelograms which are under the same height are to one another as their
bases.
Q.E.D.
In a more proper setting out of the proposition, the triangles under the same height would not have a
common side, and the parallelograms would not have a common base and side with the triangles. Since
triangles on equal bases and in the same parallels are equal (I.36), and parallelograms on equal bases and
in the same parallels are equal (I.35), and equals may be substituted in proportions (V.7), Euclid's
simplified setting out is sufficient. Nonetheless, a proper setting out does not require a more complicated
proof.
The goal of the proof is to show that three ratios, namely the ratio of the lines CB to CD, the ratio of the
triangles ACB to ACD, and the ratio of the parallelograms CE to CF, are all the same ratio. That is
The first stage of the proof shows that CB:CD = ACB:ACD. By the definition of proportion, V.Def.5,
that means for any number m and any number n that
Note that Euclid takes both m and n to be 3 in his proof. Now m BC equals the line CH, n CD equals the
line CL, m ABC equals the triangle ACH, and n ACD equals the triangle ACL. So what has to be shown
is that
But that follows from proposition I.38. So the first stage of the proof is complete.
The second stage is easier. Since the parallelograms are twice the triangles, they also have the same
ratio.
Other propositions that state fundamental proportions use the same outline for their proofs. Proposition
VI.33: arcs of circles are proportional to angles on which they stand; XI.25: parallelepipeds are
proportional to their bases; and XII.13: cylinders are proportional to their axes.
Heath remarked that "some American and German text-books adopt the less rigorous method of
appealing to the theory of limits" for the foundation for the theory of proportion used here in geometry.
Heath preferred Eudoxus' theory of proportion in Euclid's Book V as a foundation.
It is remarkable how much mathematics has changed over the last century. In the beginning of the 20th
century Heath could still gloat over the superiority of synthetic geometry, although he may have been
one of the last to do so. Now, in the 21st century, synthetic geometry has receded into near oblivion
while analysis, based on various concepts of limits, is preeminent.
It took some time to find a foundation for mathematical analysis as solid, or more solid, than geometry.
In the 17th century, the time of the creation of differential and integral calculus, geometry was seen as
the most dependable justification for calculus. In the first half of the 19th century, the concept of limit
was clarified and limits became the foundation of mathematical analysis. Heath's complaint would have
been valid then since the theory of real numbers was still without any foundation except a geometric
one, which, ultimately was based on Eudoxus' theory of proportion in Euclid's Book V. In the later 19th
century Weierstrass, Cantor, and Dedekind succeeded in founding the theory of real numbers on that of
natural numbers and a bit of set theory, so that by the beginning of the 20th century, there was a modern
foundation for mathematical analysis. All the same, this new foundation could still be called Eudoxus'
since the modern definition of real number is the same as his, but in a modern guise.
This is one of the most used propositions in the Elements. It is used frequently in Book VI starting with
the next proposition, dozens of times in Book X, and and a few times in Books XI and XIII.
Book VI introduction
© 1996, 2002
D.E.Joyce
Clark University
Proposition 2
If a straight line is drawn parallel to one of the sides of a triangle, then it cuts the sides of the
triangle proportionally; and, if the sides of the triangle are cut proportionally, then the line
joining the points of section is parallel to the remaining side of the triangle.
Let DE be drawn parallel to BC, one of the sides of the triangle ABC.
But the triangle BDE is to ADE as BD is to AD, for, being under the same height, the VI.1
perpendicular drawn from E to AB, they are to one another as their bases.
Next, let the sides AB and AC of the triangle ABC be cut proportionally, so that BD is to AD as
CE is to AE. Join DE.
With the same construction, since BD is to AD as CE is to AE, but BD is to AD as the triangle VI.1
BDE is to the triangle ADE, and CE is to AE as the triangle CDE is to the triangle ADE,
therefore the triangle BDE is to the triangle ADE as the triangle CDE is to the triangle ADE. V.11
Therefore each of the triangles BDE and CDE has the same ratio to ADE.
Therefore the triangle BDE equals the triangle CDE, and they are on the same base DE. V.9
But equal triangles which are on the same base are also in the same parallels. I.39
Therefore, if a straight line is drawn parallel to one of the sides of a triangle, then it cuts the sides of
the triangle proportionally; and, if the sides of the triangle are cut proportionally, then the line
joining the points of section is parallel to the remaining side of the triangle.
Q.E.D.
Euclid prefers to prove a pair of converses in two stages, but in some propositions, as this one, the
proofs in the two stages are almost inverses of each other, so both could be proved at once.
In this proposition we have a given triangle ABC and a line DE joining a point D on the side BC to a
point E on the side AC. The claim is that
Hence,
By propositions V.7 and V.9 the latter condition is equivalent to BDE = CDE, and that, in turn, by
propositions I.37 and I.39 is equivalent to DE || BC.
Note
It should be noted that a proportion such as BD:AD = AE:CE is not intended. In that case the sides are
cut proportionally, but the correspondence is not the intended one.
This proposition is frequently used in the rest of Book VI starting with the next proposition. It is also
used in Books XI and XII.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 3
If an angle of a triangle is bisected by a straight line cutting the base, then the segments of the
base have the same ratio as the remaining sides of the triangle; and, if segments of the base
have the same ratio as the remaining sides of the triangle, then the straight line joining the
vertex to the point of section bisects the angle of the triangle.
Let ABC be a triangle, and let the angle BAC be bisected by the straight line AD.
Then, since the straight line AC falls upon the parallels AD I.29
and EC, the angle ACE equals the angle CAD.
Again, since the straight line BAE falls upon the parallels
AD and EC, the exterior angle BAD equals the interior angle I.29
AEC.
But the angle ACE was also proved equal to the angle BAD, therefore the angle ACE also I.6
equals the angle AEC, so that the side AE also equals the side AC.
And, since AD is parallel to EC, one of the sides of the triangle BCE, therefore, proportionally VI.2
DB is to DC as AB is to AE.
V.9
Therefore AC equals AE, so that the angle AEC also equals the angle ACE. I.5
But the angle AEC equals the exterior angle BAD, and the angle ACE equals the alternate angle I.29
CAD, therefore the angle BAD also equals the angle CAD.
Therefore, if an angle of a triangle is bisected by a straight line cutting the base, then the segments of
the base have the same ratio as the remaining sides of the triangle; and, if segments of the base have
the same ratio as the remaining sides of the triangle, then the straight line joining the vertex to the
point of section bisects the angle of the triangle.
Q.E.D.
This proposition characterizes an angle bisector of an angle in a triangle as the line that partitions the
base into parts proportional to the adjacent sides. The second part of the statement of the proposition is
the converse of the first part of the statement. The proof relies on basic properties of triangles and
parallel lines developed in Book I along with the result of the previous proposition VI.2.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 4
In equiangular triangles the sides about the equal angles are proportional where the
corresponding sides are opposite the equal angles.
Let ABC and DCE be equiangular triangles having the angle ABC equal to the angle DCE,
the angle BAC equal to the angle CDE, and the angle ACB equal to the angle CED.
I say that in the triangles ABC and DEC the sides about the equal angles are proportional
where the corresponding sides are opposite the equal angles.
Then, since the sum of the angles ABC and ACB is less than
two right angles, and the angle ACB equals the angle DEC, I.17
therefore the sum of the angles ABC and DEC is less than
two right angles. Therefore BA and ED, when produced, will I.Post.5
meet. Let them be produced and meet at F.
Now, since the angle DCE equals the angle ABC, DC is parallel to FB. Again, since the angle I.28
ACB equals the angle DEC, AC is parallel to FE.
Therefore FACD is a parallelogram, therefore FA equals DC, and AC equals FD. I.34
Therefore, in equiangular triangles the sides about the equal angles are proportional where the
corresponding sides are opposite the equal angles.
Q.E.D.
In the enunciation of this proposition the term "equiangular triangles" refers to two triangles whose
corresponding angles are equal, not to two triangles each of which is equiangular (equilateral).
Euclid has placed the triangles in particular positions in order to employ this particular proof. Such
positioning is common in Book VI and is easily justified.
This proposition implies that equiangular triangles are similar, a fact proved in detail in the proof of
proposition VI.8. It also implies that triangles similar to the same triangle are similar to each other, also
proved in detail in VI.8. The latter statement is generalized in VI.21 to rectilinear figures in general.
This proposition is frequently used in the rest of Book VI starting with the next proposition, its converse.
It is also used in Books X through XIII.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 5
If two triangles have their sides proportional, then the triangles are equiangular with the equal
angles opposite the corresponding sides.
Let ABC and DEF be two triangles having their sides proportional, so that AB is to BC as DE is to EF,
BC is to CA as EF is to FD, and further BA is to AC as ED is to DF.
I say that the triangle ABC is equiangular with the triangle DEF where the equal angles are opposite
the corresponding sides, namely the angle ABC equals the angle DEF, the angle BCA equals the angle
EFD, and the angle BAC equals the angle EDF.
Construct the angle FEG equal to the angle CBA and the angle EFG equal to the angle BCA on I.23
the straight line EF and at the points E and F on it. Therefore the remaining angle at A equals
the remaining angle at G. I.32
Therefore in the triangles ABC and GEF the sides about the equal angles are proportional where VI.4
the corresponding sides are opposite the equal angles, therefore AB is to BC as GE is to EF.
Therefore each of the straight lines DE and GE has the same ratio to EF, therefore DE equals V.9
GE.
Then since DE equals GE, and EF is common, the two sides DE and EF equal the two sides GE I.8
and EF, and the base DF equals the base GF, therefore the angle DEF equals the angle GEF,
and the triangle DEF equals the triangle GEF, and the remaining angles equal the remaining
I.4
angles, namely those opposite the equal sides.
Therefore the angle DFE also equals the angle GFE, and the angle EDF equals the angle EGF.
And, since the angle DEF equals the angle GEF, and the angle GEF equals the angle ABC,
therefore the angle ABC also equals the angle DEF.
For the same reason the angle ACB also equals the angle DFE, and further, the angle at A
equals the angle at D, therefore the triangle ABC is equiangular with the triangle DEF.
Therefore, if two triangles have their sides proportional, then the triangles are equiangular with the
equal angles opposite the corresponding sides.
Q.E.D.
Of course, this proposition is the converse of the previous. We now have two characterizations of similar
triangles, either as equiangular triangles or as triangles with proportional sides. The next two
propositions give two more characterizations corresponding to characterizations of congruent triangles.
As in VI.2, a certain order is assumed for the proportionality. It is not intended, for instance, that AB:
BC = DE:EF while BC:CA = FD:EF. See the remark about VI.Def.1.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 6
If two triangles have one angle equal to one angle and the sides about the equal angles
proportional, then the triangles are equiangular and have those angles equal opposite the
corresponding sides.
Let ABC and DEF be two triangles having one angle BAC equal to one angle EDF and the sides about
the equal angles proportional, so that BA is to AC as ED is to DF.
I say that the triangle ABC is equiangular with the triangle DEF, and has the angle ABC equal to the
angle DEF, and the angle ACB equal to the angle DFE.
On the straight line DF and at the points D and F on it, construct the angle FDG equal to either I.23
of the angles BAC or EDF, and the angle DFG equal to the angle ACB.
Therefore the remaining angle at B equals the remaining angle at G. Therefore the triangle ABC I.32
is equiangular with the triangle DGF.
Therefore, proportionally BA is to AC as GD is to DF. VI.4
Therefore ED equals GD. And DF is common, therefore the two sides ED and DF equal the two V.9
sides GD and DF, and the angle EDF equals the angle GDF, therefore the base EF equals the
base GF, the triangle DEF equals the triangle DGF, and the remaining angles equal the
I.4
remaining angles, namely those opposite the equal sides.
Therefore the angle DFG equals the angle DFE, and the angle DGF equals the angle DEF.
But the angle DFG equals the angle ACB, therefore the angle ACB also equals the angle DFE.
And, by hypothesis, the angle BAC also equals the angle EDF, therefore the remaining angle at B
also equals the remaining angle at E. Therefore the triangle ABC is equiangular with the triangle I.32
DEF.
Therefore, if two triangles have one angle equal to one angle and the sides about the equal angles
proportional, then the triangles are equiangular and have those angles equal opposite the
corresponding sides.
Q.E.D.
Here's a summary of the proof. Construct a triangle DGF equiangular with triangle ABC. Then triangle
DGF is similar to triangle ABC ( VI.4), and that gives us the proportion
BA:AC = GD:DF.
BA:AC = ED:DF,
GD:DF = ED:DF
(V.11), from which it follows that GD = ED (V.9). Therefore triangles DEF and DGF are congruent,
and the rest follows easily.
This proposition is used in the proofs of propositions VI.20, VI.32, XII.1, and several times in XII.12.
Book VI introduction
© 1996, 2002
D.E.Joyce
Clark University
Proposition 7
If two triangles have one angle equal to one angle, the sides about other angles proportional,
and the remaining angles either both less or both not less than a right angle, then the triangles
are equiangular and have those angles equal the sides about which are proportional.
Let ABC and DEF be two triangles having one angle equal to one angle, the angle BAC equal
to the angle EDF, the sides about other angles ABC and DEF proportional, so that AB is to BC
as DE is to EF. And, first, each of the remaining angles at C and F less than a right angle.
I say that the triangle ABC is equiangular with the triangle DEF, the angle ABC equals the
angle DEF, and the remaining angle, namely the angle at C, equals the remaining angle, the
angle at F.
Then, since the angle A equals D, and the angle ABG equals the angle DEF, therefore the I.32
remaining angle AGB equals the remaining angle DFE.
But, by hypothesis, DE is to EF as AB is to BC, therefore AB has the same ratio to each of the V.11
straight lines BC and BG. Therefore BC equals BG, so that the angle at C also equals the angle V.9
BGC. I.5
But, by hypothesis, the angle at C is less than a right angle, therefore the angle BGC is also less I.13
than a right angle, so that the angle AGB adjacent to it is greater than a right angle.
And it was proved equal to the angle at F, therefore the angle at F is also greater than a right
angle. But it is by hypothesis less than a right angle, which is absurd.
Therefore the angle ABC is not unequal to the angle DEF. Therefore it equals it.
But the angle at A also equals the angle at D, therefore the remaining angle at C equals the I.32
remaining angle at F.
Next let each of the angles at C and F be supposed not less than a right angle.
I say again that, in this case too, the triangle ABC is equiangular with the triangle DEF.
Thus in the triangle BGC the sum of two angles is not less than two right angles, which is I.17
impossible.
Therefore, once more, the angle ABC is not unequal to the angle DEF. Therefore it equals it.
But the angle at A also equals the angle at D, therefore the remaining angle at C equals the I.32
remaining angle at F.
Therefore, if two triangles have one angle equal to one angle, the sides about other angles
proportional, and the remaining angles either both less or both not less than a right angle, then the
triangles are equiangular and have those angles equal the sides about which are proportional.
Q.E.D.
This is a side-side-angle similarity proposition for triangles. The Elements does not have the analogous
side-side-angle congruence proposition for triangles. See the note on congruence theorems after I.26 for
more about congruence theorems.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 8
If in a right-angled triangle a perpendicular is drawn from the right angle to the base, then the
triangles adjoining the perpendicular are similar both to the whole and to one another.
Let ABC be a right-angled triangle having the angle BAC right, and let AD be drawn
from A perpendicular to BC.
Since the angle BAC equals the angle BDA, for each is right, and the angle at B is
common to the two triangles ABC and DBA, therefore the remaining angle ACB equals I.32
the remaining angle DAB. Therefore the triangle ABC is equiangular with the triangle
DBA.
Therefore BC, which is opposite the right angle in the triangle ABC, is to BA, which is
opposite the right angle in the triangle DBA, as AB, which is opposite the angle at C in VI.4
the triangle ABC, is to DB, which is opposite the equal angle BAD in the triangle DBA,
and also as AC is to DA, which is opposite the angle at B common to the two triangles.
Therefore the triangle ABC is both equiangular to the triangle DBA and has the sides
about the equal angles proportional.
In the same manner we can prove that the triangle DAC is also similar to the triangle
ABC. Therefore each of the triangles DBA and DAC is similar to the whole ABC.
I say next that the triangles DBA and DAC are also similar to one another.
Since the right angle BDA equals the right angle ADC, and moreover the angle DAB
was also proved equal to the angle at C, therefore the remaining angle at B also equals I.32
the remaining angle DAC. Therefore the triangle DBA is equiangular with the triangle
ADC.
Therefore BD, which is opposite the angle DAB in the triangle DBA, is to AD, which is
opposite the angle at C in the triangle DAC equal to the angle DAB, as AD, itself VI.4
which is opposite the angle at B in the triangle DBA, is to CD, which is opposite the
angle DAC in the triangle DAC equal to the angle at B, and also as BA is to AC, these
VI.Def.1
sides opposite the right angles. Therefore the triangle DBA is similar to the triangle
DAC.
Therefore, if in a right-angled triangle a perpendicular is drawn from the right angle to the base, then
the triangles adjoining the perpendicular are similar both to the whole and to one another.
Q.E.D.
Corollary
From this it is clear that, if in a right-angled triangle a perpendicular is drawn from the right angle to
the base, then the straight line so drawn is a mean proportional between the segments of the base.
Essentially, triangles ABC and DBA are equiangular since they are right triangles with a common angle.
Therefore, they are similar. Likewise, triangles ABC and DAC are similar.
Note that Euclid verbosely draws from proposition VI.4 the conclusions that equiangular triangles are
similar and that triangles similar to the same triangle are similar to each other. The general proposition
that figures similar to the same figure are also similar to one another is proposition VI.21. There is no
reason why that proposition could not have been placed before this one.
This proposition may be used to give an alternate proof of proposition I.47. Indeed, Euclid presents such
a proof in the lemma for X.33. That proof is probably older than Euclid's as given in I.47, but Euclid's
proof has the advantage of not being dependent on Eudoxus' theory of proportion in Book V.
This proposition and its corollary are used in propositions VI.13, VI.31, X.33, and often in Book XIII.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 9
Then, since DF is parallel to a side CB of the triangle ABC, therefore, proportionally, AD is to VI.2
DC as AF is to FB.
But DC is double AD, therefore FB is also double AF, therefore AB is triple of AF.
Therefore from the given straight line AB the prescribed third part AF has been cut off.
Q.E.F.
The word "part" in this proposition means submultiple. The problem here is to divide a line AB into
some given number n of equal parts, or actually, to to find just one of these parts. Euclid takes the case
n = 3 in his proof.
Simson complained that proving the general case by using a specific case, the one-third part, "is not at
all like Euclid's manner." But it is very much Euclid's manner throughout books V and VI to prove a
general numerical statement with a specific numerical value.
Al-Nayrizi's construction
This construction is used in a few propositions in Book XIII to find a third or a fifth of a line.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 10
To cut a given uncut straight line similarly to a given cut straight line.
Let AB be the given uncut straight line, and AC the straight line cut at the points D and E, and
let them be so placed as to contain any angle. Join CB, and draw DF and EG through D and E I.31
parallel to CB, and draw DHK through D parallel to AB.
Again, since DF is parallel to a side EG of the triangle AEG, therefore, proportionally, AD is to VI.2
DE as AF is to FG.
Therefore the given uncut straight line AB has been cut similarly to the given cut straight line
AC.
Q.E.F.
In a sense, this proposition is a generalization of the last one VI.9. Prop. VI.9 cut a line into two parts
whose ratio was a given numerical ratio. This proposition cuts a line into two parts whose ratio is a
given ratio of two other lines. Both propositions rely on VI.2 as a basis to make any conclusion about the
This proposition is not used later in the Elements, but it is a basic construction of geometry.
Book VI introduction
© 1996, 2002
D.E.Joyce
Clark University
Proposition 11
Therefore a third proportional CE has been found to two given straight lines AB and AC.
Q.E.F.
If a and b are two magnitudes, then their third proportional is a magnitude c such that a:b = b:c. The
third proportional is needed whenever a duplicate ratio is needed when the ratio itself is known. The
duplicate ratio for a:b is a:c.
This construction is used in propositions VI.19, VI.22, and a few propositions in Book X.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 12
Therefore a fourth proportional HF has been found to the three given straight lines A, B, and C.
Q.E.F.
Descartes (1591-1661) is well known for his coordinate geometry which he and Fermat developed in the
16th century. This subject, also called analytic geometry, places an x-y-coordinate system on a plane so
that a curve in the plane corresponds to an equation in two variables x and y. The usual way this
correspondence is used is to convert a problem in geometry into an algebraic problem about equations.
Descartes was equally interested is using geometry to solve algebraic problems, but using a method
quite distinct from that in the Elements which began in Book II.
His idea was to take an equation in one variable and find a geometric figure which can be used to solve
the equation. The idea wasn't particularly new as even Menaechmus (fl. about 350 B.C.E) had about 50
years before Euclid intersected two parabolas to find cube roots, which are solutions to particular cubic
equations. Furthermore, about 1100, Omar Khayyam solved all cubic equations by means of parabolas
and hyperbolas. But Descartes was systematic and was able to use the relatively recent invention of
symbolic algebra to make more connections.
This proposition is used in the proofs of VI.22, VI.23, and half a dozen propositions in Book X.
Book VI introduction
© 1996, 2002
D.E.Joyce
Clark University
Proposition 13
And, since, in the right-angled triangle ADC, BD has been drawn from the right angle
perpendicular to the base, therefore BD is a mean proportional between the segments of the VI.8,Cor
base, AB and BC.
Therefore a mean proportional BD has been found to the two given straight lines AB and
BC.
Q.E.F.
This construction of the mean proportional was used before in II.4 to find a square equal to a given
rectangle. By proposition VI.17 coming up, the two constructions are equivalent. That is the mean
proportional between two lines is the side of a square equal to the rectangle contained by the two lines.
Algebraically, a : x = x : b if and only if ab = x2. Thus, x is the square root of ab.
When b is taken to have unit length, this construction gives the construction for the square root of a.
This construction is used in the proofs of propositions VI.25, X.27, and X.28.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 14
In equal and equiangular parallelograms the sides about the equal angles are reciprocally
proportional; and equiangular parallelograms in which the sides about the equal angles are
reciprocally proportional are equal.
Let AB and BC be equal and equiangular parallelograms having the angles at B equal, and let I.14
DB and BE be placed in a straight line. Therefore FB and BG are also in a straight line.
I say that, in AB and BC, the sides about the equal angles are reciprocally proportional, that is to
say, DB is to BE as BG is to BF.
Therefore in the parallelograms AB and BC the sides about the equal angles are reciprocally
proportional.
Therefore, in equal and equiangular parallelograms the sides about the equal angles are reciprocally
proportional; and equiangular parallelograms in which the sides about the equal angles are
reciprocally proportional are equal.
Q.E.D.
This proposition is used in the proofs of propositions VI.16, VI.30, and X.22.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 15
In equal triangles which have one angle equal to one angle the sides about the equal angles
are reciprocally proportional; and those triangles which have one angle equal to one angle,
and in which the sides about the equal angles are reciprocally proportional, are equal.
Let ABC and ADE be equal triangles having one angle equal to one angle, namely the angle
BAC equal to the angle DAE.
I say that in the triangles ABC and ADE the sides about the equal angles are reciprocally
proportional, that is to say, that CA is to AD as EA is to AB.
Join BD.
Since then the triangle ABC equals the triangle ADE, and ABD is
another triangle, therefore the triangle ABC is to the triangle ABD V.7
as the triangle ADE is to the triangle ABD.
Therefore in the triangles ABC and ADE the sides about the equal angles are reciprocally
proportional.
Next, let the sides of the triangles ABC and ADE be reciprocally proportional, that is to say, let
AE be to AB as CA is to AD.
Therefore each of the triangles ABC and ADE has the same ratio to ABD.
Therefore, in equal triangles which have one angle equal to one angle the sides about the equal
angles are reciprocally proportional; and those triangles which have one angle equal to one angle,
and in which the sides about the equal angles are reciprocally proportional, are equal.
Q.E.D.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 16
If four straight lines are proportional, then the rectangle contained by the extremes equals the
rectangle contained by the means; and, if the rectangle contained by the extremes equals the
rectangle contained by the means, then the four straight lines are proportional.
Let the four straight lines AB, CD, E, and F be proportional, so that AB is to CD as E is to F.
Draw AG and CH from the points A and C at right angles to the straight lines AB and CD, and I.11
make AG equal to F, and CH equal to E. I.3
Then since AB is to CD as E is to F, while E equals CH, and F equals AG, therefore AB is to V.7
CD as CH is to AG.
Therefore in the parallelograms BG and DH the sides about the equal angles are reciprocally
proportional.
But those equiangular parallelograms in which the sides about the equal angles are reciprocally VI.14
proportional are equal, therefore the parallelogram BG equals the parallelogram DH.
With the same construction, since the rectangle AB by F equals the rectangle CD by E, and the
rectangle AB by F is BG, for AG equals F, and the rectangle CD by E is DH, for CH equals E,
therefore BG equals DH.
And they are equiangular. But in equal and equiangular parallelograms the sides about the VI.14
equal angles are reciprocally proportional.
Therefore, if four straight lines are proportional, then the rectangle contained by the extremes equals
the rectangle contained by the means; and, if the rectangle contained by the extremes equals the
rectangle contained by the means, then the four straight lines are proportional.
Q.E.D.
This proposition is a special case of VI.14. It hardly needs such a protracted proof. It is used
occasionally in Book X, but the special case when the means are equal and the second figure is a square,
as enunciated in the next proposition, is used throughout Book X and frequently in Book XIII.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 17
If three straight lines are proportional, then the rectangle contained by the extremes equals the
square on the mean; and, if the rectangle contained by the extremes equals the square on the
mean, then the three straight lines are proportional.
Let the three straight lines A and B and C be proportional, so that A is to B as
B is to C.
I say that A is to B as B is to C.
With the same construction, since the rectangle A by C equals the square on
B, while the square on B is the rectangle B by D, for B equals D, therefore the
rectangle A by C equals the rectangle B by D.
But, if the rectangle contained by the extremes equals that contained by the VI.16
means, then the four straight lines are proportional.
Therefore A is to B as D is to C.
Therefore, if three straight lines are proportional, then the rectangle contained by the extremes equals
the square on the mean; and, if the rectangle contained by the extremes equals the square on the
mean, then the three straight lines are proportional.
Q.E.D.
This is obviously a special case of the previous proposition. It is used very frequently in Books X and
XIII.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 18
To describe a rectilinear figure similar and similarly situated to a given rectilinear figure on a
given straight line.
Let AB be the given straight line and CE the given rectilinear figure.
It is required to describe on the straight line AB a rectilinear figure similar and similarly
situated to the rectilinear figure CE.
VI.4
Therefore, proportionally, FD is to GB as FC is to GA, and as CD is to AB. V.16
Again, construct the angle BGH equal to the angle DFE, and the angle GBH equal to the I.23
angle FDE, on the straight line BG and at the points B and G on it.
Therefore the remaining angle at E equals the remaining angle at H. Therefore the triangle I.32
FDE is equiangular with the triangle GBH. Therefore, proportionally, FD is to GB as FE is VI.4
to GH, and as ED is to HB. V.16
But it was also proved that FD is to GB as FC is to GA, and as CD is to AB. Therefore FC is V.11
to AG as CD is to AB, and as FE is to GH, and further as ED is to HB.
And, since the angle CFD equals the angle AGB, and the angle DFE equals the angle BGH,
therefore the whole angle CFE equals the whole angle AGH.
For the same reason the angle CDE also equals the angle ABH.
And the angle at C also equals the angle at A, and the angle at E equals the angle at H.
Therefore AH is equiangular with CE, and they have the sides about their equal angles VI.Def.1
proportional. Therefore the rectilinear figure AH is similar to the rectilinear figure CE.
Therefore the rectilinear figure AH has been described similar and similarly situated to the
given rectilinear figure CE on the given straight line AB.
Q.E.F.
It is evident from the diagram, if not from the text, that AB should correspond to CD.
Although the figure has only four sides, it is clear that the method applies to figures with more than four
sides.
This proposition is used in the proofs of propositions VI.22, VI.25, and VI.28, and the corollary is used
in XII.17.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 19
Similar triangles are to one another in the duplicate ratio of the corresponding sides.
Let ABC and DEF be similar triangles having the angle at B equal to the angle at E, and V.Def.11
such that AB is to BC as DE is to EF, so that BC corresponds to EF.
I say that the triangle ABC has to the triangle DEF a ratio duplicate of that which BC has to
EF.
Therefore in the triangles ABG and DEF the sides about the equal angles are reciprocally
proportional.
But those triangles which have one angle equal to one angle, and in which the sides about
the equal angles are reciprocally proportional, are equal. Therefore the triangle ABG equals VI.15
the triangle DEF.
Now since BC is to EF as EF is to BG, and, if three straight lines are proportional, the first
has to the third a ratio duplicate of that which it has to the second, therefore BC has to BG a V.Def.9
ratio duplicate of that which BC has to EF.
But BC is to BG as the triangle ABC is to the triangle ABG, therefore the triangle ABC also VI.1
has to the triangle ABG a ratio duplicate of that which BC has to EF. V.11
But the triangle ABG equals the triangle DEF, therefore the triangle ABC also has to the V.7
triangle DEF a ratio duplicate of that which BC has to EF.
Therefore, similar triangles are to one another in the duplicate ratio of the corresponding sides.
Q.E.D.
Corollary
From this it is manifest that if three straight lines are proportional, then the first is to the third as the
figure described on the first is to that which is similar and similarly described on the second.
This proposition is used in the proof of the next one, and the corollary is used in the proofs of VI.22,
VI.31, and X.6.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 20
Similar polygons are divided into similar triangles, and into triangles equal in multitude and in
the same ratio as the wholes, and the polygon has to the polygon a ratio duplicate of that
which the corresponding side has to the corresponding side.
Let ABCDE and FGHKL be similar polygons, and let AB correspond to FG.
I say that the polygons ABCDE and FGHKL are divided into similar triangles, and into triangles equal
in multitude and in the same ratio as the wholes, and the polygon ABCDE has to the polygon FGHKL
a ratio duplicate of that which AB has to FG.
Now, since the polygon ABCDE is similar to the polygon FGHKL, therefore the angle BAE VI.Def.1
equals the angle GFL, and AB is to AE as GF is to FL.
Since then ABE and FGL are two triangles having one angle equal to one angle and the sides VI.6
about the equal angles proportional, therefore the triangle ABE is equiangular with the VI.4
triangle FGL, so that it is also similar, therefore the angle ABE equals the angle FGL. VI.Def.1
But the whole angle ABC also equals the whole angle FGH because of the similarity of the
polygons, therefore the remaining angle EBC equals the angle LGH.
And, since the triangles ABE and FGL are similar, BE is to AB as GL is to GF. Also, since
V.22
the polygons are similar, AB is to BC as FG is to GH. Therefore, ex aequali, BE is to BC as
VI.6
GL is to GH, that is, the sides about the equal angles EBC and LGH are proportional. VI.4
Therefore the triangle EBC is equiangular with the triangle LGH, so that the triangle EBC is VI.Def.1
also similar to the triangle LGH.
For the same reason the triangle ECD is also similar to the triangle LHK.
Therefore the similar polygons ABCDE and FGHKL have been divided into similar
triangles, and into triangles equal in multitude.
I say that they are also in the same ratio as the wholes, that is, in such manner that the
triangles are proportional, and ABE, EBC, and ECD are antecedents, while FGL, LGH, and
LHK are their consequents, and that the polygon ABCDE has to the polygon FGHKL a ratio
duplicate of that which the corresponding side has to the corresponding side, that is AB to
FG.
Then since the polygons are similar, the angle ABC equals the angle FGH, and AB is to BC
as FG is to GH, the triangle ABC is equiangular with the triangle FGH, therefore the angle VI.6
BAC equals the angle GFH, and the angle BCA to the angle GHF.
And, since the angle BAM equals the angle GFN, and the angle ABM also equals the angle
FGN, therefore the remaining angle AMB also equals the remaining angle FNG. Therefore I.32
the triangle ABM is equiangular with the triangle FGN.
Similarly we can prove that the triangle BMC is also equiangular with the triangle GNH.
But AM is to MC as the triangle ABM is to MBC, and as AME is to EMC, for they are to one VI.1
another as their bases.
Therefore also one of the antecedents is to one of the consequents as are all the antecedents V.12
to all the consequents, therefore the triangle AMB is to BMC as ABE is to CBE.
But AMB is to BMC as AM is to MC, therefore AM is to MC as the triangle ABE is to the V.11
triangle EBC.
For the same reason also FN is to NH as the triangle FGL is to the triangle GLH.
And AM is to MC as FN is to NH, therefore the triangle ABE is to the triangle BEC as the
V.11
triangle FGL is to the triangle GLH, and, alternately, the triangle ABE is to the triangle FGL V.16
as the triangle BEC is to the triangle GLH.
Similarly we can prove, if BD and GK are joined, that the triangle BEC is to the triangle
LGH as the triangle ECD is to the triangle LHK.
And since the triangle ABE is to the triangle FGL as EBC is to LGH, and further as ECD is
to LHK, therefore also one of the antecedents is to one of the consequents as the sum of the V.12
antecedents to the sum of the consequents. Therefore the triangle ABE is to the triangle FGL
as the polygon ABCDE is to the polygon FGHKL.
But the triangle ABE has to the triangle FGL a ratio duplicate of that which the
corresponding side AB has to the corresponding side FG, for similar triangles are in the VI.19
duplicate ratio of the corresponding sides.
Therefore the polygon ABCDE also has to the polygon FGHKL a ratio duplicate of that V.11
which the corresponding side AB has to the corresponding side FG.
Therefore, similar polygons are divided into similar triangles, and into triangles equal in multitude
and in the same ratio as the wholes, and the polygon has to the polygon a ratio duplicate of that
which the corresponding side has to the corresponding side.
Q.E.D.
Corollary
Similarly also it can be proved in the case of quadrilaterals that they are in the duplicate ratio of the
corresponding sides. And it was also proved in the case of triangles, therefore also, generally, similar
rectilinear figures are to one another in the duplicate ratio of the corresponding sides.
This proposition and its corollary are used occassionally in Books X, XII, and XIII, in particular, XII.1
and XII.8.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 21
Figures which are similar to the same rectilinear figure are also similar to one another.
Let each of the rectilinear figures A and B be similar to C.
Since A is similar to C, it is equiangular with it and has the sides about the equal angles VI.Def.1
proportional.
Again, since B is similar to C, it is equiangular with it and has the sides about the equal
angles proportional.
Therefore each of the figures A and B is equiangular with C and with C has the sides about V.11
the equal angles proportional, therefore A is similar to B.
Therefore, figures which are similar to the same rectilinear figure are also similar to one another.
Q.E.D.
This proposition is used in the proofs of propositions VI.24, VI.28, and VI.29. It also would have been
useful in the proof of VI.8.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 22
If four straight lines are proportional, then the rectilinear figures similar and similarly
described upon them are also proportional; and, if the rectilinear figures similar and similarly
described upon them are proportional, then the straight lines are themselves also proportional.
Let the four straight lines AB, CD, EF, and GH be proportional, so that AB is to CD as EF is to
GH. Let the similar and similarly situated rectilinear figures KAB and LCD be described on AB
and CD, and the similar and similarly situated rectilinear figures MF and NH be described on
EF and GH.
Take a third proportional O to AB and CD, and a third proportional P to EF and GH. VI.11
Therefore MF has the same ratio to each of the figures NH and SR, therefore NH V.9
equals SR.
But it is also similar and similarly situated to it, therefore GH equals QR.
Therefore, if four straight lines are proportional, then the rectilinear figures similar and similarly
described upon them are also proportional; and, if the rectilinear figures similar and similarly
described upon them are proportional, then the straight lines are themselves also proportional.
Q.E.D.
There is a missing step near the end of the proof, namely, the justification of the statement that GH
equals QR is missing. Just before it we have NH and SR are similar equal rectilinear figures, and we
want to conclude the corresponding sides GH and QR are equal. The needed demonstration is not
difficult to supply.
This proposition is used in the proofs of several propositions in Book X and in XII.4 in Book XII.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 23
Equiangular parallelograms have to one another the ratio compounded of the ratios of their
sides.
Let AC and CF be equiangular parallelograms having the angle BCD equal to the angle ECG.
I say that the parallelogram AC has to the parallelogram CF the ratio compounded of the ratios
of the sides.
Let them be placed so that BC is in a straight line with CG. Then DC is also in a straight line I.14
with CE.
But the ratio of K to M is compounded of the ratio of K to L and of that of L to M, so that K has
also to M the ratio compounded of the ratios of the sides.
Since then it was proved that K is to L as the parallelogram AC is to the parallelogram CH, and
L is to M as the parallelogram CH is to the parallelogram CF, therefore, ex aequali K is to M as V.22
AC is to the parallelogram CF.
But K has to M the ratio compounded of the ratios of the sides, therefore AC also has to CF the
ratio compounded of the ratios of the sides.
Therefore, equiangular parallelograms have to one another the ratio compounded of the ratios of
their sides.
Q.E.D.
This proposition is a generalization of the basic formula for the area of a rectangle, that is, the area of a
rectangle is the product of its length and width. Such a formula depends on predetermined units of
length and area so that the unit area is the area of a square whose sides have length equal to the unit
length. Euclid and other Greek mathematicians did not use predetermined units of length or area, so they
expressed this formula as a proportion. We would state that proportion as saying the ratio of the area of a
given rectangle to the area of a given square is the product of the ratios of the lengths of the sides of the
rectangle to the length of a side of the square. Of course, Euclid would say that without using the words
'area' and 'length' as follows: the ratio of the a given rectangle to a given square is the product of the
ratios of the sides of the rectangle to a side of the square. Note that his terminology for a product of
ratios involves "compounding the ratios." A natural generalization of the ratio of a rectangle to a square
is the ratio of a rectangle to a rectangle. A broader generalization is the ratio of one parallelogram to
another parallelogram having the same angles. That gives the generalization as stated in this proposition.
These areas have been treated earlier in the Elements. Back in Book I and II the basic concept was
"quadrature," that is, finding a square or other shaped figure of the same area as the given rectangle or
parallelogram. That began with Proposition I.35 which said two triangles on the same base and with the
same height are equal, and ended with I.14 which constructed a square equal to a given rectangle.
Early in this book was the proposition VI.1 generalizing I.35 which said that parallelograms with the
same height are proportional to their bases. Finally, in this proposition we have the full statement about
areas of rectangles and parallelograms.
Proposition VIII.5 states that plane numbers have to one another the ratio compounded of the ratios of
their sides. That proposition is probably a much older version that may go back to the Pythagoreans
when "all was number." The discovery of incommensurable lines showed there were serious limitations
to that version of the proposition.
In Book XI there are analogous statements for volumes of parallelopipeds. For instance, Proposition
XI.33 states that similar parallelepipeds are to one another in the triplicate ratio of their corresponding
sides. That statement for parallelepipeds is analogous to this one for parallelograms.
Although this is a basic proposition on areas, it is actually not used in the rest of the Elements.
Book VI introduction
© 1996, 2002
D.E.Joyce
Clark University
Proposition 24
In any parallelogram the parallelograms about the diameter are similar both to the whole and
to one another.
Let ABCD be a parallelogram, and AC its diameter, and let EG and HK be parallelograms
about AC.
I say that each of the parallelograms EG and HK is similar both to the whole ABCD and to
the other.
Therefore in the parallelograms ABCD and EG, the sides about the common angle BAD are
proportional.
And, since GF is parallel to DC, the angle AFG equals the angle ACD, and the angle DAC is
common to the two triangles ADC and AGF, therefore the triangle ADC is equiangular with I.29
the triangle AGF.
For the same reason the triangle ACB is also equiangular with the triangle AFE, and the
whole parallelogram ABCD is equiangular with the parallelogram EG.
Therefore in the parallelograms ABCD and EG the sides about the equal angles are VI.Def.1
proportional. Therefore the parallelogram ABCD is similar to the parallelogram EG.
For the same reason the parallelogram ABCD is also similar to the parallelogram KH.
Therefore each of the parallelograms EG and HK is similar to ABCD.
But figures similar to the same rectilinear figure are also similar to one another, therefore VI.21
the parallelogram EG is also similar to the parallelogram HK.
Therefore, in any parallelogram the parallelograms about the diameter are similar both to the whole
and to one another.
Q.E.D.
With this proposition Euclid returns to applications of areas. Back in proposition I.45 rectilinear areas
were applied to lines. In the upcoming propositions VI.28 and VI.29, rectilinear areas will be applied to
lines but the areas will fall short of or extend beyond the end of the lines. Those propositions geometric
solve two kinds of quadratic equations. This proposition is preporatory to them. It is used in the proofs
of VI.26 (its converse) and VI.29.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 25
To construct a figure similar to one given rectilinear figure and equal to another.
Let ABC be the given rectilinear figure to which the figure to be constructed must be similar,
and D that to which it must be equal.
Let there be applied to BC the parallelogram BE equal to the triangle ABC, and to CE the I.44
parallelogram CM equal to D in the angle FCE which equals the angle CBL. I.45
Take a mean proportional GH to BC and CF, and describe KGH similar and similarly VI.13
situated to ABC on GH. VI.18
Then, since BC is to GH as GH is to CF, and, if three straight lines are proportional, then
the first is to the third as the figure on the first is to the similar and similarly situated figure V.19,Cor
described on the second, therefore BC is to CF as the triangle ABC is to the triangle KGH.
Therefore also the triangle ABC is to the triangle KGH as the parallelogram BE is to the
V.11
parallelogram EF. Therefore, alternately, the triangle ABC is to the parallelogram BE as the V.16
triangle KGH is to the parallelogram EF.
But the triangle ABC equals the parallelogram BE, therefore the triangle KGH also equals (V.14)
the parallelogram EF. And the parallelogram EF equals D, therefore KGH also equals D.
And KGH is also similar to ABC. Therefore this figure KGH has been constructed similar to
the given rectilinear figure ABC and equal to the other given figure D.
Q.E.F.
Note that it isn't proposition V.14 being invoked near the end of the proof, but an alternate form of it.
See the Guide to V.14.
This proposition solves a similar problem, to find a figure with the size of one figure but the shape of
another, a problem reputedly solved by Pythagoras. It is used in the proofs of propositions VI.28 and
VI.29
Book VI introduction
© 1996, 1997
D.E.Joyce
Clark University
Proposition 26
If from a parallelogram there is taken away a parallelogram similar and similarly situated to
the whole and having a common angle with it, then it is about the same diameter with the
whole.
From the parallelogram ABCD let there be taken away the parallelogram AF similar and
similarly situated to ABCD, and having the angle DAB common with it.
VI.
But also, since ABCD and EG are similar, therefore DA is to AB as GA is to AE. Therefore Def.1
GA is to AK as GA is to AE. V.11
Therefore GA has the same ratio to each of the straight lines AK and AE.
Therefore AE equals AK the less equals the greater, which is impossible. V.9
Therefore ABCD cannot fail to be about the same diameter with AF. Therefore the
parallelogram ABCD is about the same diameter with the parallelogram AF.
Therefore, if from a parallelogram there is taken away a parallelogram similar and similarly situated
to the whole and having a common angle with it, then it is about the same diameter with the whole.
Q.E.D.
This proposition is the converse of VI.24. It is used in the proofs of the next three and for a few
propositions in Book X.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 27
Of all the parallelograms applied to the same straight line falling short by parallelogrammic
figures similar and similarly situated to that described on the half of the straight line, that
parallelogram is greatest which is applied to the half of the straight line and is similar to the
difference.
Let AB be a straight line and let it be bisected at C. Let there be applied to the straight line AB
the parallelogram AD falling short by the parallelogrammic figure DB described on the half of
AB, that is, CB.
I say that, of all the parallelograms applied to AB falling short by parallelogrammic figures
similar and similarly situated to DB, AD is greatest.
Since the parallelogram DB is similar to the parallelogram FB, therefore they are about the VI.26
same diameter.
Then, since CF equals FE, and FB is common, therefore the whole CH equals the whole KE. I.43
Add CF to each. Therefore the whole AF equals the gnomon LMN, so that the parallelogram
DB, that is, AD, is greater than the parallelogram AF.
Therefore, of all the parallelograms applied to the same straight line falling short by
parallelogrammic figures similar and similarly situated to that described on the half of the straight
line, that parallelogram is greatest which is applied to the half of the straight line and is similar to the
difference.
Q.E.D.
This proposition clarifies the limitations of the next one, VI.28. In VI.28 a construction is made to apply
a parallelogram equal to a given rectilinear figure to a line falling short by a parallelogrammic figure.
This proposition implies that that construction cannot be made if the given rectilinear figure is too large.
When that proposition is applied, the part which falls short is usually a square, not just any
parallelogram, and this and the next proposition are much more easily understood in that case. In that
case, the next proposition applies a rectangle equal to a given area to a line but falling short by a square.
And this proposition implies that can only be done if the given area is at least the square on half the line,
since that square is the greatest rectangle that can be so applied.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 28
To apply a parallelogram equal to a given rectilinear figure to a given straight line but
falling short by a parallelogram similar to a given one; thus the given rectilinear figure VI.27
must not be greater than the parallelogram described on the half of the straight line and
similar to the given parallelogram.
Let C be the given rectilinear figure, AB the given straight line, and D the given parallelogram,
and let C not be greater than the parallelogram described on the half of AB similar to the given
parallelogram D.
It is required to apply a parallelogram equal to the given rectilinear figure C to the given
straight line AB but falling short by a parallelogram similar to D.
Bisect AB at the point E. Describe EBFG similar and similarly situated to D on EB, and I.9
complete the parallelogram AG. VI.18
If then AG equals C, that which was proposed is done, for the parallelogram AG equal to the
given rectilinear figure C has been applied to the given straight line AB but falling short by a
parallelogram GB similar to D.
Construct KLMN equal to GB minus C and similar and similarly situated to D. VI.25
Now, since GB equals C and KM, therefore GB is greater than KM, therefore also GE is greater
than KL, and GF than LM.
Make GO equal to KL, and GP equal to LM, and let the parallelogram OGPQ be completed,
therefore it is equal and similar to KM.
VI.21
Therefore GQ is also similar to GB, therefore GQ is about the same diameter with GB. VI.26
Then, since BG equals C and KM, and in them GQ equals KM, therefore the remainder, the
gnomon UWV, equals the remainder C.
And, since PR equals OS, add QB to each, therefore the whole PB equals the whole OB.
But OB equals TE, since the side AE also equals the side EB, therefore TE also equals PB. I.36
Add OS to each. Therefore the whole TS equals the whole, the gnomon VWU.
But the gnomon VWU was proved equal to C, therefore TS also equals C.
Therefore there the parallelogram ST equal to the given rectilinear figure C has been applied to
the given straight line AB but falling short by a parallelogram QB similar to D.
Q.E.F.
When this proposition is used, the given parallelgram D usually is a square. Then the problem is to cut
the line AB at a point S so that the rectangle AS by SB equals the given rectilinear figure C. This special
case can be proved with the help of the propositions in Book II. See the Guide to proposition II.5 for
more details.
The proof of the current propostion is difficult to follow. It is simplified when we take the special case
mentioned above, namely, when the given parallelogram D is a square. The simplified proof is easier to
follow since the rest of the parallelograms mentioned all become rectangles.
We can understand the meaning of this construction more easily if we interpret it algebraically. Let a
stand for the known quantity AB, and c the known quantity C. Then let x and y stand for the unknown
quantities SB and SA. Then this construction finds x and y so that their sum is a and their product is c.
In terms of the single variable x, the construction solves the quadratic equation ax – x2 = C.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 29
To apply a parallelogram equal to a given rectilinear figure to a given straight line but
exceeding it by a parallelogram similar to a given one.
Let C be the given rectilinear figure, AB be the given straight line, and D the parallelogram to
which the excess is required to be similar.
It is required to apply a parallelogram equal to the the rectilinear figure C to the straight line
AB but exceeding it by a parallelogram similar to D.
Bisect AB at E. Describe the parallelogram BF on EB similar and similarly situated to D, and VI.25
construct GH equal to the sum of BF and C and similar and similarly situated to D.
Now, since GH is greater than FB, therefore KH is also greater than FL, and KG greater
thanFE.
Produce FL and FE. Make FLM equal to KH, and FEN equal to KG. Complete MN. Then MN
is both equal and similar to GH.
But GH is similar to EL, therefore MN is also similar to EL, therefore EL is about the same VI.21
diameter with MN. VI.26
Since GH equals the sum of EL and C, while GH equals MN, therefore MN also equals the
sum of EL and C.
Subtract EL from each. Therefore the remainder, the gnomon XWV, equals C.
I.36
Now, since AE equals EB, therefore AN equals NB, that is, LP. I.43
Therefore the parallelogram AO equal to the given rectilinear figure C has been applied to the
given straight line AB but exceeding it by a parallelogram QP similar to D, since PQ is also VI.24
similar to EL.
Q.E.F.
The construction in this proposition is a generalization of that described in the Guide for II.6. In that
proposition, the figure D is a square.
Like the last proposition, this one is more easily understood when the given parallelogram D is a square.
As in the last proposition we can understand the meaning of this construction more easily if we interpret
it algebraically. Let a stand for the known quantity AB and x stand for the unknown quantity BP. Then
this construction finds x so that (a + x) x = C. In other words it solves the quadratic equation ax + x2 = C.
The construction of this proposition is used in the next proposition to cut a straight line in extreme and
mean ratio.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 30
But it is also equiangular with it, therefore in BF and AD the sides about the equal angles are VI.14
reciprocally proportional. Therefore FE is to ED as AE is to EB.
The construction given here cuts a line into two parts A and B so that (A + B):B = B:A. By proposition
VI.17, that condition is equivalent to making the rectangle A + B by A equal the square B by B, and that
is the construction of proposition II.11.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 31
In right-angled triangles the figure on the side opposite the right angle equals the sum of the
similar and similarly described figures on the sides containing the right angle.
Let ABC be a right-angled triangle having the angle BAC right.
I say that the figure on BC equals the sum of the similar and similarly described figures
on BA and AC.
And, since three straight lines are proportional, the first is to the third as the figure on the VI.19,Cor
first is to the similar and similarly described figure on the second.
For the same reason also, BC is to CD as the figure on BC is to that on CA, so that, in
addition, BC is to the sum of BD and DC as the figure on BC is to the sum of the similar V.24
and similarly described figures on BA and AC.
But BC equals the sum of BD and DC, therefore the figure on BC equals the sum of the
similar and similarly described figures on BA and AC.
Therefore, in right-angled triangles the figure on the side opposite the right angle equals the sum of
the similar and similarly described figures on the sides containing the right angle.
Q.E.D.
This proposition is a generalization of I.47 where the squares in I.47 are replaced by any similar
rectilinear figures.
Proclus says that this proposition is Euclid's own, and the proof may be his, but the result, if not the
proof, was known long before Euclid, at least in the time of Hippocrates a century before Euclid.
The broad problem Hippocrates was investigating was that of quadrature of a circle, also called squaring
a circle, which is to find a square (or any other polygon) with the same area as a given circle.
Hippocrates did not solve that problem, but he did solve a related one involving lunes. A lune (also
called a crescent) is a region of nonoverlap of two intersecting circles. Hippocrates did not succeed in
squaring an arbitrary lune, but he did succeed in a couple special cases. Here is a summary of the
simplist case.
Note that there are three segments of circles in the diagram; two
of them are small, namely, AGB with base AB, and BHC with
base BC, and one is large, namely, AFC with base AC. The first
two are congruent, and the third is similar to them since all
three are segments in quarters of circles.
Hippocrates then uses a version of this proposition VI.31—generalized so the figures don't have to be
rectilinear but may have curved sides— to conclude that the sum of the two small segments,
AGB + BHC, equals the large segment AFC, since the bases of the small segments are sides of a right
triangle while the base of the large segment is the triangle's hypotenuse.
Therefore, the lune, which is the semicircle minus the large segment, equals the semicircle minus the
sum of the small segments. But the semicircle minus the sum of the small segments is just the right
triangle ABC. Thus, a rectilinear figure (the triangle) has been found equal to the lune, as required.
Note that at Hippocrates' time, Eudoxus' theory of proportion had not been developed, so the
understanding of the theory of similar figures (Book V) was not as complete as it was after Eudoxus.
Also, Eudoxus' principle of exhaustion (see Book XI and proposition X.1) for finding areas of curved
figures was still to come, so the concept of area of curved figures was on shaky ground, too. Such a
situation is common in mathematics—mathematics advances into new territory long before the
foundations of mathematics are developed to logically justify those advances.
Book VI introduction
© 1996, 2002
D.E.Joyce
Clark University
Proposition 32
If two triangles having two sides proportional to two sides are placed together at one angle so
that their corresponding sides are also parallel, then the remaining sides of the triangles are in
a straight line.
Let ABC and DCE be two triangles having the two sides AB and AC proportional to the two
sides DC and DE, so that AB is to AC as DC is to DE, and AB parallel to DC, and AC parallel to
DE.
For the same reason the angle CDE also equals the angle
ACD, so that the angle BAC equals the angle CDE.
And, since ABC and DCE are two triangles having one angle, the angle at A, equal to one angle,
the angle at D, and the sides about the equal angles proportional, so that AB is to AC as DC is to VI.6
DE, therefore the triangle ABC is equiangular with the triangle DCE. Therefore the angle ABC
equals the angle DCE.
But the angle ACD was also proved equal to the angle BAC, therefore the whole angle ACE
equals the sum of the two angles ABC and BAC.
Add the angle ACB to each. Therefore the sum of the angles ACE and ACB equals the sum of
the angles BAC, ACB, and CBA.
But the sum of the angles BAC, ABC, and ACB equals two right angles, therefore the sum of the I.32
angles ACE and ACB also equals two right angles.
Therefore with a straight line AC, and at the point C on it, the two straight lines BC and CE not
lying on the same side make the sum of the adjacent angles ACE and ACB equal to two right I.14
angles. Therefore BC is in a straight line with CE.
Therefore, if two triangles having two sides proportional to two sides are placed together at one angle
so that their corresponding sides are also parallel, then the remaining sides of the triangles are in a
straight line.
Q.E.D.
The corresponding sides mentioned in the statement of the proposition are supposed to be directed in the
same direction, even though that is not explicitly stated.
This proposition is used in the proof of proposition XIII.17 which inscribes a regular dodecahedron in a
sphere.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Proposition 33
Angles in equal circles have the same ratio as the circumferences on which they stand whether
they stand at the centers or at the circumferences.
Let ABC and DEF be equal circles, and let the angles BGC and EHF be angles at their centers G and
H, and the angles BAC and EDF angles at the circumferences.
I say that the circumference BC is to the circumference EF as the angle BGC is to the angle EHF, and
as the angle BAC is to the angle EDF.
Make any number of consecutive circumferences CK and KL equal to the circumference BC,
and any number of consecutive circumferences FM, MN equal to the circumference EF, and
join GK and GL and HM and HN.
Then, since the circumferences BC, CK, and KL equal one another, the angles BGC, CGK,
and KGL also equal one another. Therefore, whatever multiple the circumference BL is of III.27
BC, the angle BGL is also that multiple of the angle BGC.
For the same reason, whatever multiple the circumference NE is of EF, the angle NHE is also
that multiple of the angle EHF.
If the circumference BL equals the circumference EN, then the angle BGL also equals the
angle EHN; if the circumference BL is greater than the circumference EN, then the angle III.27
BGL is also greater than the angle EHN; and, if less, less.
There being then four magnitudes, two circumferences BC and EF, and two angles BGC and
EHF, there have been taken, of the circumference BC and the angle BGC equimultiples,
namely the circumference BL and the angle BGL, and of the circumference EF and the angle
EHF equimultiples, namely the circumference EN and the angle EHN.
And it has been proved that, if the circumference BL is in excess of the circumference EN,
the angle BGL is also in excess of the angle EHN; if equal, equal; and if less, less.
Therefore the circumference BC is to EF as the angle BGC is to the angle EHF. V.Def.5
But the angle BGC is to the angle EHF as the angle BAC is to the angle EDF, for they are V.15
doubles respectively. III.20
Therefore also the circumference BC is to the circumference EF as the angle BGC is to the
angle EHF, and the angle BAC to the angle EDF.
Therefore, angles in equal circles have the same ratio as the circumferences on which they stand
whether they stand at the centers or at the circumferences.
Q.E.D.
This proposition stands apart from the rest in this book since it depends on none of them, but it is like
the first proposition VI.1 which established a proportion between lines and plane figures since it
establishes a proportion between angles and portions of circumferences cut off by those angles.
This proposition is used in three consecutive propositions in Book XIII starting with XIII.8 to convert
statements about arcs to statements about angles. Incidentally, all three have to do with regular
pentagons inscribed in circles.
Book VI introduction
© 1996
D.E.Joyce
Clark University
Table of contents
● Definitions (22)
● Propositions (39)
Guide
Definitions
Definition 1
A unit is that by virtue of which each of the things that exist is called one.
Definition 2
A number is a multitude composed of units.
Definition 3
A number is a part of a number, the less of the greater, when it measures the greater;
Definition 4
Definition 5
The greater number is a multiple of the less when it is measured by the less.
Definition 6
An even number is that which is divisible into two equal parts.
Definition 7
An odd number is that which is not divisible into two equal parts, or that which differs by a unit
from an even number.
Definition 8
An even-times even number is that which is measured by an even number according to an even
number.
Definition 9
An even-times odd number is that which is measured by an even number according to an odd
number.
Definition 10
An odd-times odd number is that which is measured by an odd number according to an odd
number.
Definition 11
A prime number is that which is measured by a unit alone.
Definition 12
Numbers relatively prime are those which are measured by a unit alone as a common measure.
Definition 13
A composite number is that which is measured by some number.
Definition 14
Numbers relatively composite are those which are measured by some number as a common
measure.
Definition 15
A number is said to multiply a number when that which is multiplied is added to itself as many
times as there are units in the other.
Definition 16
And, when two numbers having multiplied one another make some number, the number so
produced be called plane, and its sides are the numbers which have multiplied one another.
Definition 17
And, when three numbers having multiplied one another make some number, the number so
produced be called solid, and its sides are the numbers which have multiplied one another.
Definition 18
A square number is equal multiplied by equal, or a number which is contained by two equal
numbers.
Definition 19
And a cube is equal multiplied by equal and again by equal, or a number which is contained by
three equal numbers.
Definition 20
Numbers are proportional when the first is the same multiple, or the same part, or the same parts,
of the second that the third is of the fourth.
Definition 21
Similar plane and solid numbers are those which have their sides proportional.
Definition 22
A perfect number is that which is equal to the sum its own parts.
Propositions
Proposition 1
When two unequal numbers are set out, and the less is continually subtracted in turn from the
greater, if the number which is left never measures the one before it until a unit is left, then the
original numbers are relatively prime.
Proposition 2
To find the greatest common measure of two given numbers not relatively prime.
Corollary. If a number measures two numbers, then it also measures their greatest common
measure.
Proposition 3
To find the greatest common measure of three given numbers not relatively prime.
Proposition 4
Any number is either a part or parts of any number, the less of the greater.
Proposition 5
If a number is part of a number, and another is the same part of another, then the sum is also the
same part of the sum that the one is of the one.
Proposition 6
If a number is parts of a number, and another is the same parts of another, then the sum is also the
same parts of the sum that the one is of the one.
Proposition 7
If a number is that part of a number which a subtracted number is of a subtracted number, then
the remainder is also the same part of the remainder that the whole is of the whole.
Proposition 8
If a number is the same parts of a number that a subtracted number is of a subtracted number,
then the remainder is also the same parts of the remainder that the whole is of the whole.
Proposition 9
If a number is a part of a number, and another is the same part of another, then alternately,
whatever part of parts the first is of the third, the same part, or the same parts, the second is of the
fourth.
Proposition 10
If a number is a parts of a number, and another is the same parts of another, then alternately,
whatever part of parts the first is of the third, the same part, or the same parts, the second is of the
fourth.
Proposition 11
If a whole is to a whole as a subtracted number is to a subtracted number, then the remainder is to
the remainder as the whole is to the whole.
Proposition 12
If any number of numbers are proportional, then one of the antecedents is to one of the
consequents as the sum of the antecedents is to the sum of the consequents.
Proposition 13
If four numbers are proportional, then they are also proportional alternately.
Proposition 14
If there are any number of numbers, and others equal to them in multitude, which taken two and
two together are in the same ratio, then they are also in the same ratio ex aequali.
Proposition 15
If a unit number measures any number, and another number measures any other number the same
number of times, then alternately, the unit measures the third number the same number of times
that the second measures the fourth.
Proposition 16
If two numbers multiplied by one another make certain numbers, then the numbers so produced
equal one another.
Proposition 17
If a number multiplied by two numbers makes certain numbers, then the numbers so produced
have the same ratio as the numbers multiplied.
Proposition 18
If two number multiplied by any number make certain numbers, then the numbers so produced
have the same ratio as the multipliers.
Proposition 19
If four numbers are proportional, then the number produced from the first and fourth equals the
number produced from the second and third; and, if the number produced from the first and
fourth equals that produced from the second and third, then the four numbers are proportional.
Proposition 20
The least numbers of those which have the same ratio with them measure those which have the
same ratio with them the same number of times; the greater the greater; and the less the less.
Proposition 21
Numbers relatively prime are the least of those which have the same ratio with them.
Proposition 22
The least numbers of those which have the same ratio with them are relatively prime.
Proposition 23
If two numbers are relatively prime, then any number which measures one of them is relatively
prime to the remaining number.
Proposition 24
If two numbers are relatively prime to any number, then their product is also relatively prime to
the same.
Proposition 25
If two numbers are relatively prime, then the product of one of them with itself is relatively prime
to the remaining one.
Proposition 26
If two numbers are relatively prime to two numbers, both to each, then their products are also
relatively prime.
Proposition 27
If two numbers are relatively prime, and each multiplied by itself makes a certain number, then
the products are relatively prime; and, if the original numbers multiplied by the products make
certain numbers, then the latter are also relatively prime.
Proposition 28
If two numbers are relatively prime, then their sum is also prime to each of them; and, if the sum
of two numbers is relatively prime to either of them, then the original numbers are also relatively
prime.
Proposition 29
Any prime number is relatively prime to any number which it does not measure.
Proposition 30
If two numbers, multiplied by one another make some number, and any prime number measures
the product, then it also measures one of the original numbers.
Proposition 31
Any composite number is measured by some prime number.
Proposition 32
Any number is either prime or is measured by some prime number.
Proposition 33
Given as many numbers as we please, to find the least of those which have the same ratio with
them.
Proposition 34
Proposition 35
If two numbers measure any number, then the least number measured by them also measures the
same.
Proposition 36
To find the least number which three given numbers measure.
Proposition 37
If a number is measured by any number, then the number which is measured has a part called by
the same name as the measuring number.
Proposition 38
If a number has any part whatever, then it is measured by a number called by the same name as
the part.
Proposition 39
To find the number which is the least that has given parts.
Book VII is the first of the three books on number theory. It begins with the 22 definitions used in these
books. The important definitions being those for unit and number, part and multiple, even and odd,
prime and relatively prime, proportion, and perfect number. The topics in Book VII are antenaresis and
the greatest common divisor, proportions of numbers, relatively prime numbers and prime numbers, and
the least common multiple.
The basic construction for Book VII is antenaresis, also called the Euclidean algorithm, a kind of
reciprocal subtraction. Beginning with two numbers, the smaller, whichever it is, is repeatedly
subtracted from the larger until a single number is left. This algorithm, studied in propositions VII.1
througth VII.3, results in the greatest common divisor of two or more numbers.
Propositions V.5 through V.10 develop properties of fractions, that is, they study how many parts one
number is of another in preparation for ratios and proportions. The next group of propositions VII.11
through VII.19 develop the theory of proportions for numbers.
Propositions VII.20 through VII.29 discuss representing ratios in lowest terms as relatively prime
numbers and properties of relatively prime numbers. Properties of prime numbers are presented in
propositions VII.30 through VII.32. Book VII finishes with least common multiples in propositions
VII.33 through VII.39.
Postulates are as necessary for numbers as they are for geometry. Missing postulates occurs as early as
proposition VII.2. In its proof, Euclid constructs a decreasing sequence of whole positive numbers, and,
apparently, uses a principle that conclude that the sequence must stop, that is, there cannot be an infinite
decreasing sequence of numbers. If that is the principle he uses, then it ought to be stated as a postulate
for numbers.
Numbers are so familiar that it hardly occurs to us that the theory of numbers needs axioms, too. In fact,
that field was one of the last to receive a careful scrutiny, and axioms for numbers weren't developed
until the late 19th century. By that time foundations for the rest of mathematics were laid upon either
geometry or number theory or both, and only geometry had axioms. About the same time that
foundations for number theory were developed, a new subject, set theory, was created by Cantor, and
mathematics was refounded in terms of set theory.
The foundations of number theory will be discussed in the Guides to the various definitions and
propositions.
Book VII
introduction
© 1996, 2002
D.E.Joyce
Clark University
Definition 22
A perfect number is that which is equal to the sum of its own parts.
For example, the number 28 is perfect because its parts (that is, proper divisors) 1, 2, 4, 7, and 14 sum to
28. The four smallest perfect numbers were known to the ancient Greek mathematicians. They are 6, 28,
496, and 8128. In proposition IX.36 Euclid gives a construction of even perfect numbers.
The divisors of these even perfect numbers can be listed in two columns, illustrated here for the divisors
of 496.
1 31
2 62
4 124
8 248
16 (496)
The first column lists powers of 2 from 20 up through 24. The sum of these powers of 2 is 31, which is
one less than 25. That number 31 appears at the top of the second column, and its repeated doubles up
through 496 appear on the second column. In such a tableau, the sum of all the numbers, except the last,
will equal the last.
The question of odd perfect numbers was not solved by Euclid. Probably the oldest open conjecture in
mathematics is that there are no ood perfect numbers. There is no proof yet, but it is known that if there
is an odd perfect number, then it has to be immensely huge.
© 1997, 2002
D.E.Joyce
Clark University
Definitions 1 and 2
Def. 1. A unit is that by virtue of which each of the things that exist is called one.
These 23 definitions at the beginning of Book VII are the definitions for all three books VII through IX
on number theory. Some won't be used until Books VIII or IX.
These first two definitions are not very constructive towards a theory of numbers. The numbers in
definition 2 are meant to be whole positive numbers greater than 1, and definition 1 is meant to define
the unit as 1. The word "monad," derived directly from the Greek, is sometimes used instead of "unit."
Euclid treats the unit, 1, separately from numbers, 2, 3, and so forth. This makes his proofs awkward in
some cases. Chrysippus (280–207), a Stoic philosopher, claimed that 1 is a number, but his
pronouncement was not accepted for some time.
Throughout these three books on number theory Euclid exhibits numbers as lines. In the diagram above,
if A is the unit, then BE is the number 3. But, just because he draws them as lines does not mean they are
lines, and he never calls them lines.
It is not clear what the nature of these numbers is supposed to be. But their nature is irrelevant. Euclid
could illustrate the unit as a line or as any other magnitude, and numbers would then be illustrated as
multiples of that unit.
There is a major distinction between lines and numbers. Lines are infinitely divisible, but numbers are
not, in particular, the unit is not divisible into smaller numbers.
Euclid has no postulates to elaborate the concept of number (other than the Common Notions which are
meant to apply to numbers as well as magnitudes of various kinds). Indeed, mathematicians did not
develop foundations for number theory until the late nineteenth century. Peano's axioms for numbers are
the best known. The most important of Peano's axioms is the principle of mathematical induction which
states that
Euclid does not use the principle of mathematical induction, but he does implicitly use a similar property
of numbers, namely, that any decreasing sequence of numbers is finite. That property is known variously
as the "well-ordering principle" for numbers and the "descending chain condition." We will discuss it
later in more detail.
© 1997, 2003
D.E.Joyce
Clark University
Definitions 3 through 5
Def. 3. A number is a part of a number, the less of the greater, when it measures the greater;
Def. 5. The greater number is a multiple of the less when it is measured by the less.
These definitions are in preparation for the definition of proportion of numbers given in VII.Def.20. In
the current definitions, the possible relations between a pair of numbers, m and n, are classified. Later in
Book VII, the term "ratio" will be used for this relation.
In all three of these definitions, the concept of "measures" is assumed to be understood. There is more to
these definitions than meets the eye, though, at least part of the intent is evident.
If u is the unit, then 2 is represented as AB while 6 is represented by CF. As AB measures CF three times
by CD, DE, and DF, therefore 2 is a part of 6, namely, the one-third part since it measures 6 three times.
We can also use the same figure as an illustration of VII.Def.5 to see that 6 is a multiple of 2, in
particular, the third multiple of 2.
Definition VII.Def.4 is less clear, but its intent can be read from the use to which it's put in VII.Def.20
for proportions of numbers. For an example, consider the numbers 4 and 6. The number 4 does not
measure the number 6, but it is parts of 6.
Here, 4 is represented as AC while 6 is represented as DG. Clearly, AC does not measure DG. The way
this definition is used in VII.Def.4, just the knowledge that "4 is parts of 6" is not enough, what is also
needed is how many parts of 6 is 4. This will be needed to define a proportion such as 4:6 = 6:9. That
proportion is supposed to hold since 4 is the same parts of 6 as 6 is of 9, namely 2 third parts. Thus, one
number being parts of another also carries along with it how many of what parts.
There is one more difficulty with this definition. It seems obvious that when one number m is less than
another n, then in all cases m would be parts of n, namely m consists of m one-nth parts of n. Yet, the
proposition VII.4 has a proof to show that m is either a part or parts of n.
Divisors
Where Euclid would say that m is a part of n, modern mathematicians would say that m is a proper
divisor of n. A divisor of n is any whole number m (including 1) that divides n in the sense that there is
another number k such that mk = n. A proper divisor of n is any divisor except n itself.
© 1997, 2002
D.E.Joyce
Clark University
Def. 7. An odd number is that which is not divisible into two equal parts, or that which differs by
a unit from an even number.
Def. 8. An even-times even number is that which is measured by an even number according to an
even number.
Def. 9. An even-times odd number is that which is measured by an even number according to an
odd number.
Def. 10. An odd-times odd number is that which is measured by an odd number according to an
odd number.
Definition 6 for "even number" is clear: the number n is even if it is of the form m + m.
Definition 7 for "odd number" has two statements. The first can be taken as a definition of odd number,
a number which is not divisible into two equal parts, that is to say not an even number.
The other statement is not a definition for odd number, since one has already been given, but an
unproved statement. It is easy to recognize that something has to be proved, since if we make the
analogous definitions for another number, say 10, then analogous statement is false. Suppose we say a
"decade number" is one divisible by 10, and and "undecade number' is one not divisible by 10. Then it is
not the case that an undecade number differs by a unit from a decade number; the number 13, for
instance, is not within 1 of a decade number.
The unproved statement that a number differing from an even number by 1 is an odd number ought to be
proved. That statment is used in proposition IX.22 and several propositions that follow it. It could be
proved using, for instance, a principle that any decreasing sequence of numbers is finite.
Definitions 8-10 are also clear. A product of two even numbers is an even-times even number; a product
of an even and an odd number is an even-times odd number; and a product of of two odd numbers is an
odd-times odd number. Note that a number like 12 is both even-times even and even-times odd being at
The numbers which are even-times even but not even-times odd are just the powers of 2: 4, 8, 16, 32,
etc. These are the numbers which are even-times even only, and they occur in proposition IX.32.
© 1997, 2002.
D.E.Joyce
Clark University
Definitions 11 through 14
Def. 11. A prime number is that which is measured by a unit alone.
Def. 12. Numbers relatively prime are those which are measured by a unit alone as a common
measure.
Def. 14. Numbers relatively composite are those which are measured by some number as a
common measure.
Prime numbers form a very important class of numbers, and much of number theory is devoted to their
analysis. The only proper divisor of a prime number is 1. The first few prime numbers are, of course, 2,
3, 5, 7, 11. Those numbers that aren't prime are composite, for instance, 4, 6, 8, 9, 10. The number 1
holds a special position. For Euclid, it was the unit rather than a number. For modern mathematicians 1
is also a unit, but in a different sense of the word, since it has a reciprocal, namely, itself.
Numbers are relatively prime if their only common divisor is 1. For example, 6 and 35 are relatively
prime (although neither is a prime number in itself). This situation is also phrased as "6 is prime to 35."
For another example, the three numbers 6, 10, and 15 are relatively prime since no number (except 1)
divides all three. If the numbers aren't relatively prime, then they're called "relatively composite," a term
rarely used now.
© 1997
D.E.Joyce
Clark University
Definitions 15 through 19
Def. 15. A number is said to multiply a number when that which is multiplied is added to itself as
many times as there are units in the other.
Def. 16. And, when two numbers having multiplied one another make some number, the number
so produced be called plane, and its sides are the numbers which have multiplied one another.
Def. 17. And, when three numbers having multiplied one another make some number, the number
so produced be called solid, and its sides are the numbers which have multiplied one another.
Def. 18. A square number is equal multiplied by equal, or a number which is contained by two
equal numbers.
Def. 19. And a cube is equal multiplied by equal and again by equal, or a number which is
contained by three equal numbers.
Notice that Euclid doesn't define addition and subtraction. Those operations are assumed to be
understood. But multiplication and proportion are defined, and proportion is defined next in VII.Def.20.
Definition 15 defines multiplication in terms of addition as a kind of composition. For instance, if 3 is
multiplied by 6, then since 6 is 1+1+1+1+1+1, therefore, 3 multiplied by 6 is 3+3+3+3+3+3. The first
proposition on multiplication is VII.16 which says multiplication is commutative. For our example, that
would say 3 multiplied by 6 equals 6 multiplied by 3, which is 6+6+6.
Figurate numbers
Plane numbers are the composite numbers. Each composite number can be a plane number in at least
one way, but most in more than one way. For instance, 16 can be viewed as a plane number either
with sides 2 and 8 or with sides 4 and 4, that is, as a square number.
Perhaps for the Pythagoreans, the most important figures were the triangular
numbers: 3, 6, especially 10, 15, 21, etc. Each could be formed from the previous by
adding a new row one unit longer. So, for instance, 10 = 1 + 2 + 3 + 4. For some
reason, Euclid doesn't mention triangular numbers. Indeed, he doesn't address sums of
arithmetic progressions at all, a subject of interest in many ancient cultures. Euclid
does give the sum of a geometric progression, that is, a continued proportion, in
proposition IX.35.
Definition 18 defines solid numbers. For example, if 18 is presented as 3 times 3 times 2, then it is
given as a solid number with three sides 3, 3, and 2. Solid numbers can be represented as a
configuration of dots or cubes in three dimensions.
Squares and cubes are are described as certain symmetric plane and solid numbers. Of course, some
numbers, such as 64, can be simultaneously squares and cubes.
© 1997
D.E.Joyce
Clark University
Definition 20
Def. 20. Numbers are proportional when the first is the same multiple, or the same part, or the
same parts, of the second that the third is of the fourth.
Definition V.20 for proportionality of numbers is not the same as the definition of proportionality for
magnitudes in Book V given in V.Def.5. This definition for numbers was probably the earlier one, but as
not all magnitudes are commensurable, it cannot adequately define proportionality for magnitudes.
This definition VII.20 is given by cases. The various cases correspond to defintions VII.Def.3 through
VII.Def.6 for part, parts, and multiple.
When four numbers, j, k, m, and n, are proportional, we'll write that symbolically as j:k = m:n,
In the first case, j is the same multiple of k as m is of n. An example of this is the proportion
12:6 = 22:11, where 12 is twice 6 and 22 is twice 11.
The second case is inverse to the first, j is the same part of k as m is of n. For an example take the
proportion 6:12 = 22:11, where 6 is one half of 12, and 11 is one half of 22.
For an example of the third case, consider 12:16 = 21:28. Since the first is the same parts of the second,
namely 3 parts of 4, as the the third is of the fourth, the proportion holds. Actually, there should be a
fourth case (inverse to the third case) when the second is the same parts of the first as the fourth is of the
third, as 16:12 = 28:21. Of course, these cases could be merged into one by considering 1 to be a number
and not distinguishing when the first is greater or less than the second.
Ratios of numbers
Although the word "ratio" doesn't appear in this definition, it appears frequently beginning in
proposition VII.14. In book VII ratio is restricted to the use of saying when one ratio is the same as
another, that is, there is a proportion as defined in this definition. In Book VIII, duplicate ratios,
triplicate ratios, and other compounded ratios appear. Definitions for these concepts are not explicitly
given, but once the concept of proportion has been defined, they have the same defintion given in Book
V for duplicate and triplicate ratio in V.Def.9-10. Compound ratios aren't defined in Book V, but they
can be understood by their use. See the Guide to V.Def.3. Various other definitions that go along with
ratios and proportions were given in Book V, for instance, alternate ratios, inverse ratios, taken jointly,
taken separately, and ex aequali. These definitions are also not repeated here in Book VII.
Very soon in these books on number theory Euclid begins to rely on properties of proportion proved in
Book V using the other definition of proportion. That these are valid for proportions of numbers could
be verified individually or by showing that the two definitions of proportion are equivalent for numbers.
Proportions of numbers first appear in proposition VII.11, but all the propositions from VII.4 through
VII.10 are in preparation for the study of numeric proportions.
© 1997, 2002
D.E.Joyce
Clark University
Definition 21
Def. 21. Similar plane and solid numbers are those which have their sides proportional.
To illustrate similar solid numbers, consider the two numbers 240 and 810 when represented as 4 times
6 times 10 and 6 times 9 times 15, respectively.
Proposition VIII.19 shows that the ratio of two similar solid numbers is the triplicate ratio of the
corresponding sides. In this example, the ratio 240:810 is triplicate of the ratio 4:6.
© 2002
D.E.Joyce
Clark University
Proposition 1
When two unequal numbers are set out, and the less is continually subtracted in turn from the
greater, if the number which is left never measures the one before it until a unit is left, then the
original numbers are relatively prime.
The less of two unequal numbers AB and CD being continually subtracted from the
greater, let the number which is left never measure the one before it until a unit is left.
I say that AB and CD are relatively prime, that is, that a unit alone measures AB and VII.Def.12
CD.
If AB and CD are not relatively prime, then some number E measures them. Let CD,
measuring BF, leave FA less than itself, let AF, measuring DG, leave GC less than
itself, and let GC, measuring FH, leave a unit HA.
But it also measures the whole BA, therefore it measures the remainder
AF. But AF measures DG, therefore E also measures DG. But it also
measures the whole DC, therefore it also measures the remainder CG.
Therefore no number measures the numbers AB and CD. Therefore AB and CD are VII.Def.12
relatively prime.
Therefore, when two unequal numbers are set out, and the less is continually subtracted in turn from
the greater, if the number which is left never measures the one before it until a unit is left, then the
original numbers are relatively prime.
Q.E.D.
Modern terminology uses the word "divides" rather than "measures," and the notation a | b is used to
abbreviate the phrase "a divides b."
This proposition assumes that 1 is the result of an antenaresis process. Antenaresis, also called the
Euclidean algorithm, is a kind of reciprocal subtraction. Beginning with two numbers, the smaller,
whichever it is, is repeatedly subtracted from the larger.
If the initial two numbers are a1 (AB in the proof) and a2 (CD), with a1 greater than a2, then the first
stage is to repeatedly subtract a2 from a1 until a remainder a3 (AF) less than a2 is found. That can be
stated algebraically as
a1 = m1 a2 + a3
where m1 is the number of times that a2 was subtracted from a1.
a2 = m2 a3 + a4.
For the hypotheses of this proposition, the algorithm stops when a remainder of 1 occurs:
a =m a + 1.
n-1 n-1 n
(In Euclid's proof, a is a5 which is AH.) The conclusion is that a1 and a2 are relatively prime.
n
The proof is not difficult. It depends on the observation that if b divides (that is, measures) both c and d,
then b divides their difference c - d. So, if some number b divides both a1 and a2, then it divides the
remainder a3, too. And since it divides both a2 and a3, it divides the remainder a4. And so forth, with the
final conclusion that b divides the last remainder 1.
Since there is no number b (and by "number" is meant a number greater than 1) which divides 1, there is
no number that divides both a1 and a2. Therefore a1 and a2 are relatively prime.
© 1996
D.E.Joyce
Clark University
Proposition 2
To find the greatest common measure of two given numbers not relatively prime.
Let AB and CD be the two given numbers not relatively prime.
But, if CD does not measure AB, then, when the less of the numbers AB
and CD being continually subtracted from the greater, some number is left
which measures the one before it.
VII.
For a unit is not left, otherwise AB and CD would be relatively prime, Def.12
which is contrary to the hypothesis. VII.1
Therefore some number is left which measures the one before it.
Now let CD, measuring BE, leave EA less than itself, let EA, measuring DF, leave FC less
than itself, and let CF measure AE.
Since then, CF measures AE, and AE measures DF, therefore CF also measures DF. But it
measures itself, therefore it also measures the whole CD.
But CD measures BE, therefore CF also measures BE. And it also measures EA, therefore
it measures the whole BA.
But it also measures CD, therefore CF measures AB and CD. Therefore CF is a common
measure of AB and CD.
If CF is not the greatest common measure of AB and CD, then some number G, which is
greater than CF, measures the numbers AB and CD.
Now, since G measures CD, and CD measures BE, therefore G also measures BE. But it
also measures the whole BA, therefore it measures the remainder AE.
But AE measures DF, therefore G also measures DF. And it measures the whole DC,
therefore it also measures the remainder CF, that is, the greater measures the less, which is
impossible.
Therefore no number which is greater than CF measures the numbers AB and CD.
Therefore CF is the greatest common measure of AB and CD.
Corollary
From this it is manifest that, if a number measures two numbers, then it also measures their greatest
common measure.
Euclid again uses antenaresis (the Euclidean algorithm) in this proposition, this time to find the greatest
common divisor of two numbers that aren't relatively prime. Had Euclid considered the unit (1) to be a
number, he could have merged these two propositions into one.
The greatest common divisor of two numbers m and n is the largest number which divides both. It's
usually denoted GCD(m, n). It can be found by antenaresis by repeatedly subtracting the smaller,
whichever it happens to be at the time, from the larger until the smaller divides the larger.
As an illustration consider the problem of computing the greatest common divisor of 884 and 3009.
First, repeatedly subtract 884 from 3009 until the remainder is less than 884. An equivalent numerical
operation is to divide 884 into 3009; you'll get the same remainder. In this case after subtracting 884
three times, the remainder is 357. The two numbers under our consideration are now 884 and 357.
Repeatedly subtract 357 from 884 to get the remainder 170. Repeatedly subtract 170 from 357 to get the
remainder 17. Finally, stop since 17 divides 170. We've found GCD(884,3009) equals 17.
The stages of the algorithm are the same as in VII.1 except that the final remainder a , which divides
n+1
the previous number a , is not 1.
n
a1 = m1 a2 + a3
a2 = m2 a3 + a4
...
a -1 = m -1 a + a +1.
n n n n
In the first part of the proof, Euclid shows that since a divides a , it also divides a , ... , a2, and a1.
n+1 n n-1
Therefore a is a common divisor of a2 and a1. In the last part of the proof, Euclid shows that if any
n+1
number d divides both a2 and a1, then it also divides a3, ... , a , and a . Therefore a is the greatest
n n+1 n+1
common divisor. The last part of the proof also shows that every common divisor divides the greatest
common divisor as noted in the corollary.
Euclid makes many implicit assumptions about numbers. For instance, he assumes that if m < n, then m
can be repeatedly subtracted from n until there is eventually a remainder less than or equal to m. He
seems to have recognized that magnitudes need not have this property since the property is used as a
qualifier in the definition of ratios (V.Def.4), but he didn't recognize its importance for numbers. There
are, in fact, nonstandard models of number theory which satisfy the usual properties of numbers, but do
not have this property. In such models, there are numbers than can be decreased by 1 infinitely many
times but not ever reach 1. The existance of such models implies an axiom is needed to exclude such
behavior.
There is a similar assumption that the process of antenaresis eventually reaches an end when applied to
numbers. Euclid certainly knew it needn't halt for magnitudes since its halting is used as a criterion for
incommensurability (X.2).
There needs to be an explicit axiom to cover these situations. One such axiom is a descending chain
condition which states that there is no infinite decreasing sequence of numbers
This proposition and its corollary are used in the next two propositions.
Note how similar this proposition is to X.3, even having the same diagram and the same corollary. The
terminology is slightly different and X.3 deals with magnitudes rather than numbers.
© 1996, 2002
D.E.Joyce
Clark University
Proposition 3
To find the greatest common measure of three given numbers not relatively prime.
Let A, B, and C be the three given numbers not relatively prime.
If D is not the greatest common measure of A, B, and C, then some number E, greater than
D, measures the numbers A, B, and C.
Therefore no number which is greater than D measures the numbers A, B, and C. Therefore
D is the greatest common measure of A, B, and C.
Since A, B, and C are not relatively prime, therefore some number measures them.
Now that which measures A, B, and C also measures A and B, and therefore measures D,
the greatest common measure of A and B. But it measures C also, therefore some number VII.2,Cor.
measures the numbers D and C. Therefore D and C are not relatively prime.
Then, since E measures D, and D measures A and B, therefore E also measures A and B.
But it measures C also, therefore E measures A, B, and C. Therefore E is a common
measure of A, B, and C.
If E is not the greatest common measure of A, B, and C, then some number F, greater than
E, measures the numbers A, B, and C.
Now, since F measures A, B, and C, it also measures A and B, therefore it measures the
greatest common measure of A and B. But the greatest common measure of A and B is D, VII.2,Cor.
therefore F measures D.
And it measures C also, therefore F measures D and C. Therefore it also measures the
greatest common measure of D and C. But the greatest common measure of D and C is E, VII.2,Cor.
therefore F measures E, the greater the less, which is impossible.
Therefore no number which is greater than E measures the numbers A, B, and C. Therefore
E is the greatest common measure of A, B, and C.
Q.E.D.
A common modern notation for the greatest common divisor of two numbers a and b is GCD(a, b).
Also, the notation a | b is typically used to indicate that a divides b.
The proof that this construction works is simplified if 1 is considered to be a number. Then, two
numbers are relatively prime when their GCD is 1, and Euclid's first case in the proof is subsumed in the
second.
Let d = GCD(a, b), and let e = GCD(d, c). Since e | d, d | a, and d | b, it follows that e | a and e | b, so e,
in fact, is a common divisor of a, b, and c.
In order to show that e is the greatest common divisor, let f be any common divisor of a, b, and c. Then
as f | a and f | b, therefore f | GCD(a, b), that is, f | d. Also, as f | d and f | c, therefore f | GCD(d, c), that is
f | e. Therefore e is the greatest common divisor of a, b, and c. Q.E.D.
© 1996
D.E.Joyce
Clark University
Proposition 4
Any number is either a part or parts of any number, the less of the greater.
Let A and BC be two numbers, and let BC be the less.
Then, if BC is divided into the units in it, then each unit of those in BC is VII.Def.4
some part of A, so that BC is parts of A.
Next let A and BC not be relatively prime, then BC either measures, or does
not measure, A.
Now if BC measures A, then BC is a part of A. But, if not, take the greatest common VII.
measure D of A and BC, and divide BC into the numbers equal to D, namely BE, EF, and Def.3
FC. VII.2
Now, since D measures A, therefore D is a part of A. But D equals each of the numbers BE,
EF, and FC, therefore each of the numbers BE, EF, and FC is also a part of A, so that BC is
parts of A.
Therefore, any number is either a part or parts of any number, the less of the greater.
Q.E.D.
This proposition says that if b is a smaller number than a, then b is either a part of a, that is, b is a unit
fraction of a, or b is parts of a, that is, a proper fraction, but not a unit fraction, of a. For instance, 2 is
one part of 6, namely, one third part; but 4 is parts of 6, namely, 2 third parts of 6.
It seems obvious that when one number b is less than another a, then in all cases b would be parts of a,
namely b consists of b of the ath parts of a. For instance, 4 consists of 4 sixth parts of 6. Yet, the proof of
this proposition ignores that possibility, except in the special case when b and a are relatively prime. In
the case of 4 and 6, the proof will find that 4 is 2 third parts of 6. Thus, it appears that a satisfactory
answer to the question "How mary parts of a is b?" requires finding the least number of parts.
© 1996, 2002
D.E.Joyce
Clark University
Proposition 5
If a number is part of a number, and another is the same part of another, then the sum is also
the same part of the sum that the one is of the one.
Let the number A be a part of BC, and another number D be the same part of another number EF that
A is of BC.
I say that the sum of A and D is also the same part of the sum of BC and EF that A is of BC.
Since, whatever part A is of BC, D is also the same part of EF, therefore, there
are as many numbers equal to D in EF as there are in BC equal to A.
Divide BC into the numbers equal to A, namely BG and GC, and EF into the
numbers equal to D, namely EH and HF. Then the multitude of BG and GC
equals the multitude of EH and HF.
And, since BG equals A, and EH equals D, therefore the sum of BG and EH also
equals the sum of A and D. For the same reason the sum of GC and HF also
equals the sum of A and D.
Therefore there are as many numbers in BC and EF equal to A and D as there are in BC equal to A.
Therefore, the sum of BC and EF is the same multiple of the sum of A and D that BC is of A.
Therefore, the sum of A and D is the same part of the sum of BC and EF that A is of BC.
Therefore, if a number is part of a number, and another is the same part of another, then the sum is
also the same part of the sum that the one is of the one.
Q.E.D.
This is the first of four propositions that deal with distributivity of division and multiplication over
addition and subtraction. This one says division distributes over addition. Algebraically, if a = b/n and
d = e/n, then a + d = (b + e)/n. As a single equation, this says
This proposition is used in the proofs of five of the next seven propositions.
© 1996
D.E.Joyce
Clark University
Proposition 6
If a number is parts of a number, and another is the same parts of another, then the sum is also
the same parts of the sum that the one is of the one.
Let the number AB be parts of the number C, and another number DE be the same parts of
another number F that AB is of C.
I say that the sum of AB and DE is also the same parts of the sum of C and F that AB is of C.
Divide AB into the parts of C, namely AG and GB, and divide DE into the
parts of F, namely DH, and HE. Then the multitude of AG and GB equals
the multitude of DH and HE.
And since DH is the same part of F that AG is of C, therefore the sum of AG and DH is the
same part of the sum of C and F that AG is of C. For the same reason, the sum of GB and HE is VII.5
the same parts of the sum of C and F that GB is of C.
Therefore the sum of AB and DE is the same parts of the sum of C and F that AB is of C.
Therefore, if a number is parts of a number, and another is the same parts of another, then the sum is
also the same parts of the sum that the one is of the one.
Q.E.D.
This proposition says multiplication by fractions distributes over addition. Algebraically, if a = (m/n)b
and d = (m/n)e then a + d = (m/n)(b + e). As an equation, this says
© 1996
D.E.Joyce
Clark University
Proposition 7
If a number is that part of a number which a subtracted number is of a subtracted number, then
the remainder is also the same part of the remainder that the whole is of the whole.
Let the number AB be that part of the number CD which AE subtracted is of CF subtracted.
I say that the remainder EB is also the same part of the remainder FD that the whole AB is of
the whole CD.
But, by hypothesis, AB is the same part of CD that AE is of CF, therefore AB is the same part of
CD that it is of GF. Therefore GF equals CD.
Subtract CF from each. Then the remainder GC equals the remainder FD.
Now since EB is the same part of GC that AE is of CF, and GC equals FD, therefore EB is the
same part of FD that AE is of CF.
But AB is the same part of CD that AE is of CF, therefore the remainder EB is the same part of
the remainder FD that the whole AB is of the whole CD.
Therefore, if a number is that part of a number which a subtracted number is of a subtracted number,
then the remainder is also the same part of the remainder that the whole is of the whole.
Q.E.D.
This proposition is like VII.5 except it deals with subtraction instead of addition. It says division
© 1996
D.E.Joyce
Clark University
Proposition 8
If a number is the same parts of a number that a subtracted number is of a subtracted number,
then the remainder is also the same parts of the remainder that the whole is of the whole.
Let the number AB be the same parts of the number CD that AE subtracted is of CF subtracted.
I say that the remainder EB is also the same parts of the remainder FD that the whole AB is of
the whole CD.
Now since AL is the same part of CF that GK is of CD, and CD is greater than CF, therefore
GK is also greater than AL.
Then GK is the same part of CD that GM is of CF. Therefore the remainder MK is the same VII.7
part of the remainder FD that the whole GK is of the whole CD.
Again, since EL is the same part of CF that KH is of CD, and CD is greater than CF, therefore
HK is also greater than EL.
Therefore KN is the same part of CF that KH is of CD. Therefore the remainder NH is the same VII.7
part of the remainder FD that the whole KH is of the whole CD.
But the remainder MK was proved to be the same part of the remainder FD that the whole GK
is of the whole CD, therefore the sum of MK and NH is the same parts of DF that the whole
HG is of the whole CD.
But the sum of MK and NH equals EB, and HG equals BA, therefore the remainder EB is the
same parts of the remainder FD that the whole AB is of the whole CD.
Therefore, if a number is the same parts of a number that a subtracted number is of a subtracted
number, then the remainder is also the same parts of the remainder that the whole is of the whole.
Q.E.D.
This proposition says multiplication by fractions distributes over subtraction. Algebraically, if a = (m/n)
b and d = (m/n)e, then a + d = (m/n)(b + e). The sample value taken for m/n in the proof is 2/3.
© 1996
D.E.Joyce
Clark University
Proposition 9
If a number is a part of a number, and another is the same part of another, then alternately,
whatever part of parts the first is of the third, the same part, or the same parts, the second is of
the fourth.
Let the number A be a part of the number BC, and and another number D be the same part of
another number EF that A is of BC.
Divide BC into the numbers equal to A, namely BG and GC, and divide
EF into those equal to D, namely EH and HF. Then the multitude of BG
and GC equals the multitude of EH and HF.
Now, since the numbers BG and GC equal one another, and the numbers EH and HF also
equal one another, while the multitude of BG and GC equals the multitude of EH and HF, VII.5
therefore GC is the same part or parts of HF that BG is of EH, so that, in addition, the sum BC VII.6
is the same part or parts of the sum EF that BG is of EH.
But BG equals A, and EH equals D, therefore BC is the same part or parts of EF that A is of D.
Therefore, if a number is a part of a number, and another is the same part of another, then
alternately, whatever part of parts the first is of the third, the same part, or the same parts, the second
is of the fourth.
Q.E.D.
In this proposition, Euclid shows that if a = b/n, and d = e/n, and if a = (m/n)d, then b = (m/n)e. The
sample value taken for 1/n in the proof is 1/2.
© 1996
D.E.Joyce
Clark University
Proposition 10
If a number is parts of a number, and another is the same parts of another, then alternately,
whatever part of parts the first is of the third, the same part, or the same parts, the second is of
the fourth.
Let the number AB be parts of the number C, and another number DE be the same parts of
another number F.
Divide AB into the parts of C, namely AG and GB, and divide DE into the
parts of F, namely DH and HE. Then the multitude of AG and GB equals
the multitude of DH and HE.
VII.9
For the same reason, C is the same part or the same parts of F as GB is of HE, so that, in VII.5
addition, C is the same part or the same parts of F as AB is of DE. VII.6
Therefore, if a number is parts of a number, and another is the same parts of another, then
alternately, whatever part of parts the first is of the third, the same part, or the same parts, the second
is of the fourth.
Q.E.D.
In this proposition, Euclid shows that if a = (m/n)b, and d = (m/n)e, and if a = (p/q)d, then b = (p/q)e.
The sample value taken for m/n in the proof is 2/3.
© 1996
D.E.Joyce
Clark University
Proposition 11
I say that the remainder EB is to the remainder FD as the whole AB is to the whole CD.
VII.
Since AB is to CD as AE is to CF, therefore AE is the same part or parts of CF
Def.20
as AB is of CD. Therefore the remainder EB is the same part or parts of FD that VII.7
AB is of CD. VII.8
This proposition is the numerical analogue of proposition V.19 for general magnitudes. Algebraically, if
a:c = e:f, then a – e:c – f = a:c.
Note that Euclid only deals with two cases, when AB is a part or parts of CD, and leaves out the other
two, when CD is a part or parts of AB.
© 1996
D.E.Joyce
Clark University
Proposition 12
If any number of numbers are proportional, then one of the antecedents is to one of the
consequents as the sum of the antecedents is to the sum of the consequents.
Let A, B, C, and D be as many numbers as we please in proportion, so that A is to B as C is
to D.
VII.
Since A is to B as C is to D, therefore A is the same part or parts of B
Def.20
as C is of D. Therefore the sum of A and C is the same part or parts VII.5
of the sum of B and D that A is of B. VII.6
Therefore, if any number of numbers are proportional, then one of the antecedents is to one of the
consequents as the sum of the antecedents is to the sum of the consequents.
Q.E.D.
© 1996
D.E.Joyce
Clark University
Proposition 13
If four numbers are proportional, then they are also proportional alternately.
Let the four numbers A, B, C, and D be proportional, so that A is to B as C is to D.
Therefore A is to C as B is to D. VII.Def.20
Therefore, if four numbers are proportional, then they are also proportional alternately.
Q.E.D.
This is the numerical analogue of proposition V.16 for magnitudes. It says that if a : b = c : d, then a :
c = b : d.
This proposition is used frequently in Books VII through IX starting with the next proposition.
© 1996
D.E.Joyce
Clark University
Proposition 14
If there are any number of numbers, and others equal to them in multitude, which taken two
and two together are in the same ratio, then they are also in the same ratio ex aequali.
Let there be as many numbers as we please A, B, and C, and others equal to them in multitude
D, E, and F, which taken two and two are in the same ratio, so that A is to B as D is to E, and
B is to C as E is to F.
Therefore, if there are any number of numbers, and others equal to them in multitude, which taken
two and two together are in the same ratio, then they are also in the same ratio ex aequali.
Q.E.D.
This is the numerical analogue of V.22 for magnitudes. It says that if x1:x2 = y1:y2, x2:x3 = y2:y3, ... , and
x -1:xn = y -1:yn, then x1:xn = y1:yn. Euclid takes n to be 3 in his proof.
n n
The proof is straightforward, and a simpler proof than the one given in V.22 for magnitudes. Note that at
one point, the missing analogue of proposition V.11 is used: from the two proportions B : E = C : F and
B : E = A : D, the conclusion A : D = C : F is drawn. Similar missing analogues of propositions from
Book V are used in other proofs in book VII. See, for instance, VII.19 where V.7 and V.9 are used.
This proposition is used occasionally in Books VIII and IX starting with VIII.1.
© 1996
D.E.Joyce
Clark University
Proposition 15
If a unit measures any number, and another number measures any other number the same
number of times, then alternately, the unit measures the third number the same number of times
that the second measures the fourth.
Let the unit A measure any number BC, and let another number D measure any other number
EF the same number of times.
I say that, alternately also, the unit measures the number D the same number of times that BC
measures EF.
Divide BC into the units in it, BG, GH, and HC, and divide EF into the numbers EK, KL, and
LF equal to D. Then the multitude of BG, GH, and HC equals the multitude of EK, KL, and
LF.
And, since the units BG, GH, and HC equal one another, and the numbers EK, KL, and LF
also equal one another, while the multitude of the units BG, GH, and HC equals the multitude
of the numbers EK, KL, and LF, therefore the unit BG is to the number EK as the unit GH is to
the number KL, and as the unit HC is to the number LF.
Since one of the antecedents is to one of the consequents as the sum of the antecedents is to VII.12
the sum of the consequents, therefore the unit BG is to the number EK as BC is to EF.
But the unit BG equals the unit A, and the number EK equals the number D. Therefore the unit
A is to the number D as BC is to EF. Therefore the unit A measures the number D the same
number of times that BC measures EF.
Therefore, if a unit number measures any number, and another number measures any other number
the same number of times, then alternately, the unit measures the third number the same number of
times that the second measures the fourth.
Q.E.D.
This proposition expresses the commutativity of multiplication. If a number e is b times d, that is, 1
measures b the same number of times that b measures d, then e also is d times b. In other words, bd = db.
The next proposition states this commutativity more explicitly.
This proposition is used in the next proposition and a few others in Books VII and IX.
© 1996
D.E.Joyce
Clark University
Proposition 16
If two numbers multiplied by one another make certain numbers, then the numbers so produced
equal one another.
Let A and B be two numbers, and let A multiplied by B make C, and B multiplied by A make
D.
But the unit E measures the number B the same number of times that A measures C, therefore
A measures each of the numbers C and D the same number of times.
Therefore C equals D.
Therefore, if two numbers multiplied by one another make certain numbers, then the numbers so
produced equal one another.
Q.E.D.
This proposition describes the commutativity mentioned in the last proposition more explicitly, ab = ba.
It is used in VII.18 and a few others in Book VII.
© 1996
D.E.Joyce
Clark University
Proposition 17
If a number multiplied by two numbers makes certain numbers, then the numbers so produced
have the same ratio as the numbers multiplied.
Let the number A multiplied by the two numbers B and C make D and E.
I say that B is to C as D is to E.
But the unit F also measures the number A according to the units in it, therefore the unit F
measures the number A the same number of times that B measures D. Therefore the unit F VII.Def.20
is to the number A as B is to D.
VII.
For the same reason the unit F is to the number A as C is to E, therefore B is to D as C is to Def.20
E.
(V.11)
Therefore, if a number multiplied by two numbers makes certain numbers, then the numbers so
produced have the same ratio as the numbers multiplied.
Q.E.D.
Algebraically, b : c = ab : ac.
This proposition is used very frequently in Books VII through IX starting with the next proposition.
© 1996
D.E.Joyce
Clark University
Proposition 18
If two numbers multiplied by any number make certain numbers, then the numbers so produced
have the same ratio as the multipliers.
Let two numbers A and B multiplied by any number C make D and E.
I say that A is to B as D is to E.
Therefore, if two numbers multiplied by any number make certain numbers, then the numbers so
produced have the same ratio as the multipliers.
Q.E.D.
b : c = ab : ac,
this one says
b : c = ba : ca.
The only difference is the order of multiplication, but VII.16 says multiplication is commutative, so that
order is irrelevant.
This proposition is used in the next one and occasionally in Book VIII.
© 1996
D.E.Joyce
Clark University
Proposition 19
If four numbers are proportional, then the number produced from the first and fourth equals
the number produced from the second and third; and, if the number produced from the first
and fourth equals that produced from the second and third, then the four numbers are
proportional.
Let A, B, C, and D be four numbers in proportion, so that A is to B as C is to D, and let A
multiplied by D make E, and let B multiplied by C make F.
I say that A is to B as C is to D.
VII.17
But G is to E as C is to D, and G is to F as A is to B, therefore A is to B as C is to D. VII.18
(V.11)
Therefore, if four numbers are proportional, then the number produced from the first and fourth
equals the number produced from the second and third; and, if the number produced from the first
and fourth equals that produced from the second and third, then the four numbers are proportional.
Q.E.D.
Algebraically, a : b = c : d if and only if ad = bc. These algebraic expressions are meaningful when the
variables are all numbers, but not when they are magnitudes in general. They can be interpreted,
however, when they are lines, and proposition VI.16 is the analogue in that case.
Twice in this proof Euclid makes conclusions about proportions for numbers that he has neither stated
nor proved. These places are indicated by (V.11), (V.9), and (V.7) in the margins, the analogous
justifications for magnitudes. Some of the propositions in Book V for magnitudes are stated in proved in
Book VII for numbers, in particular, V.16 and VII.13 correspond, and V.22 and VII.14 correspond. But
many of the propositions in Book V have no analogue in Book VII, such as V.11, V.9, and V.7.
Now it could be that Euclid considered the missing statements as being obvious, as Heath claims, but
being obvious is usually not a reason for Euclid to omit a proposition. Furthermore, other propositions in
the next three books assume properties about proportions of numbers without having proofs of those
propositions. One explanation is that the books on number theory, including this one, are older, and
when the material in Book V was developed by Eudoxus, or when it was incorporated into the Elements
by Euclid, more careful attention was made to fundamental propositions like V.7, V.9, and V.11.
This proposition is used frequently in Books VII and IX starting with VII.24.
© 1996
D.E.Joyce
Clark University
Proposition 20
The least numbers of those which have the same ratio with them measure those which have the
same ratio with them the same number of times; the greater the greater; and the less the less.
Let CD and EF be the least numbers of those which have the same ratio with A and B.
Now, since the numbers CG and GD equal one another, and the
numbers EH and HF also equal one another, while the multitude of
CG and GD equals the multitude of EH and HF, therefore CG is to
EH as GD is to HF.
Since one of the antecedents is to one of the consequents as the sum of the antecedents is VII.12
to the sum of the consequents, therefore CG is to EH as CD is to EF.
Therefore CG and EH are in the same ratio with CD and EF, being less than they, which is
impossible, for by hypothesis CD and EF are the least numbers of those which have the VII.4
same ratio with them. Therefore CD is not parts of A, therefore it is a part of it.
And EF is the same part of B that CD is of A, therefore CD measures A the same number VII.13
of times that EF measures B. VII.Def.20
Therefore, the least numbers of those which have the same ratio with them measure those which have
the same ratio with them the same number of times; the greater the greater; and the less the less.
Q.E.D.
This proposition says that given a ratio a:b, if c:d is the same ratio and the least among all those ratios
with the same ratio, then, first of all, c divides a, and d divides b, but also, c divides a the same number
of times that d divides b. For example, the ratio 91:132 is the same ratio as 7:11, which is least among
all the ratios equal to 91:132, that is to say 91:132 reduces to 7:11 in lowest terms, therefore 7 divides 91
the same number of times that 11 divides 132, namely, 13 times.
The proof goes along like this. Suppose a:b reduces to c:e in lowest terms. In order to show that c
divides a, assume that it doesn't, assume that c = (m/n)a. Since a:b is the same ratio as c:e, therefore
e = (m/n)d. But then c/m = (1/n)a, and e/m = (1/n)b. Therefore c/m:e/m is the same ratio as a:b, which
shows that c:e is not in lowest terms, a contradiction. Therefore c does divide a, and e divides b the same
number of times.
This proposition is used frequently in Books VII through IX starting with the next proposition.
© 1996, 2002
D.E.Joyce
Clark University
Proposition 21
Numbers relatively prime are the least of those which have the same ratio with them.
Let A and B be numbers relatively prime.
I say that A and B are the least of those which have the same ratio with them.
If not, there are some numbers less than A and B in the same
ratio with A and B. Let them be C and D.
Since, then, the least numbers of those which have the same
ratio measure those which have the same ratio the same
number of times, the greater the greater, and the less the VII.20
less, that is, the antecedent the antecedent and the
consequent the consequent, therefore C measures A the
same number of times that D measures B.
Let there be as many units in E as the times that C measures A. Then D also measures B
according to the units in E.
Therefore E measures A and B which are relatively prime, which is impossible. VII.Def.12
Therefore there are no numbers less than A and B which are in the same ratio with A and
B. Therefore A and B are the least of those which have the same ratio with them.
Therefore, numbers relatively prime are the least of those which have the same ratio with them.
Q.E.D.
The next proposition is the converse of this one. Together they say that a ratio a:b is reduced to lowest
terms if and only if a is relatively prime to b.
Although it appears that this proposition is pairs of numbers and their ratios, it is used in proposition
VII.33 with any quantity of numbers. Stated in terms of three numbers a, b, and c, that proposition says
that of all triples with the same ratio as a, b, and c, have, the triple of relatively prime numbers is least.
This proposition is used frequently in Books VII through IX starting with VII.24.
© 1996
D.E.Joyce
Clark University
Proposition 22
The least numbers of those which have the same ratio with them are relatively prime.
Let A and B be the least numbers of those which have the same ratio with them.
Since C measures A according to the units in D, therefore C multiplied by D makes A. For VII.Def.15
the same reason C multiplied by E makes B.
Thus the number C multiplied by the two numbers D and E makes A and B, therefore D is VII.17
to E as A is to B.
Therefore D and E are in the same ratio with A and B, being less than they, which is
impossible. Therefore no number measures the numbers A and B.
Therefore, the least numbers of those which have the same ratio with them are relatively prime.
Q.E.D.
This proposition is the converse of the last one. Together they say that a ratio a:b is reduced to lowest
© 1996
D.E.Joyce
Clark University
Proposition 23
If two numbers are relatively prime, then any number which measures one of them is relatively
prime to the remaining number.
Let A and B be two numbers relatively prime, and let any number C measure A.
Therefore no number measures the numbers C and B. Therefore C and B are relatively
prime.
Therefore, if two numbers are relatively prime, then any number which measures one of them is
relatively prime to the remaining number.
Q.E.D.
© 1996
D.E.Joyce
Clark University
Proposition 24
If two numbers are relatively prime to any number, then their product is also relatively prime
to the same.
Let the two numbers A and B [each] be relatively prime to a number C, and let A
multiplied by B make D.
But, if the product of the extremes equal that of the means, then the four numbers are VII.19
proportional. Therefore E is to A as B is to F.
But A and E are relatively prime, numbers which are relatively prime are also the least of
those which have the same ratio, and the least numbers of those which have the same ratio VII.21
with them measure those which have the same ratio the same number of times, the greater
the greater, and the less the less, that is, the antecedent the antecedent and the consequent VII.20
the consequent, therefore E measures B.
But it also measures C, therefore E measures B and C which are relatively prime, which is VII.Def.12
impossible.
Therefore no number measures the numbers C and D. Therefore C and D are relatively
prime.
Therefore, if two numbers are relatively prime to any number, then their product is also relatively
prime to the same.
Q.E.D.
Assume that two numbers a and b are each relatively prime to a third number c.
Suppose their product ab is not relatively prime to c. Then there is some number e (greater than 1) that
divides both ab and c. Now, since e divides c, and c is relatively prime to a, therefore, by VII.23, e is
also relatively prime to a.
Let f be the number ab/e. Then e:a = b:f. Since e and a are relatively prime, then, by VII.21, e:a is in
lowest terms. Therefore, by VII.20, e divides b. But then e divides both b and c contradicting the
assumption that b and c are relatively prime.
© 1996, 2002
D.E.Joyce
Clark University
Proposition 25
If two numbers are relatively prime, then the product of one of them with itself is relatively
prime to the remaining one.
Let A and B be two numbers relatively prime, and let A multiplied by itself make C.
Make D equal to A.
Since A and B are relatively prime, and A equals D, therefore D and B are
also relatively prime. Therefore each of the two numbers D and A is VII.24
relatively prime to B. Therefore the product of D and A is also relatively
prime to B.
But the number which is the product of D and A is C. Therefore C and B are
relatively prime.
Therefore, if two numbers are relatively prime, then the product of one of them with itself is relatively
prime to the remaining one.
Q.E.D.
This is a special case of the previous proposition. It is used in VII.27 and IX.15.
© 1996
D.E.Joyce
Clark University
Proposition 26
If two numbers are relatively prime to two numbers, both to each, then their products are also
relatively prime.
Let the two numbers A and B be relatively prime to the two numbers C and D, both to each,
and let A multiplied by B make E, and let C multiplied by D make F.
Therefore the product of C and D is also relatively prime to E. But the product of C and D is VII.24
F. Therefore E and F are relatively prime.
Therefore, if two numbers are relatively prime to two numbers, both to each, then their products are
also relatively prime.
Q.E.D.
The proof of this proposition uses proposition VII.24 twice. If a and b are both relatively prime to both c
and d, then so is their product ab. Now since c and d are both relatively prime to ab, therefore so is their
product cd.
© 1996, 2002
D.E.Joyce
Clark University
Proposition 27
If two numbers are relatively prime, and each multiplied by itself makes a certain number, then
the products are relatively prime; and, if the original numbers multiplied by the products make
certain numbers, then the latter are also relatively prime.
Let A and B be two relatively prime numbers, let A multiplied by itself make C, and multiplied
by C make D, and let B multiplied by itself make E, and multiplied by E make F.
I say that C and E are relatively prime, and that D and F are
relatively prime.
Since, then, the two numbers A and C are relatively prime to the two numbers B and E, both to
each, therefore the product of A and C is relatively prime to the product of B and E. And the VII.26
product of A and C is D, and the product of B and E is F.
Therefore, if two numbers are relatively prime, and each multiplied by itself makes a certain number,
then the products are relatively prime; and, if the original numbers multiplied by the products make
certain numbers, then the latter are also relatively prime.
Q.E.D.
The proposition states that if two numbers are relatively prime, then their powers are also relatively
prime. Explicitly, it only says that their squares are relatively prime, and their cubes are relatively prime,
but the way it is used in VIII.2, any powers need to be relatively prime.
The proof of this proposition uses the last two propositions. Assume that a and b are relatively prime.
Then applying VII.25 twice, we first get a2 and b relatively prime, then we get a2 and b2 relatively
prime.
Again, by VII.25, a and b2 are relatively prime. Now, a is relatively prime to b2, and b is relatively
prime to a2, so by VII.26, a3 is relatively prime to b3.
© 1996, 2002
D.E.Joyce
Clark University
Proposition 28
If two numbers are relatively prime, then their sum is also prime to each of them; and, if the
sum of two numbers is relatively prime to either of them, then the original numbers are also
relatively prime.
Let two relatively prime numbers AB and BC be added.
I say that their sum AC is also relatively prime to each of the numbers AB
and BC.
Since then D measures CA and AB, therefore it also measures the remainder
BC. But it also measures BA, therefore D measures AB and BC which are VII.Def.12
relatively prime, which is impossible.
Now, since D measures each of the numbers AB and BC, therefore it also
measures the whole CA. But it measures AB, therefore D measures CA and VII.Def.12
AB which are relatively prime, which is impossible.
Therefore, if two numbers are relatively prime, then their sum is also prime to each of them; and, if
the sum of two numbers is relatively prime to either of them, then the original numbers are also
relatively prime.
Q.E.D.
© 1996
D.E.Joyce
Clark University
Proposition 29
Any prime number is relatively prime to any number which it does not measure.
Let A be a prime number, and let it not measure B.
Now, since C measures B and A, therefore it also measures A which is prime, though it is not the same
as it, which is impossible. Therefore no number measures B and A.
Therefore, any prime number is relatively prime to any number which it does not measure.
Q.E.D.
This proposition is used in the next one and in propositions IX.12 and IX.36.
© 1996
D.E.Joyce
Clark University
Proposition 30
If two numbers, multiplied by one another make some number, and any prime number
measures the product, then it also measures one of the original numbers.
Let the two numbers A and B multiplied by one another make C, and let any
prime number D measure C.
Therefore D is to A as B is to E. VII.19
But D and A are relatively prime, relatively prime numbers are also least,
and the least measure the numbers which have the same ratio the same VII.21
number of times, the greater the greater and the less the less, that is, the
antecedent the antecedent and the consequent the consequent, therefore D VII.20
measures B.
Similarly we can also show that, if D does not measure B, then it measures
A. Therefore D measures one of the numbers A or B.
Therefore, if two numbers, multiplied by one another make some number, and any prime number
measures the product, then it also measures one of the original numbers.
Q.E.D.
This proposition states that if p is a prime number, then whenever p divides a product of two numbers,
then it divides at least one of them. This is actually a property that characterizes prime numbers, that is
to say, no composite number has this property. (For if c is a composite number, c ab, so c divides the
product but it doesn't divide either factor.)
The form of the proof is interesting. Euclid shows that if d doesn't divide a, then d does divide b, and
similarly, if d doesn't divide b, then d does divide a. Therefore, it divides either one or the other.
Suppose d does not divide a. Then, by VII.29, d is relatively prime to a. Let e be the number ab/d. Then
d:a = b:e. By VII.21, the ratio d:a is in lowest terms, and so, by VII.20, d divides b.
© 1996, 2002
D.E.Joyce
Clark University
Proposition 31
But if it is composite, some number measures it. Let a number C measure it. VII.Def.11,13
Therefore some prime number will be found which measures the one before it, which
also measures A.
Euclid does not explain why there can't be an infinite sequence of numbers where each number divides
the previous. He simply says that is impossible. Some justification is required such as the principle
Euclid uses elsewhere that any decreasing sequence of numbers is finite.
This proposition is used in the next one and in propositions IX.13 and IX.20.
© 1996
D.E.Joyce
Clark University
Proposition 32
After the previous proposition, this one really doesn't need to be stated at all.
© 1996
D.E.Joyce
Clark University
Proposition 33
Given as many numbers as we please, to find the least of those which have the same ratio with
them.
Let A, B, and C be the given numbers, as many as we please.
It is required to find the least of those which have the same ratio with A, B, and C.
Therefore the numbers E, F, and G measure the numbers A, B, and C respectively VII.16
according to the units in D. Therefore E, F, and G measure A, B, and C the same number
of times. Therefore E, F, and G are in the same ratio with A, B, and C. VII.Def.20
I say next that they are the least that are in that ratio.
If E, F, and G are not the least of those which have the same ratio with A, B, and C, then
there are numbers less than E, F, and G in the same ratio with A, B, and C. Let them be H,
K, and L. Therefore H measures A the same number of times that the numbers K and L
measure the numbers B and C respectively.
Let there be as many units in M as the times that H measures A. Then the numbers K and L
also measure the numbers B and C respectively according to the units in M.
Now, since H measures A according to the units in M, therefore H multiplied by M makes VII.Def.15
A. For the same reason also E multiplied by D makes A.
Therefore the product of E and D equals the product of H and M. Therefore E is to H as M VII.19
is to D.
But E is greater than H, therefore M is also greater than D. And it measures A, B, and C,
which is impossible, for by hypothesis, D is the greatest common measure of A, B, and C.
Therefore there cannot be any numbers less than E, F, and G which are in the same ratio
with A, B, and C. Therefore E, F, and G are the least of those which have the same ratio
with A, B, and C.
Q.E.D.
This proposition is unusual in that it discusses a ratio a:b:c of three (or more) numbers. It also has the
proportion a:b:c = e:f:g. These multiterm ratios and proportions may have been left over from an earlier
time. Euclid argues that the proportion holds because e, f, and g measure a, b, and c, respectively, the
same number, d, times. By the definition of proportion, that observation directly implies
The desired proportion, a:b:c = e:f:g is an alternate form of that multiple proportion. See V.Def.13 for
the definition of alternate ratios.
This proposition is used in the next one and several propositions in Book VIII starting with VIII.6.
© 1996, 2002
D.E.Joyce
Clark University
Proposition 34
Let there be as many units in E as the times that A measures D, and as many units in F as
the times that B measures D.
VII.
Then A multiplied by E makes D, and B multiplied by F makes D. Therefore the product Def.15
of A and E equals the product of B and F. Therefore A is to B as F is to E.
VII.19
But A and B are relatively prime, primes are also least, and the least measure the numbers
VII.21
which have the same ratio the same number of times, the greater the greater and the less VII.20
the less, therefore B measures E as the consequent the consequent.
Therefore A and B do not measure any number less than C. Therefore C is the least that is
measured by A and B.
Next, let A and B not be relatively prime. Take F and E, the least numbers of those which
VII.33
have the same ratio with A and B. Therefore the product of A and E equals the product of VII.19
B and F.
Let there be as many units in G as the times that A measures D, and as many units in H as
the times that B measures D.
Then A multiplied by G makes D, and B multiplied by H makes D. Therefore the product VII.19
of A and G equals the product of B and H. Therefore A is to B as H is to G.
But F and E are least, and the least measure the numbers which have the same ratio the VII.20
same number of times, the greater the greater and the less the less, therefore E measures G.
But E measures G, therefore C also measures D, the greater the less, which is impossible.
Therefore A and B do not measure any number less than C. Therefore C is the least that is
measured by A and B.
Q.E.D.
The least common multiple of two numbers a and b is the smallest number that they both divide. It is
denoted LCM(a, b). This proposition construct it as the product divided by the greatest common divisor:
Let a and b be the two numbers. There are two cases depending on whether they are relatively prime or
not.
Case 1. Suppose a and b are relatively prime. An indirect proof shows that their least common multiple
is their product ab. If not, then there is a smaller number d which both a and b divide. Since a:b = (d/b):
(d/a), and a:b is in lowest terms (since a and b are relatively prime), therefore b divides d/a. Also, b:(d/
a) = ab:d, so ab divides d, but d is smaller than ab, a contradiction. Thus, when a and b are relatively
prime, their least common multiple is their product.
Case 2. Suppose a and b are not relatively prime. Reduce the ratio a:b to its lowest terms f:e using the
previous proposition VII.33. Then ae = bf. Let c denote this product. (Note that f = a/GCD(a, b), and
e = b/GCD(a, b), so c = ab/GCD(a, b).) Both a and b divide c, therefore c is a common multiple of a and
b. Suppose that it's not the least common multiple. Then there is a smaller number d which both a and b
divide. Now
and f:e is in lowest terms, therfore e divides d/a. But e:(d/a) = ae:d, therefore ae also divides d. But
c = ae, and d is less than c, a contradiction. Thus, LCM(a, b) = ab/LCM(a, b).
© 1996, 2002
D.E.Joyce
Clark University
Proposition 35
If two numbers measure any number, then the least number measured by them also measures
the same.
Let the two numbers A and B measure any number CD, and let E be the least that they measure.
Therefore, if two numbers measure any number, then the least number measured by them also
measures the same.
Q.E.D.
Assume both a and b divide c. Let e be their least common multiple. Suppose that e does not divide c.
Then repeatedly subtract e from c to get c = ke + f, where the remainder f is less than e and k is some
number. Since a and b both divide c and e, they also divide f making f a smaller common multiple than
the least common multiple e, a contradiction. Thus the least common multiple also divides c.
© 1996, 2002
D.E.Joyce
Clark University
Proposition 36
Take D the least number measured by the two numbers A and B. VII.34
Since A, B, and C measure E, therefore A and B measure E. Therefore the least number VII.35
measured by A and B also measures E.
But D is the least number measured by A and B, therefore D measures E, the greater the less,
which is impossible.
Therefore A, B, and C do not measure any number less than D. Therefore D is the least that A,
B, and C measure.
Since A, B, and C measure F, therefore A and B measure F. Therefore the least number
measured by A and B also measures F. But D is the least number measured by A and B, VII.35
therefore D measures F. But C also measures F, therefore D and C measure F, so that the least
number measured by D and C also measures F.
But E is the least number measured by C and D, therefore E measures F, the greater the less,
which is impossible.
Therefore A, B, and C do not measure any number which is less than E. Therefore E is the
least that is measured by A, B, and C.
Q.E.D.
© 1996
D.E.Joyce
Clark University
Proposition 37
If a number is measured by any number, then the number which is measured has a part called
by the same name as the measuring number.
Let the number A be measured by any number B.
Therefore, alternately, the unit D measures the number B the same number of times as C
measures A. Therefore, whatever part the unit D is of the number B, the same part is C of A
also. But the unit D is a part of the number B called by the same name as it, therefore C is also VII.15
a part of A called by the same name as B, so that A has a part C which is called by the same
name as B.
Therefore, if a number is measured by any number, then the number which is measured has a part
called by the same name as the measuring number.
Q.E.D.
This proposition says that if b divides a, then a has a one-bth part (namely, a/b). For example, 3 divides
12, therefore 12 has a one-third part.
© 1996
D.E.Joyce
Clark University
Proposition 38
If a number has any part whatever, then it is measured by a number called by the same name
as the part.
Let the number A have any part whatever, B, and let C be a number called by the same name
as the part B.
Therefore, alternately, the unit D measures the number B the same number of times that C VII.15
measures A. Therefore C measures A.
Therefore, if a number has any part whatever, then it is measured by a number called by the same
name as the part.
Q.E.D.
This proposition says that if a has a one-cth part of a, then c divides a. For example, 12 has a one-third
part, 3 divides 12. This is a converse of the last proposition.
© 1996
D.E.Joyce
Clark University
Proposition 39
To find the number which is the least that has given parts.
Let A, B, and C be the given parts.
It is required to find the number which is the least that will have the parts A, B, and C.
But A, B, and C are parts called by the same name as D, E, and F, therefore G has the parts A,
B, and C.
If not, there is some number H less than G which has the parts A, B, and C.
Since H has the parts A, B, and C, therefore H is measured by numbers called by the same
name as the parts A, B, and C. But D, E, and F are numbers called by the same name as the VII.38
parts A, B, and C, therefore H is measured by D, E, and F.
And it is less than G, which is impossible. Therefore there is no number less than G that has
the parts A, B, and C.
Q.E.D.
The wording of the proposition is somewhat unclear, but an example will show its intent.
Suppose you want to find the smallest number with given parts, say, a fourth part and a sixth part. Then
take the LCM(4,6) which is 12. The number 12 has a 1/4 part, namely 3, and a 1/6 part, namely 2.
© 1996, 2002
D.E.Joyce
Clark University
Table of contents
● Propositions (27)
Propositions
Proposition 1.
If there are as many numbers as we please in continued proportion, and the extremes of them are
relatively prime, then the numbers are the least of those which have the same ratio with them.
Proposition 2.
To find as many numbers as are prescribed in continued proportion, and the least that are in a
given ratio.
Corollary. If three numbers in continued proportion are the least of those which have the same
ratio with them, then the extremes are squares, and, if four numbers, cubes.
Proposition 3.
If as many numbers as we please in continued proportion are the least of those which have the
same ratio with them, then the extremes of them are relatively prime.
Proposition 4.
Given as many ratios as we please in least numbers, to find numbers in continued proportion
which are the least in the given ratios.
Proposition 5.
Plane numbers have to one another the ratio compounded of the ratios of their sides.
Proposition 6.
If there are as many numbers as we please in continued proportion, and the first does not measure
the second, then neither does any other measure any other.
Proposition 7.
If there are as many numbers as we please in continued proportion, and the first measures the
last, then it also measures the second.
Proposition 8.
If between two numbers there fall numbers in continued proportion with them, then, however
many numbers fall between them in continued proportion, so many also fall in continued
proportion between the numbers which have the same ratios with the original numbers.
Proposition 9.
If two numbers are relatively prime, and numbers fall between them in continued proportion,
then, however many numbers fall between them in continued proportion, so many also fall
between each of them and a unit in continued proportion.
Proposition 10.
If numbers fall between two numbers and a unit in continued proportion, then however many
numbers fall between each of them and a unit in continued proportion, so many also fall between
the numbers themselves in continued proportion.
Proposition 11.
Between two square numbers there is one mean proportional number, and the square has to the
square the duplicate ratio of that which the side has to the side.
Proposition 12.
Between two cubic numbers there are two mean proportional numbers, and the cube has to the
cube the triplicate ratio of that which the side has to the side.
Proposition 13.
If there are as many numbers as we please in continued proportion, and each multiplied by itself
makes some number, then the products are proportional; and, if the original numbers multiplied
by the products make certain numbers, then the latter are also proportional.
Proposition 14.
If a square measures a square, then the side also measures the side; and, if the side measures the
side, then the square also measures the square.
Proposition 15.
If a cubic number measures a cubic number, then the side also measures the side; and, if the side
measures the side, then the cube also measures the cube.
Proposition 16.
If a square does not measure a square, then neither does the side measure the side; and, if the side
does not measure the side, then neither does the square measure the square.
Proposition 17.
If a cubic number does not measure a cubic number, then neither does the side measure the side;
and, if the side does not measure the side, then neither does the cube measure the cube.
Proposition 18.
Between two similar plane numbers there is one mean proportional number, and the plane
number has to the plane number the ratio duplicate of that which the corresponding side has to
the corresponding side.
Proposition 19.
Between two similar solid numbers there fall two mean proportional numbers, and the solid
number has to the solid number the ratio triplicate of that which the corresponding side has to the
corresponding side.
Proposition 20.
If one mean proportional number falls between two numbers, then the numbers are similar plane
numbers.
Proposition 21.
If two mean proportional numbers fall between two numbers, then the numbers are similar solid
numbers.
Proposition 22.
If three numbers are in continued proportion, and the first is square, then the third is also square.
Proposition 23.
If four numbers are in continued proportion, and the first is a cube, then the fourth is also a cube.
Proposition 24.
If two numbers have to one another the ratio which a square number has to a square number, and
the first is square, then the second is also a square.
Proposition 25.
If two numbers have to one another the ratio which a cubic number has to a cubic number, and
the first is a cube, then the second is also a cube.
Proposition 26.
Similar plane numbers have to one another the ratio which a square number has to a square
number.
Proposition 27.
Similar solid numbers have to one another the ratio which a cubic number has to a cubic number.
Select topic
Book VIII
introduction
© 1996
D.E.Joyce
Clark University
Proposition 1
If there are as many numbers as we please in continued proportion, and the extremes of them
are relatively prime, then the numbers are the least of those which have the same ratio with
them.
Let there be as many numbers as we please, A, B, C, and D, in continued proportion, and let
the extremes of them, A and D, be relatively prime.
I say that A, B, C, and D are the least of those which have the same ratio with them.
If not, let E, F, G, and H be less than A, B, C, and D, and in the same ratio with them.
But A and D are relatively prime, numbers which are relatively prime are also least, and the
least numbers measure those which have the same ratio the same number of times, the greater VII.21
the greater and the less the less, that is, the antecedent the antecedent and the consequent the VII.20
consequent. Therefore A measures E, the greater the less, which is impossible.
Therefore E, F, G, and H, which are less than A, B, C, and D, are not in the same ratio with
them. Therefore A, B, C, and D are the least of those which have the same ratio with them.
Therefore, if there are as many numbers as we please in continued proportion, and the extremes of
them are relatively prime, then the numbers are the least of those which have the same ratio with them.
Q.E.D.
Euclid doesn't define a continued proportion. In this proposition we consider only continued proportions
with a constant ratio, proportions of the form
In proposition VIII.4 continued proportions that don't have a constant ratio will be considered.
A modern expression for this situation is to say that the numbers a1, a2, a3, ..., an-1:a are in a geometric
n
progression or a geometric sequence. The ratio of any consecutive pair in a geometric progression is
constant.
Many of the propositions in Books VIII and IX treat geometric progressions. The sum of a geometric
progression is found in proposition IX.35.
This proposition says that if the end numbers are relatively prime in a continued proportion with
constant ratio, then there is no continued proportion of the same length and same ratio having smaller
numbers. We could say that the continued proportion is in lowest terms. The proposition generalizes
VII.21 when there are only two numbers in continued proportion, that is, a ratio. VII.21 says if the two
numbers are relatively prime, then the ratio is in lowest terms.
Notice that the example of a continued proportion given above, 1250:750 = 750:450 = 450:270 =
270:162, is not in lowest terms, since all the numbers may be halved to get continued proportion of the
same length and same ratio but with smaller numbers.
This proposition is used in the next one and in VIII.9. The converse of this proposition is VIII.3.
© 1996, 2002
D.E.Joyce
Clark University
Proposition 2
To find as many numbers as are prescribed in continued proportion, and the least that are in a
given ratio.
Let the ratio of A to B be the given ratio in least numbers.
It is required to find numbers in continued proportion, as many as may be prescribed, the least
that are in the ratio of A to B.
Again, since A multiplied by B makes D, and B multiplied by itself makes E, therefore the
numbers A and B multiplied by B make the numbers D and E respectively.
But C is to D as A is to B, therefore A is to B as F is to G.
I say next that they are the least numbers that are so.
Since A and B are the least of those which have the same ratio with them, and the least of VII.22
those which have the same ratio are relatively prime, therefore A and B are relatively prime.
And the numbers A and B multiplied by themselves respectively make the numbers C and E,
and multiplied by the numbers C and E respectively make the numbers F and K, therefore C VII.27
and E and F and K are relatively prime respectively.
But, if there are as many numbers as we please in continued proportion, and the extremes of
them are relatively prime, then they are the least of those which have the same ratio with VIII.1
them. Therefore C, D, and E and F, G, H and K are the least of those which have the same
ratio with A and B.
Q.E.D.
Corollary.
From this it is manifest that, if three numbers in continued proportion are the least of those which have
the same ratio with them, then the extremes are squares, and, if four numbers, cubes.
The problem is to construct n numbers in a continued proportion in lowest terms with a given constant
ratio. If the ratio is a:b in lowest terms, then the numbers in the continued proportion are
For instance, the five numbers in a continued proportion in lowest terms with a ratio of 2:3 form the
sequence 24, 23 . 3, 22 . 32, 2 . 33, and 34, that is, the sequence 16, 24, 36, 54, and 81.
The proof is not difficult. First, adjacent terms do have the correct ratio. Also, since a and b are
n-1 n-1
relatively prime, proposition VII.27 implies that the end terms a and b are relatively prime. The
This proposition and its corollary are used in several propositions in Book VIII starting with the next and
in proposition IX.15 in the next book.
© 1996, 2002
D.E.Joyce
Clark University
Proposition 3
If as many numbers as we please in continued proportion are the least of those which have the
same ratio with them, then the extremes of them are relatively prime.
Let as many numbers as we please, A, B, C, and D, in continued proportion be the least of
those which have the same ratio with them.
Take two numbers E and F, the least that are in the ratio of A, B, C and D, then three others VII.33
G, H and K with the same property, and others, more by one continually, until the
multitude taken becomes equal to the multitude of the numbers A, B, C, and D. Let them be
VIII.2
L, M, N, and O.
VII.22
Since E and F are the least of those which have the same ratio with them, therefore they
are relatively prime. And, since the numbers E and F multiplied by themselves respectively
VIII.2,
make the numbers G and K, and multiplied by the numbers G and K respectively make the
Cor
numbers L and O, therefore both G and K and L and O are relatively prime.
VII.27
And, since A, B, C, and D are the least of those which have the same ratio with them, while
L, M, N, and O are the least that are in the same ratio with A, B, C, and D, and the
multitude of the numbers A, B, C, and D equals the multitude of the numbers L, M, N, and
O, therefore the numbers A, B, C, and D equal the numbers L, M, N, and O respectively.
Therefore A equals L, and D equals O.
And L and O are relatively prime. Therefore A and D are also relatively prime.
Therefore, if as many numbers as we please in continued proportion are the least of those which have
the same ratio with them, then the extremes of them are relatively prime.
Q.E.D.
This proposition, the converse of VIII.1, says if a continued proportion with constant ratio is in lowest
terms, then its end numbers are relatively prime.
The proof begins by using the previous proposition VIII.2 to construct the continued proportion in
lowest terms, which must be the same as the given continued proportion, and it has relatively prime end
numbers.
© 1996, 2002
D.E.Joyce
Clark University
Proposition 4
Given as many ratios as we please in least numbers, to find numbers in continued proportion
which are the least in the given ratios.
Let the given ratios in least numbers be that of A to B, that of C to D, and that of E to F.
It is required to find numbers in continued proportion which are the least that are in the
ratio of A to B, in the ratio of C to D, and in the ratio of E to F.
Let A measure H as many times as B measures G, and let D measure K as many times as C
measures G.
I say next that they are also the least that have this property.
If H, G, K, and L are not the least numbers continuously proportional in the ratios of A to
B, of C to D, and of E to F, let them be N, O, M, and P.
Then since A is to B as N is to O, while A and B are least, and the least numbers measure
those which have the same ratio the same number of times, the greater the greater and the VII.20
less the less, that is, the antecedent the antecedent and the consequent the consequent,
therefore B measures O.
For the same reason C also measures O. Therefore B and C measure O. Therefore the least VII.35
number measured by B and C also measures O.
But G is the least number measured by B and C, therefore G measures O, the greater the
less, which is impossible. Therefore there are no numbers less than H, G, K, and L which
are continuously in the ratio of A to B, of C to D, and of E to F.
Take M, the least number measured by E and K. Let H and G measure N and O as many
times as K measures M, respectively, and let F measure P as many times as E measures M.
Again, since E measures M the same number of times that F measures P, therefore E is to
VII.13
F as M is to P. Therefore N, O, M, and P are continuously proportional in the ratios of A to VII.Def.20
B, of C to D, and of E to F.
I say next that they are also the least that are in the ratios A, B, C, D, E, and F.
If not, there are numbers less than N, O, M, and P continuously proportional in the ratios
A, B, C, D, E, and F. Let them be Q, R, S, and T.
Now since Q is to R as A is to B, while A and B are least, and the least numbers measure
those which have the same ratio with them the same number of times, the antecedent the VII.20
antecedent and the consequent the consequent, therefore B measures R. For the same
reason C also measures R, therefore B and C measure R.
Therefore the least number measured by B and C also measures R. But G is the least VII.35
number measured by B and C, therefore G measures R.
Therefore the least number measured by E and K also measures S. But M is the least
number measured by E and K, therefore M measures S, the greater the less, which is VII.35
impossible.
Therefore there are no numbers less than N, O, M, and P continuously proportional in the
ratios of A to B, of C to D, and of E to F. Therefore N, O, M, and P are the least numbers
continuously proportional in the ratios A, B, C, D, E, and F.
Q.E.D.
This is a generalization of VIII.2 to a more general concept of continued proportion. In this proposition
we consider continued proportions having ratios that aren't necessarily constant. These, perhaps, should
be called continued ratios. For example, the continued ratio 5:10:20 has a constant ratio of 1:2, but the
continued ratio 5:10:30 does not; its first ratio is 1:2 while its second ratio is 1:3. Note that the continued
ratio 5:10:30 is not the least with those given ratios since 1:2:6 is smaller with the same ratios of 1:2 and
1:3.
The problem here is to construct the smallest continued ratio having the specified ratios.
An examination of the proof shows that Euclid has a general process to attach two continued proportions
into one long one with with the same ratios. Take, for example, the problem of placing the continued
ratio 3:7:2:6 in front of the continued ratio 10:4:5 to make a seven-term continued ratio where the first
four terms have the ratio 3:7:2:6 and the last three terms have the ratio 10:4:5. The resulting seven-term
ratio should be least with the given ratios. The problem is that the last term of the first ratio, 6, does not
equal the first term of the second ratio, 10. The solution is to increase the numbers in each ratio to match
these numbers. Since 30 = LCM(6, 10), that can be done by multiplying each of the terms of the first
ratio by 5 and each of the terms of the second ratio by 3. The resulting ratios, 15:35:10:30 and 30:12:15
can then be merged into the desired ratio 15:35:10:30:12:15.
We start with three given ratios a:b, c:d, and e:f, all in lowest terms. First, the two ratios a:b and c:d are
merged into a three-term ratio h:g:k so that h:g = a:b and g:k = c:d. These are defined equationally as
g = LCM(b, c), h = (g/b)a, and k = (g/c)d. Then h:g:k has the proper ratios.
Next, the three-term ratio h:g:k is merged with the ratio e:f to get a four-term ratio n:o:m:p so that n:
o = a:b, o:m = c:d, and m:p = c:d. These are defined equationally as m = LCM(e, k), n = (m/k)h, and
o = (m/k)g, and p = (m/e)f. Again, n:o:m:p has the proper ratios.
The bulk of the proof consists of showing that the resulting ratio n:o:m:p is the least with the given
ratios.
© 1996, 2002
D.E.Joyce
Clark University
Proposition 5
Plane numbers have to one another the ratio compounded of the ratios of their sides.
Let A and B be plane numbers, and let the numbers C and D be the sides of A, and E and F the
sides of B.
I say that A has to B the ratio compounded of the ratios of the sides.
Multiply D by E to make L.
Again, since E multiplied by D makes L, and further multiplied by F makes B, therefore D is VII.17
to F as L is to B. But D is to F as H is to K, therefore H is to K as L is to B.
But G has to K the ratio compounded of the ratios of the sides, therefore A also has to B the
ratio compounded of the ratios of the sides.
Therefore, plane numbers have to one another the ratio compounded of the ratios of their sides.
Q.E.D.
Compounded ratios
Compound ratios as such only appear in a few places in the Elements. They appear here in this
proposition, and in VI.23, an analogous proposition for rectangles and parallelograms. But duplicate and
triplicate ratios are also special kinds of compound ratios, and they are used in Books VI, VIII, IX, X,
XI, and XII. Duplicate and triplicate ratios were defined in general in V.9-10, where they are defined as
the ratio of the ends of a continued proportion. That is, if a:b = b:c, then the duplicate ratio of a:b is a:c.
For lines, constructing the third proportional c needed to duplicate the ratio a:b is done in proposition
VI.11, but for numbers, the third proportional can be constructed by VIII.2.
The ratio compounded from two given ratios a:b and b:c is just the ratio a:c. But if the middle term b is
not shared by the two given ratios, then equal ratios must be found that do have a shared middle term.
To find the ratio compounded from two given ratios a:b and c:d, first find e, f, and g so that e:f = a:b and
f:g = c:d. Then, the ratio compounded from the ratios a:b and c:d will be the same as the ratio
compounded from the ratios e:f and f:g, namely e:g. For numbers, this construction was done in the
previous proposition VIII.4.
Let the plane number a be the product cd of its sides, and let the plane number b be the product ef of its
sides. Use VIII.4 to construct a continued ratio g:h:k so that g:h = c:e and h:k = d:f so that g:k is the ratio
compounded of the ratios c:e and d:f of the sides.
Since a = cd, therefore c:e = a:de, and so g:h = a:de. Since b = ef, therefore d:f = de:b, and so h:k = de:b.
From the two proportions g:h = a:de and h:k = de:b therefore, ex aequali, g:k = a:b. Thus, ratio the plane
numbers is the ratio compounded of the ratios of their sides.
The application of VIII.4 to find the least numbers continuously in the ratios c:d and e:f actually makes
the proof more difficult. Here's a slightly shorter proof. Since
therefore, the ratio compounded from the ratios c:e and d:f of the sides is the ratio of the plane numbers
a:b.
© 1996, 2002
D.E.Joyce
Clark University
Proposition 6
If there are as many numbers as we please in continued proportion, and the first does not
measure the second, then neither does any other measure any other.
Let there be as many numbers as we please, A, B, C, D, and E, in continued proportion,
and let A not measure B.
I say that neither does any other measure any [later] other.
Now it is manifest that A, B, C, D, and E do not measure one another in order, for A does
not even measure B.
I say, then, that neither does any other measure any [later] other.
And since A is to B as F is to G, while A does not measure B, therefore neither does F VII.Def.20
measure G. Therefore F is not a unit, for the unit measures any number.
Now F and H are relatively prime. And F is to H as A is to C, therefore neither does A VIII.3
measure C.
Similarly we can prove that neither does any other measure any other.
Therefore, if there are as many numbers as we please in continued proportion, and the first does not
measure the second, then neither does any other measure any other.
Q.E.D.
The proposition as stated isn't quite correct. For example, the numbers 24, 12, 6 and 3 are in continued
proportion, and 24 does not divide 12, but each of the others does divide others, for instance, 3 divides 6.
But none of the others divide others later in the sequence.
Consider a sequence of numbers in continued proportion where the first number does not divide the
second. Since any number in that sequence has to the next number in the sequence the same ratio as the
first has to the second, therefore no number divides the next.
Suppose that some number in the sequence divides a later number. We may call that the former number
a since it divides the next number in the sequence, and call the number it divides c. Take the continued
proportion a, b, ..., c and, using VII.33, reduce it a continued proportion f, g, ..., h in lowest terms. Since
that's in lowest terms, f and h are relatively prime. Since a:b = f:g, and a does not divide b, therefore f
does not divide g. Since f does not divide g, in particular f does not equal 1, but f and h are relatively
prime by VIII.3, therefore f does not divide h. Finally, since a:c = f:h, therefore a does not divide c
either.
© 1996, 2002
D.E.Joyce
Clark University
Proposition 7
If there are as many numbers as we please in continued proportion, and the first measures the
last, then it also measures the second.
Let there be as many numbers as we please, A, B, C, and D, in continued proportion, and let A
measure D.
Therefore, if there are as many numbers as we please in continued proportion, and the first measures
the last, then it also measures the second.
Q.E.D.
© 1996
D.E.Joyce
Clark University
Proposition 8
If between two numbers there fall numbers in continued proportion with them, then, however
many numbers fall between them in continued proportion, so many also fall in continued
proportion between the numbers which have the same ratios with the original numbers.
Let the numbers C and D fall between the two numbers A and B in continued
proportion with them, and make E in the same ratio to F as A is to B.
But G and L are relatively prime, numbers which are relatively prime are also VII.21
least, and the least numbers measure those which have the same ratio the same
number of times, the greater the greater and the less the less, that is, the
VII.20
antecedent the antecedent and the consequent the consequent.
Therefore, if between two numbers there fall numbers in continued proportion with them, then,
however many numbers fall between them in continued proportion, so many also fall in continued
proportion between the numbers which have the same ratios with the original numbers.
Q.E.D.
This proposition implies, among other things, that there is no number which forms a mean proportional
between a number n and the number 2n, for if there were, there would be a number m so that 2, m, and 4
would form a continued proportion, but the only number between 2 and 4 is 3, and 2, 3, and 4 do not
form a continued proportion. (If 1 is considered to be a number, the argument simplifies.) In modern
terminology, this conclusion says the square root of 2 is not a rational number. See proposition X.9 for
implications of this conclusion for imcommensurability of line segments.
Suppose that a:b = e:f and the sequence a, c, d, ..., b are in continued proportion. Use VII.33 to reduce
that sequence to lowest terms g, h, k, ..., l. According to VIII.3, the ends of that sequence g and l are
relatively prime. Since a:b = e:f, we also have g:l = e:f. But g is relatively prime to l, so the ratio g:l is in
lowest terms (VII.21), and therefore g divides e the same number of times, say n, that l divides f
(VII.20). Then the sequence ng, nh, nk, ..., nl is in continued proportion and starts with e and ends with f
as required.
Although this proposition is not used in Book VIII, it is used in the first six propositions of Book IX.
© 1996, 2002
D.E.Joyce
Clark University
Proposition 9
If two numbers are relatively prime, and numbers fall between them in continued proportion,
then, however many numbers fall between them in continued proportion, so many also fall
between each of them and a unit in continued proportion.
Let A and B be two numbers relatively prime, and let C and D fall between them in
continued proportion, and let the unit E be set out.
I say that, as many numbers fall between A and B in continued proportion as fall between
either of the numbers A or B and the unit in continued proportion.
Take two numbers F and G, the least that are in the ratio of A, C, D, and B, three numbers
H, K, and L with the same property, and others more by one continually, until their VIII.2
multitude equals the multitude of A, C, D, and B. Let them be M, N, O, and P.
It is now manifest that F multiplied by itself makes H and multiplied by H makes M, while VIII.2,Cor
G multiplied by itself makes L and multiplied by L makes P.
And, since M, N, O, and P are the least of those which have the same ratio with F and G,
and A, C, D, and B are also the least of those which have the same ratio with F and G,
while the multitude of the numbers M, N, O, and P equals the multitude of the numbers A, VIII.1
C, D, and B, therefore M, N, O, and P equal A, C, D, and B respectively. Therefore M
equals A, and P equals B.
Now, since F multiplied by itself makes H, therefore F measures H according to the units
in F. But the unit E also measures F according to the units in it, therefore the unit E
measures the number F the same number of times as F measures H.
But it was also proved that the unit E is to the number F as F is to H, therefore the unit E
is to the number F as F is to H, and as H is to M. But M equals A, therefore the unit E is to
the number F as F is to H, and as H is to A. For the same reason also the unit E is to the
number G as G is to L and as L is to B.
Therefore as many numbers fall between A and B in continued proportion as fall between
each of the numbers A and B and the unit E in continued proportion.
Therefore, If two numbers are relatively prime, and numbers fall between them in continued
proportion, then, however many numbers fall between them in continued proportion, so many also fall
between each of them and a unit in continued proportion.
Q.E.D.
Suppose that relatively prime numbers a and b are the ends of a continued proportion with n terms, and
that f / g is the ratio for the continued proportion in lowest terms. Then Euclid shows that a is the n–1st
power of f, and b is the n–1st power of g. The argument is that the sequence
is in continued proportion with the correct ratio with relatively prime ends, so by VIII.1 they're the same
sequence.
© 1996, 2002
D.E.Joyce
Clark University
Proposition 10
If numbers fall between two numbers and a unit in continued proportion, then however many
numbers fall between each of them and a unit in continued proportion, so many also fall
between the numbers themselves in continued proportion.
Let the numbers D and E and the numbers F and G respectively fall between the two
numbers A and B and the unit C in continued proportion.
I say that, as many numbers have fallen between each of the numbers A and B and the unit
C in continued proportion as fall between A and B in continued proportion.
Now, since the unit C is to the number D as D is to E, therefore the unit C measures the
number D the same number of times as D measures E. But the unit C measures the number VII.Def.20
D according to the units in D, therefore the number D also measures E according to the
units in D. Therefore D multiplied by itself makes E.
Again, since C is to the number D as E is to A, therefore the unit C measures the number D
the same number of times as E measures A. But the unit C measures the number D
according to the units in D, therefore E also measures A according to the units in D.
Therefore D multiplied by E makes A.
For the same reason also F multiplied by itself makes G, and multiplied by G makes B.
And, since D multiplied by itself makes E and multiplied by F makes H, therefore D is to VII.17
F as E is to H.
Again, since D multiplied by the numbers E and H makes A and K respectively, therefore VII.17
E is to H as A is to K. But E is to H as D is to F, therefore D is to F as A is to K.
Again, since the numbers D and F multiplied by H make K and L respectively, therefore D VII.18
is to F as K is to L. But D is to F as A is to K, therefore A is to K as K is to L. Further, since
F multiplied by the numbers H and G makes L and B respectively, therefore H is to G as L
VII.17
is to B.
Therefore, as many numbers as fall between each of the numbers A and B and the unit C in
continued proportion, so many also fall between A and B in continued proportion.
Therefore, if numbers fall between two numbers and a unit in continued proportion, then however
many numbers fall between each of them and a unit in continued proportion, so many also fall
between the numbers themselves in continued proportion.
Q.E.D.
This is a partial converse to the previous proposition; it doesn't require that the generating numbers d
and f be relatively prime in order to conclude that the sequence
is in continued proportion.
© 1996
D.E.Joyce
Clark University
Proposition 11
Between two square numbers there is one mean proportional number, and the square has to
the square the duplicate ratio of that which the side has to the side.
Let A and B be square numbers, and let C be the side of A, and D of B.
I say that between A and B there is one mean proportional number, and A has to B the ratio
duplicate of that which C has to D.
Multiply C by D to make E.
Now, since A is a square and C is its side, therefore C multiplied by itself makes A. For the
same reason also, D multiplied by itself makes B.
I say next that A also has to B the ratio duplicate of that which C has to D.
Since A, E, and B are three numbers in proportion, therefore A has to B the ratio duplicate of V.Def.9
that which A has to E.
But A is to E as C is to D, therefore A has to B the ratio duplicate of that which the side C has
to D.
Therefore, between two square numbers there is one mean proportional number, and the square has
to the square the duplicate ratio of that which the side has to the side.
Q.E.D.
Between c2 and d2 is the mean proportional cd, and the ratio c2:d2 is the duplicate ratio of c:d. The
argument for the latter statement is that c2:d2 is compounded of the two ratios c2:cd and cd:d2, but both
of those are the same ratio as c:d.
© 1996, 2002
D.E.Joyce
Clark University
Proposition 12
Between two cubic numbers there are two mean proportional numbers, and the cube has to the
cube the triplicate ratio of that which the side has to the side.
Let A and B be cubic numbers, and let C be the side of A, and D of B.
I say that between A and B there are two mean proportional numbers, and A has to K the
ratio triplicate of that which C has to D.
Now, since A is a cube, and C its side, and C multiplied by itself makes E, therefore C
multiplied by itself makes E and multiplied by E makes A. For the same reason also D
multiplied by itself makes G and multiplied by G makes B.
And, since C multiplied by the numbers C and D makes E and F respectively, therefore C is VII.17
to D as E is to F. For the same reason also C is to D as F is to G. Again, since C multiplied
by the numbers E and F makes A and H respectively, therefore E is to F as A is to H. But E
VII.18
is to F as C is to D. Therefore C is to D as A is to H.
Again, since the numbers C and D multiplied by F make H and K respectively, therefore C VII.18
is to D as H is to K. Again, since D multiplied by each of the numbers F and G makes K and
B respectively, therefore F is to G as K is to B. VII.17
I say next that A also has to B the ratio triplicate of that which C has to D.
Since A, H, K, and B are four numbers in proportion, therefore A has to B the ratio triplicate V.Def.10
of that which A has to H.
But A is to H as C is to D, therefore A also has to B the ratio triplicate of that which C has to
D.
Therefore, Between two cubic numbers there are two mean proportional numbers, and the cube has to
the cube the triplicate ratio of that which the side has to the side.
Q.E.D.
© 1996
D.E.Joyce
Clark University
Proposition 13
If there are as many numbers as we please in continued proportion, and each multiplied by
itself makes some number, then the products are proportional; and, if the original numbers
multiplied by the products make certain numbers, then the latter are also proportional.
Let there be as many numbers as we please, A, B, and C, in continued proportion, so that A is
to B as B is to C. Let A, B, and C multiplied by themselves make D, E, and F, and multiplied
by D, E, and F let them make G, H, and K.
Then, in manner similar to the foregoing, we can prove that D, L, and E and G, M, N, and H
are continuously proportional in the ratio of A to B, and further E, O, and F and H, P, Q, and K
are continuously proportional in the ratio of B to C.
Now A is to B as B is to C, therefore D, L, and E are also in the same ratio with E, O, and F,
and further G, M, N, and H in the same ratio with H, P, Q, and K. And the multitude of D, L, VII.14
and E equals the multitude of E, O, and F and that of G, M, N, and H to that of H, P, Q, and K,
therefore, ex aequali D is to E as E is to F, and G is to H as H is to K.
Therefore, if there are as many numbers as we please in continued proportion, and each multiplied by
itself makes some number, then the products are proportional; and, if the original numbers multiplied
by the products make certain numbers, then the latter are also proportional.
Q.E.D.
The proposition says that if the terms of a continued proportion are squared or cubed, then the resulting
sequences of numbers are also in continued proportion.
Suppose that the original continued proportion has three terms: a, b, c. Then form two more sequences
Each of these are in continued proportion with the same ratio as the original sequence. The alternate
terms in the second sequence form the continued proportion of the squares of the original sequence
where the ratio is duplicate of the original ratio. Likewise, every third term of the third sequence make
up a continued proportion of the cubes of the original sequence where the ratio is triplicate of the
original ratio.
© 1996, 2002
D.E.Joyce
Clark University
Proposition 14
If a square measures a square, then the side also measures the side; and, if the side measures
the side, then the square also measures the square.
Let A and B be square numbers, let C and D be their sides, and let A measure B.
With the same construction, we can in a similar manner prove that A, E, and B are
continuously proportional in the ratio of C to D. And since C is to D as A is to E, and C VII.Def.20
measures D, therefore A also measures E.
Therefore, if a square measures a square, then the side also measures the side; and, if the side
measures the side, then the square also measures the square.
Q.E.D.
© 1996
D.E.Joyce
Clark University
Proposition 15
If a cubic number measures a cubic number, then the side also measures the side; and, if the
side measures the side, then the cube also measures the cube.
Let the cubic number A measure the cube B, and let C be the side of A and D the side of B.
Now it is manifest that E, F, and G and A, H, K, and B are continuously proportional in the VIII.11
ratio of C to D. And, since A, H, K, and B are continuously proportional, and A measures B VIII.12
and G therefore it also measures H. VIII.7
With the same construction, we can prove in a similar manner that A, H, K, and B are
continuously proportional in the ratio of C to D. And, since C measures D, and C is to D VII.Def.20
as A is to H, therefore A also measures H, so that A measures B also.
Therefore, if a cubic number measures a cubic number, then the side also measures the side; and, if
the side measures the side, then the cube also measures the cube.
Q.E.D.
© 1996
D.E.Joyce
Clark University
Proposition 16
If a square does not measure a square, then neither does the side measure the side; and, if the
side does not measure the side, then neither does the square measure the square.
Let A and B be square numbers, and let C and D be their sides, and let A not measure B.
If A measures B, then C also measures D. But C does not measure D, therefore neither does A VIII.14
measure B.
Therefore, if a square does not measure a square, then neither does the side measure the side; and, if
the side does not measure the side, then neither does the square measure the square.
Q.E.D.
This is simply the contrapositive of VIII.14. It is unclear why papyrus was wasted to state and prove it.
© 1996
D.E.Joyce
Clark University
Proposition 17
If a cubic number does not measure a cubic number, then neither does the side measure the
side; and, if the side does not measure the side, then neither does the cube measure the cube.
Let the cubic number A not measure the cubic number B, and let C be the side of A, and D of
B.
If A measures B, then C also measures D. But C does not measure D, therefore neither does A VIII.15
measure B.
Therefore, if a cubic number does not measure a cubic number, then neither does the side measure the
side; and, if the side does not measure the side, then neither does the cube measure the cube.
Q.E.D.
"Contrariwise," continued Tweedledee, "if it was so, it would be; but as it isn't, it ain't. That's logic."
© 1996
D.E.Joyce
Clark University
Proposition 18
Between two similar plane numbers there is one mean proportional number, and the plane
number has to the plane number the ratio duplicate of that which the corresponding side has to
the corresponding side.
Let A and B be two similar plane numbers, and let the numbers C and D be
the sides of A, and E and F of B.
Now, since similar plane numbers are those which have their sides VII.Def.21
proportional, therefore C is to D as E is to F.
I say then that between A and B there is one mean proportional number,
and A has to B the ratio duplicate of that which C has to E, or as D has to
F, that is, of that which the corresponding side has to the corresponding
side.
I say next that A also has to B the ratio duplicate of that which the
corresponding side has to the corresponding side, that is, of that which C
has to E or D has to F.
Therefore, between two similar plane numbers there is one mean proportional number, and the plane
number has to the plane number the ratio duplicate of that which the corresponding side has to the
corresponding side.
Q.E.D.
Assume that the similar plane numbers are cd and ef so that c:d = e:f, or, alternately, c:e = d:f. Then c:
e = cd:de, and d:f = de:ef, therefore the ratio of the plane numbers cd:ef is compounded of the ratios of
the corresponding sides c:e and d:f. Also, since the ratios of the corresponding sides are the same, the
ratio of the plane numbers is the duplicate of the ratio of the sides. Furthermore, since cd:de = de:ef, the
number de is a mean proportional between the two plane numbers cd and ef.
This proposition is used in several of the remaining propositions in this book beginning with the next. It
is also used in the first two propositions of Book IX. Proposition VIII.20 is a partial converse of this one.
© 1996, 2002
D.E.Joyce
Clark University
Proposition 19
Between two similar solid numbers there fall two mean proportional numbers, and the solid
number has to the solid number the ratio triplicate of that which the corresponding side has to
the corresponding side.
Let A and B be two similar solid numbers, and let C, D, and E be the sides of A, and let F,
G, and H be the sides of B.
Now, since similar solid numbers are those which have their sides proportional, therefore VII.Def.21
C is to D as F is to G, and D is to E as G is to H.
I say that between A and B there fall two mean proportional numbers, and A has to B the
ratio triplicate of that which C has to F, D has to G, and E has to H.
Now, since C and D are in the same ratio with F and G, and K is the product of C and D, VII.
and L the product of F and G, K and L are similar plane numbers, therefore between K and Def.21
L there is one mean proportional number M. VIII.18
Therefore M is the product of D and F was proved in the theorem preceding. VIII.18
Now, since A is a solid number, and C, D, and E are its sides, therefore E multiplied by the
product of C and D makes A. But the product of C and D is K, therefore E multiplied by K
makes A. For the same reason also H multiplied by L makes B.
Now, since E multiplied by K makes A, and further also multiplied by M makes N, VII.17
therefore K is to M as A is to N.
Therefore A, N, O, and B are continuously proportional in the aforesaid ratios of the sides.
I say that A also has to B the ratio triplicate of that which the corresponding side has to the
corresponding side, that is, of the ratio which the number C has to F, or D has to G, and
also E has to H.
Since A, N, O, and B are four numbers in continued proportion, therefore A has to B the
ratio triplicate of that which A has to N. But it was proved that A is to N as C is to F, as D V.Def.10
is to G, and as E is to H.
Therefore A also has to B the ratio triplicate of that which the corresponding side has to the
corresponding side, that is, of the ratio which the number C has to F, D has to G, and also
E has to H.
Therefore, between two similar solid numbers there fall two mean proportional numbers, and the
solid number has to the solid number the ratio triplicate of that which the corresponding side has to
the corresponding side.
Q.E.D.
Assume cde and fgh are similar solid numbers so that c:d:e = f:g:h, or, expressed alternately,
are in continued proportion giving two mean proportionals between cde and fgh. Also, the ratio cde:fgh
is compounded of the three equal ratios of the sides c:f, d:g, and e:h, so it is a triplicate ratio of each.
This proposition is used in a few propositions in Books VIII and IX starting with VIII.25. Proposition
VIII.21 is a partial converse of this one.
© 1996, 2002
D.E.Joyce
Clark University
Proposition 20
If one mean proportional number falls between two numbers, then the numbers are similar
plane numbers.
Let one mean proportional number C fall between the two numbers A and B.
Take D and E, the least numbers of those which have the same ratio with A and C. Then D VII.33
measures A the same number of times that E measures C. VII.20
Let there be as many units in G as times that E measures B. Then E measures B according to
the units in G. Therefore G multiplied by E makes B.
Therefore B is plane, and E and G are its sides. Therefore A and B are plane numbers.
Therefore A and B are similar plane numbers, for their sides are proportional.
Therefore, if one mean proportional number falls between two numbers, then the numbers are similar
plane numbers.
Q.E.D.
This is a partial converse of VIII.18. It says that if two numbers have a mean proportional, then they can
be viewed as two similar plane numbers.
An example might clarify the details. The variables refer to the outline of the proof below. The numbers
a =18 and b =50 have a mean proportional c = 30. When a:c is converted to lowest terms, the result is d:
e = 3:5. Then f, which is a/d, equals 6, and the number a = 18 is seen as the plane number d = 3 by f = 6.
Also g, which is d/c, equals 10, and the number b = 50 is seen as the plane number e = 5 by g = 10. The
sides of these plane numbers, 3 by 6 and 5 by 10, are proportional.
Suppose two numbers a and b have a mean proportional c. Reduce the ratio a:c to lowest terms d:e.
Then d divides a the same number of times e divides c; call that number f. Then a is a plane number with
sides d and f.
Now since, c:b is the same ratio as a:c, it also reduces to the ratio d:e in lowest terms. Therefore, d
divides c the same number of times that e divides b; call that number g. Then b is a plane number with
sides e and g.
Furthermore, the two plane numbers a and b are similar since we can show their sides are proportional
as follows. From the three proportions d:e = a:c (which follows from a = fd and c = fe), a:c = c:b (since c
is a mean proportional), and c:b = f:g (which follows from g = ef and b = ec), therefore, d:e = f:g, and
alternately, d:f = e:g. Thus, the two plane numbers have proportional sides.
This proposition is used in the next two propositions and also IX.2.
© 1996, 2002
D.E.Joyce
Clark University
Proposition 21
If two mean proportional numbers fall between two numbers, then the numbers are similar
solid numbers.
Let two mean proportional numbers C and D fall between the two numbers A and B.
Now, since one mean proportional number F has fallen between E and G, therefore E and G VIII.20
are similar plane numbers.
Therefore it is manifest from the theorem before this that E, F, and G are continuously
proportional in the ratio of H to L, and that of K to M.
Now, since E, F, and G are the least of the numbers which have the same ratio with A, C,
and D, and the multitude of the numbers E, F, and G equals the multitude of the numbers A, VII.14
C, and D, therefore, ex aequali E is to G as A is to D.
But E and G are relatively prime, primes are also least, and the least measure those which VII.21
have the same ratio with them the same number of times, the greater the greater and the less
the less, that is, the antecedent the antecedent and the consequent the consequent, therefore
VII.20
E measures A the same number of times that G measures D.
Let there be as many units in N as times that E measures A. Then N multiplied by E makes
A. But E is the product of H and K, therefore N multiplied by the product of H and K makes
A.
Again, since E, F, and G are the least of the numbers which have the same ratio as C, D,
and B, therefore E measures C the same number of times that G measures B.
Let there be as many units in O as times that E measures C. Then G measures B according
to the units in O, therefore O multiplied by G makes B.
But G is the product of L and M, therefore O multiplied by the product of L and M makes B.
Therefore B is solid, and L, M, and O are its sides. Therefore A and B are solid.
And H, K, and N are the sides of A, and O, L, and M the sides of B. Therefore A and B are
similar solid numbers.
Therefore, if two mean proportional numbers fall between two numbers, then the numbers are similar
solid numbers.
Q.E.D.
This is a partial converse of VIII.19. It says that if two numbers have two mean proportionals, then they
can be viewed as two similar solid numbers. It's proof is analogous the previous proposition dealing with
plane numbers, but naturally, it is longer and more involved.
© 1996
D.E.Joyce
Clark University
Proposition 22
If three numbers are in continued proportion, and the first is square, then the third is also
square.
Let A, B, and C be three numbers in continued proportion, and let A the first be square.
Therefore, if three numbers are in continued proportion, and the first is square, then the third is also
square.
Q.E.D.
This proposition is used in a few propositions in this and the next book starting with VIII.24.
© 1996
D.E.Joyce
Clark University
Proposition 23
If four numbers are in continued proportion, and the first is a cube, then the fourth is also a
cube.
Let A, B, C, and D be four numbers in continued proportion, and let A be a cube.
Therefore, if four numbers are in continued proportion, and the first is a cube, then the fourth is also
a cube.
Q.E.D.
This proposition is used in several propositions in this and the next book starting with VIII.25.
© 1996
D.E.Joyce
Clark University
Proposition 24
If two numbers have to one another the ratio which a square number has to a square number,
and the first is square, then the second is also a square.
Let the two numbers A and B have to one another the ratio which the square number C has to
the square number D, and let A be square.
Since C and D are square, C and D are similar plane numbers. VIII.18
Therefore one mean proportional number falls between C and D.
Therefore, if two numbers have to one another the ratio which a square number has to a square
number, and the first is square, then the second is also a square.
Q.E.D.
© 1996
D.E.Joyce
Clark University
Proposition 25
If two numbers have to one another the ratio which a cubic number has to a cubic number, and
the first is a cube, then the second is also a cube.
Let the two numbers A and B have to one another the ratio which the cubic number C has to
the cubic number D, and let A be a cube.
Since as many numbers fall in continued proportion between those which have the same ratio
with C and D as fall between C and D, therefore two mean proportional numbers E and F fall VIII.18
between A and B.
Since, then, the four numbers A, E, F, and B are in continued proportion, and A is a cube, VIII.23
therefore B is also a cube.
Therefore, if two numbers have to one another the ratio which a cubic number has to a cubic number,
and the first is a cube, then the second is also a cube.
Q.E.D.
This proposition is analogous to the previous one about squares. Its proof is straightforward.
© 1996
D.E.Joyce
Clark University
Proposition 26
Similar plane numbers have to one another the ratio which a square number has to a square
number.
Let A and B be similar plane numbers.
I say that A has to B the ratio which a square number has to a square number.
Take D, E, and F, the least numbers of those which have the same ratio with A, C, and VII.33 or VIII.2
B
Then the extremes of them D and F are square. And since D is to F as A is to B, and D
and F are square, therefore A has to B the ratio which a square number has to a square VIII.2,Cor
number.
Therefore, similar plane numbers have to one another the ratio which a square number has to a
square number.
Q.E.D.
© 1996
D.E.Joyce
Clark University
Proposition 27
Similar solid numbers have to one another the ratio which a cubic number has to a cubic
number.
Let A and B be similar solid numbers.
I say that A has to B the ratio which cubic number has to cubic number.
Since A and B are similar solid numbers, therefore two mean proportional numbers C and VIII.19
D fall between A and B.
Take E, F, G, and H, the least numbers of those which have the same ratio with A, C, D, VII.33 or
and B, and equal with them in multitude. VIII.2
Therefore the extremes of them, E and H, are cubes. And E is to H as A is to B, therefore A VIII.2,
also has to B the ratio which a cubic number has to a cubic number. Cor.
Therefore, similar solid numbers have to one another the ratio which a cubic number has to a cubic
number.
Q.E.D.
This proposition is analogous to the previous proposition about similar plane numbers.
© 1996
D.E.Joyce
Clark University
Table of contents
● Propositions (36)
Propositions
Proposition 1.
If two similar plane numbers multiplied by one another make some number, then the product is
square.
Proposition 2.
If two numbers multiplied by one another make a square number, then they are similar plane
numbers.
Proposition 3.
If a cubic number multiplied by itself makes some number, then the product is a cube.
Proposition 4.
If a cubic number multiplied by a cubic number makes some number, then the product is a cube.
Proposition 5.
If a cubic number multiplied by any number makes a cubic number, then the multiplied number
is also cubic.
Proposition 6.
If a number multiplied by itself makes a cubic number, then it itself is also cubic.
Proposition 7.
If a composite number multiplied by any number makes some number, then the product is solid.
Proposition 8.
If as many numbers as we please beginning from a unit are in continued proportion, then the third
from the unit is square as are also those which successively leave out one, the fourth is cubic as
are also all those which leave out two, and the seventh is at once cubic and square are also those
which leave out five.
Proposition 9.
If as many numbers as we please beginning from a unit are in continued proportion, and the
number after the unit is square, then all the rest are also square; and if the number after the unit is
cubic, then all the rest are also cubic.
Proposition 10.
If as many numbers as we please beginning from a unit are in continued proportion, and the
number after the unit is not square, then neither is any other square except the third from the unit
and all those which leave out one; and, if the number after the unit is not cubic, then neither is
any other cubic except the fourth from the unit and all those which leave out two.
Proposition 11.
If as many numbers as we please beginning from a unit are in continued proportion, then the less
measures the greater according to some one of the numbers which appear among the proportional
numbers.
Corollary. Whatever place the measuring number has, reckoned from the unit, the same place
also has the number according to which it measures, reckoned from the number measured, in the
direction of the number before it.
Proposition 12.
If as many numbers as we please beginning from a unit are in continued proportion, then by
whatever prime numbers the last is measured, the next to the unit is also measured by the same.
Proposition 13.
If as many numbers as we please beginning from a unit are in continued proportion, and the
number after the unit is prime, then the greatest is not measured by any except those which have
a place among the proportional numbers.
Proposition 14.
If a number is the least that is measured by prime numbers, then it is not measured by any other
prime number except those originally measuring it.
Proposition 15.
If three numbers in continued proportion are the least of those which have the same ratio with
them, then the sum of any two is relatively prime to the remaining number.
Proposition 16.
If two numbers are relatively prime, then the second is not to any other number as the first is to
the second.
Proposition 17.
If there are as many numbers as we please in continued proportion, and the extremes of them are
relatively prime, then the last is not to any other number as the first is to the second.
Proposition 18.
Given two numbers, to investigate whether it is possible to find a third proportional to them.
Proposition 19.
Given three numbers, to investigate when it is possible to find a fourth proportional to them.
Proposition 20.
Prime numbers are more than any assigned multitude of prime numbers.
Proposition 21.
If as many even numbers as we please are added together, then the sum is even.
Proposition 22.
If as many odd numbers as we please are added together, and their multitude is even, then the
sum is even.
Proposition 23.
If as many odd numbers as we please are added together, and their multitude is odd, then the sum
is also odd.
Proposition 24.
If an even number is subtracted from an even number, then the remainder is even.
Proposition 25.
If an odd number is subtracted from an even number, then the remainder is odd.
Proposition 26.
If an odd number is subtracted from an odd number, then the remainder is even.
Proposition 27.
If an even number is subtracted from an odd number, then the remainder is odd.
Proposition 28.
If an odd number is multiplied by an even number, then the product is even.
Proposition 29.
If an odd number is multiplied by an odd number, then the product is odd.
Proposition 30.
If an odd number measures an even number, then it also measures half of it.
Proposition 31.
If an odd number is relatively prime to any number, then it is also relatively prime to double it.
Proposition 32.
Each of the numbers which are continually doubled beginning from a dyad is even-times even
only.
Proposition 33.
If a number has its half odd, then it is even-times odd only.
Proposition 34.
If an [even] number neither is one of those which is continually doubled from a dyad, nor has its
half odd, then it is both even-times even and even-times odd.
Proposition 35.
If as many numbers as we please are in continued proportion, and there is subtracted from the
second and the last numbers equal to the first, then the excess of the second is to the first as the
excess of the last is to the sum of all those before it.
Proposition 36.
If as many numbers as we please beginning from a unit are set out continuously in double
proportion until the sum of all becomes prime, and if the sum multiplied into the last makes some
number, then the product is perfect.
Select topic
Elements
Introduction –
© 1996
D.E.Joyce
Clark University
Proposition 1
If two similar plane numbers multiplied by one another make some number, then the product is
square.
Let A and B be two similar plane numbers, and let A multiplied by B make C.
Since as many number fall in continued proportion between those which have the same ratio, VIII.8
therefore one mean proportional number falls between D and C also.
Therefore, if two similar plane numbers multiplied by one another make some number, then the
product is square.
Q.E.D.
Although this is the first proposition in Book IX, it and the succeeding propositions continue those of
Book VIII without break.
To illustrate this proposition, consider the two similar plane numbers a = 18 and b = 8, as illustrated in
the Guide to VII.Def.21. According to VIII.18, there is a mean proportional between them, namely, 12.
And the square of the mean proportional is their product, ab = 144.
It is not clear why the proof of the proposition does not use the fact that the square of the mean
proportional of two numbers equals their product, but instead uses slightly more complicated reasoning.
Let a and b be the given similar plane numbers. Then there is a mean proportional between them
(VIII.18). And, since a:b = a2:ab, therefore there is also a mean proportional between a2 and ab (VIII.1).
But since a2 is a square, therefore ab is also a square (VIII.22). Thus, the product of the original similar
plane numbers is a square.
The next proposition IX.2 is the converse of this one. This proposition is used in X.29.
© 1996, 2002
D.E.Joyce
Clark University
Proposition 2
If two numbers multiplied by one another make a square number, then they are similar plane
numbers.
Let A and B be two numbers, and let A multiplied by B make the square number C.
Therefore one mean proportional number falls between D and C. And D is to C as A is to B, VIII.18
therefore one mean proportional number falls between A and B also. VIII.8
But, if one mean proportional number falls between two numbers, then they are similar plane VIII.20
numbers, therefore A and B are similar plane numbers.
Therefore, if two numbers multiplied by one another make a square number, then they are similar
plane numbers.
Q.E.D.
As an example to illustrate this proposition, take any square number, such as 202 = 400. It can be
factored as a product of two numbers in several ways. One such factorization is as a = 50 times b = 8.
These two numbers have a mean proportional between them, namely, 20, so by VIII.20, they are similar
plane numbers. (The actual shapes given by that proposition make 8 to be 2 by 4, and 50 to be 5 by 10.)
Let a and b be two numbers whose product ab is a square. Now, both a2 and ab are square numbers
which means that they're similar plane numbers. By VIII.8, there's a mean proportional between them.
But a2:ab = a:b, so there is also a mean proportional between a and b (VIII.8). Therefore, a and b are
similar plane figures (VIII.20).
As in the last proof, this one can be shortened. When the product ab is a square, say e2, then a mean
proportional between a and b is e.
Book IX introduction
© 1996, 2002
D.E.Joyce
Clark University
Proposition 3
If a cubic number multiplied by itself makes some number, then the product is a cube.
Let the cubic number A multiplied by itself make B.
Take C, the side of A. Multiply C by itself make D. It is then manifest that C multiplied by
D makes A.
Therefore between the unit and the number A two mean proportional numbers C and D
have fallen in continued proportion.
Again, since A multiplied by itself makes B, therefore A measures B according to the units
in itself. But the unit also measures A according to the units in it, therefore the unit is to A VII.Def.20
as A is to B.
But between the unit and A two mean proportional numbers have fallen, therefore two VIII.8
mean proportional numbers also fall between A and B.
But, if two mean proportional numbers fall between two numbers, and the first is a cube, VIII.23
then the second is also a cube. And A is a cube, therefore B is also a cube.
Therefore, if a cubic number multiplied by itself makes some number, then the product is a cube.
Q.E.D.
Modern algebra certainly makes short work of this proposition: (c3)2 = (c2)3.
Book IX introduction
© 1996
D.E.Joyce
Clark University
Proposition 4
If a cubic number multiplied by a cubic number makes some number, then the product is a cube.
Let the cubic number A multiplied by the cubic number B make C.
Therefore, if a cubic number multiplied by a cubic number makes some number, then the product is
cubic.
Q.E.D.
Book IX introduction
© 1996
D.E.Joyce
Clark University
Proposition 5
If a cubic number multiplied by any number makes a cubic number, then the multiplied number
is also cubic.
Let the cubic number A multiplied by any number B make the cubic number C.
And since D and C are cubes, therefore they are similar solid numbers. Therefore two mean
VIII.19
proportional numbers fall between D and C. And D is to C as A is to B, therefore two mean VIII.8
proportional numbers fall between A and B, too.
Therefore, if a cubic number multiplied by any number makes a cubic number, then the multiplied
number is also cubic.
Q.E.D.
This proposition is a converse of the previous one. When ab = c, and a is a cube, the previous propsition
said that if b is a cube, then c is also, while this proposition says that if c is a cube, then b is also.
Book IX introduction
© 1996
D.E.Joyce
Clark University
Proposition 6
If a number multiplied by itself makes a cubic number, then it itself is also cubic.
Let the number A multiplied by itself make the cubic number B.
But the unit also measures A according to the units in it. Therefore the unit is to A as A is
to B. And, since A multiplied by B makes C, therefore B measures C according to the units VII.Def.20
in A.
And, since B and C are cubes, therefore they are similar solid numbers.
Therefore there are two mean proportional numbers between B and C. And B is to C as A VIII.19
is to B. Therefore there are two mean proportional numbers between A and B also. VIII.8
Therefore, if a number multiplied by itself makes a cubic number, then it itself is also cubic.
Q.E.D.
Assume that a2 is a cube. Since a3 is also a cube, therefore there are two mean proportionals between
them (VIII.19). But we have the proportion a:a2 = a2:a3, so there are also two mean proportionals
between a and a2 (VIII.8). And since a2 is a cube, therefore a is also a cube (VIII.23).
Book IX introduction
© 1996, 2002
D.E.Joyce
Clark University
Proposition 7
If a composite number multiplied by any number makes some number, then the product is solid.
Let the composite number A multiplied by any number B make C.
Therefore, if a composite number multiplied by any number makes some number, then the product is
solid.
Q.E.D.
Numbers with at least two factors are plain numbers; those with at least three are solid numbers.
Perhaps Euclid takes extra steps that we would miss because he sees "d measures a a number e times" as
saying something different from the product of d and e equals a."
Book IX introduction
© 1996
D.E.Joyce
Clark University
Proposition 8
If as many numbers as we please beginning from a unit are in continued proportion, then the
third from the unit is square as are also those which successively leave out one, the fourth is
cubic as are also all those which leave out two, and the seventh is at once cubic and square are
also those which leave out five.
Let there be as many numbers as we please, A, B, C, D, E, and F,
beginning from a unit and in continued proportion.
Similarly we can prove that all those which leave out one are square.
I say next that C, the fourth from the unit, is cubic are also all those which
leave out two.
Therefore, if as many numbers as we please beginning from a unit are in continued proportion, then
the third from the unit is square as are also those which successively leave out one, the fourth is cubic
as are also all those which leave out two, and the seventh is at once cubic and square are also those
which leave out five.
Q.E.D.
every sixth, a6, a12, a18, a24, etc., is both a square and a cube.
Book IX introduction
© 1996
D.E.Joyce
Clark University
Proposition 9
If as many numbers as we please beginning from a unit are in continued proportion, and the
number after the unit is square, then all the rest are also square; and if the number after the
unit is cubic, then all the rest are also cubic.
Now it was proved that B, the third from the unit, is IX.8
square as are all those which leave out one.
Since A, B, and C are in continued proportion, and A is square, therefore C is also square.
Again, since B, C, and D are in continued proportion, and B is square, therefore D is also VIII.22
square. Similarly we can prove that all the rest are also square.
Now it was proved that C, the fourth from the unit, is a cube as are all those which leave out IX.8
two.
Since the unit is to A as A is to B, therefore the unit measures A the same number of times as
A measures B. But the unit measures A according to the units in it, therefore A also measures
B according to the units in itself, therefore A multiplied by itself makes B.
And A is cubic. But, if cubic number multiplied by itself makes some number, then the IX.3
product is also a cube, therefore B is also a cube.
And, since the four numbers A, B, C, and D are in continued proportion, and A is a cube, VIII.23
therefore D also is a cube.
For the same reason E is also a cube, and similarly all the rest are cubes.
Therefore, if as many numbers as we please beginning from a unit are in continued proportion, and
the number after the unit is square, then all the rest are also square; and if the number after the unit
is cubic, then all the rest are also cubic.
Q.E.D.
This proposition says that if a number is a square then all its powers are squares, too. Likewise for
cubes.
Book IX introduction
© 1996
D.E.Joyce
Clark University
Proposition 10
If as many numbers as we please beginning from a unit are in continued proportion, and the
number after the unit is not square, then neither is any other square except the third from the
unit and all those which leave out one; and, if the number after the unit is not cubic, then
neither is any other cubic except the fourth from the unit and all those which leave out two.
Let there be as many numbers as we please, A, B, C, D, E, and F,
beginning from a unit and in continued proportion, and let A, the
number after the unit, not be square.
I say that neither are any other square except the third from the unit
and those which leave out one.
And B is to C as A is to B,
therefore A and B have to one
another the ratio which a square VIII.26 converse
number has to a square number, so
that A and B are similar plane
numbers.
I say that neither are any other cubes except the fourth from the unit
and those which leave out two.
Therefore, if as many numbers as we please beginning from a unit are in continued proportion, and
the number after the unit is not square, then neither is any other square except the third from the unit
and all those which leave out one; and, if the number after the unit is not cubic, then neither is any
other cubic except the fourth from the unit and all those which leave out two.
Q.E.D.
Book IX introduction
© 1996
D.E.Joyce
Clark University
Proposition 11
If as many numbers as we please beginning from a unit are in continued proportion, then the
less measures the greater according to some one of the numbers which appear among the
proportional numbers.
Let there be as many numbers as we please, B, C, D, and E, beginning from the unit A and in
continued proportion.
But the unit A measures D according to the units in it, therefore B also measures E according
to the units in D, so that B the less measures E the greater according to some number of those
which have place among the proportional numbers.
Therefore, if as many numbers as we please beginning from a unit are in continued proportion, then
the less measures the greater according to some one of the numbers which appear among the
proportional numbers.
Q.E.D.
Corollary
And it is manifest that, whatever place the measuring number has, reckoned from the unit, the same
place also has the number according to which it measures, reckoned from the number measured, in the
direction of the number before it.
k n n-k
This proposition, along with the comment made in the corollary, says that a divides a the number a
times.
The corollary is used in the next proposition while the proposition itself is used in the one following that.
Book IX introduction
© 1996
D.E.Joyce
Clark University
Proposition 12
If as many numbers as we please beginning from a unit are in continued proportion, then by
whatever prime numbers the last is measured, the next to the unit is also measured by the same.
Let A, B, C, and D be as many numbers as we please beginning from a unit in
continued proportion.
Now E is prime, and any prime number is relatively prime to any which it
does not measure, therefore E and A are relatively prime. And, since E VII.29
measures D, let it measure it according to F, therefore E multiplied by F
makes D.
But, further, A multiplied by itself makes B, therefore the product of E and H IX.8
equals the square on A. Therefore E is to A as A is to H. VII.19
But A and E are relatively prime, primes are also least, and the least measure
those which have the same ratio the same number of times, the antecedent the
VII.21
antecedent and the consequent the consequent, therefore E measures A VII.20
antecedent antecedent. But, again, it also does not measure it, which is
impossible.
Therefore E and A are not relatively prime. Therefore they are relatively VII.Def.14
composite. But numbers relatively composite are measured by some number.
Therefore, if as many numbers as we please beginning from a unit are in continued proportion, then
by whatever prime numbers the last is measured, the next to the unit is also measured by the same.
Q.E.D.
k
This proposition says that if a prime number p divides a power a of a number a, then it divides the
number a itself.
k
The proof is both elegant and inelegant. The elegant part is the reduction step from p dividing a to p
k
dividing a -1. There are two inelegant parts. One is that the reduction step is applied three times starting
with k equal to 4. The other is that three unnecessary statements are tacked on to the end of the proof
after the goal is already reached.
k
Assume that a prime number p divides a power a of a number a. Suppose that p does not divide a. Then
p is relatively prime to a (VII.29). From the proportion
k k-1
(a /p):a = a:p,
k k-1 k-1
we see that the ratio (a /p):a reduces to a:p in lowest terms (VII.21). Therefore, p divides a
(VII.20).
Apply this reduction step repeatedly until the conclusion p divides a is reached.
Book IX introduction
© 1996, 2002
D.E.Joyce
Clark University
Proposition 13
If as many numbers as we please beginning from a unit are in continued proportion, and the
number after the unit is prime, then the greatest is not measured by any except those which
have a place among the proportional numbers.
Let there be as many numbers as we please, A, B, C, and D, beginning from a unit and in
continued proportion, and let A, the number after the unit, be prime.
I say that D, the greatest of them, is not measured by any other number except A, B, or C.
If possible, let it be measured by E, and let E not be the same with any of the numbers A, B, or
C.
But any composite number is measured by some prime number, therefore E is measured by VII.31
some prime number.
I say next that it is no measured by any other prime except A. If E is measured by another, and
E measures D, then that other measures D, so that it also measures A, which is prime, though IX.12
it is not the same with it, which is impossible. Therefore [only the prime] A measures E.
If F is the same with one of the numbers A, B, or C, and measures D according to E, then one
of the numbers A, B, or C also measures D according to E. But one of the numbers A, B, or C IX.11
measures D according to some one of the numbers A, B, or C, therefore E is also the same
with one of the numbers A, B or C, which is contrary to the hypothesis.
Similarly we can prove that F is measured by A, by proving again that F is not prime.
If it is, and measures D, then it also measures A, which is prime, though it is not the same with IX.12
it, which is impossible. Therefore F is not prime, so it is composite.
But any composite number is measured by some prime number, therefore F is measured by VII.31
some prime number.
If any other prime number measures F, and F measures D, then that other also measures D, so
that it also measures A, which is prime, though it is not the same with it, which is impossible. IX.12
Therefore [only the prime] A measures F.
But, further, A multiplied by C makes D, therefore the product of A and C equals the product IX.11
of E and F.
Similarly, then, we can prove that G is not the same with any of the numbers A or B, and that
it is measured by A. And, since F measures C according to G, therefore F multiplied by G
makes C.
But, further, A multiplied by B makes C, therefore the product of A and B equals the product IX.11
of F and G. Therefore, proportionally A is to F as G is to B. VII.19
And, since G measures B according to H, therefore G multiplied by H makes B. But, further, A IX.8
multiplied by itself makes B, therefore the product of H and G equals the square on A.
Therefore, if as many numbers as we please beginning from a unit are in continued proportion, and
the number after the unit is prime, then the greatest is not measured by any except those which have a
place among the proportional numbers.
Q.E.D.
This proposition says that the only numbers that can divide a power of a prime are smaller powers of
that prime.
The proof involves a reduction step like that in the proof of the previous proposition.
k
Suppose a number e divides a power p of a prime number p, but e does not equal any lower power of p.
First note that e can't be prime itself, since then it would divide p (IX.12), which it doesn't.
k
Then e is composite. Then some prime number q divides e (VII.31). Then q also divides p , which it
implies q divides p. Therefore, the only prime that can divide e is p.
The rest of the proof is repeated reduction of the power k. Since e is not 1, it is divisible by p. Let g be e/
k
p. Then g divides p -1, but is not any lower power of p. Then the same argument can be applied.
Continue in this manner until some number divides p but is not 1 or p, a contradiction. Thus, the only
numbers that can divide a power of a prime are smaller powers of the prime.
Book IX introduction
© 1996, 2002.
D.E.Joyce
Clark University
Proposition 14
If a number is the least that is measured by prime numbers, then it is not measured by any
other prime number except those originally measuring it.
Let the number A be the least that is measured by the prime numbers B, C, and D.
But, if two numbers multiplied by one another make some number, and any prime number
measures the product, then it also measures one of the original numbers, therefore each of B, VII.30
C, and D measures one of the numbers E or F.
Now they do not measure E, for E is prime and not the same with any one of the numbers B,
C, or D. Therefore they measure F, which is less than A, which is impossible, for A is by
hypothesis the least number measured by B, C, and D.
Therefore, if a number is the least that is measured by prime numbers, then it is not measured by any
other prime number except those originally measuring it.
Q.E.D.
This proposition states that the least common multiple of a set of prime numbers is not divisible by any
other prime. The least common multiple is actually the product of those primes, but that isn't mentioned.
The proof is clear, and it depends on VII.30, that if a prime divides a product, then it divides one of the
factors.
Book IX introduction
© 1996
D.E.Joyce
Clark University
Proposition 15
If three numbers in continued proportion are the least of those which have the same ratio with
them, then the sum of any two is relatively prime to the remaining number.
Let A, B, and C, three numbers in continued proportion, be the least of those which have the
same ratio with them.
I say that the sum of any two of the numbers A, B, and C is relatively prime to the remaining
number, that is, A plus B is relatively prime to C, B plus C is relatively prime to A, and A plus
C is relatively prime to B.
It is then manifest that DE multiplied by itself makes A, and multiplied by EF makes B, and Cor. to
that EF multiplied by itself makes C. VIII.2
Now, since DE and EF are least, therefore they are relatively prime. But, if two numbers are
VII.22
relatively prime, then their sum is also relatively prime to each, therefore DF is relatively VII.28
prime to each of the numbers DE and EF.
But, further, DE is also relatively prime to EF, therefore DF and DE are relatively prime to
EF. But, if two numbers are relatively prime to any number, then their product is also VII.24
relatively prime to the other, so that the product of FD and DE is relatively prime to EF, VII.25
hence the product of FD and DE is also relatively prime to the square on EF.
But the product of FD and DE is the square on DE together with the product of DE and EF,
therefore the sum of the square on DE and the product of DE and EF is relatively prime to the II.3
square on EF.
Since DF is relatively prime to each of the numbers DE and EF, therefore the square on DF is VII.24
also relatively prime to the product of DE and EF. VII.25
But the sum of the squares on DE and EF together with twice the product of DE and EF
equals the square on DF, therefore the sum of the squares on DE and EF together with twice II.4
the product of DE and EF is relatively prime to the product of DE and EF.
Taken separately, the sum of the squares on DE and EF together with the product of DE and
EF is relatively prime to the product of DE and EF.
Therefore, taken separately again, the sum of the squares on DE and EF is relatively prime to
the product of DE and EF.
Therefore, if three numbers in continued proportion are the least of those which have the same ratio
with them, then the sum of any two is relatively prime to the remaining number.
Q.E.D.
Let a, b, c be three numbers in continued proportion. Then according to VIII.2, they are of the form
where d and e are relatively prime. Then the sum, d + e, is relatively prime to both d and e (VII.28).
Now, since both d and d + e are relatively prime to e, so is their product d2 + de relatively prime to e
(VII.24), and therefore to e2 (VII.25). Thus, a + b is relatively prime to c.
Next, since d + e is relatively prime to both d and e, so is its square (d + e)2 relatively prime to the
product de (VII.24 and VII.25). That is, d2 + e2 + 2de is relatively prime to de. Subtract 2de to conclude
that d2 + e2 is relatively prime to de. Thus, b is relatively prime to a + c.
Book IX introduction
© 1996, 2002
D.E.Joyce
Clark University
Proposition 16
If two numbers are relatively prime, then the second is not to any other number as the first is to
the second.
Let the two numbers A and B be relatively prime.
If possible as A is to B, let B be to C.
But it also measures itself, therefore A measures A and B which are relatively prime, which is
absurd.
Therefore B is not to C as A is to B.
Therefore, if two numbers are relatively prime, then the second is not to any other number as the first
is to the second.
Q.E.D.
Let a and b be relatively prime. Then the ratio a:b is in lowest terms. Suppose that ratio is the same as
the ratio b:c. Then the antecedent of the ratio a:b, namely a, divides the antecedent of the ratio b:c,
namely b. But a cannot divide b since they're relatively prime.
Book IX introduction
© 1996, 2002
D.E.Joyce
Clark University
Proposition 17
If there are as many numbers as we please in continued proportion, and the extremes of them
are relatively prime, then the last is not to any other number as the first is to the second.
Let there be as many numbers as we please, A, B, C, and D, in continued proportion, and let
the extremes of them, A and D, be relatively prime.
But A and D are prime, primes are also least, and the
least numbers measure those which have the same ratio
the same number of times, the antecedent the antecedent VII.21
and the consequent the consequent. Therefore A VII.20
measures B. And A is to B as B is to C. Therefore B also
measures C, so that A also measures C.
Therefore, if there are as many numbers as we please in continued proportion, and the extremes of
them are relatively prime, then the last is not to any other number as the first is to the second.
Q.E.D.
This proposition generalizes the previous proposition from a ratio of two terms to a continued proportion
of arbitrarily many. It says that a continued proportion in lowest terms cannot be extended.
Consider a continued proportion in lowest terms with the first term a relatively prime to the last term d,
and having the ratio a:b. Suppose it can be extended to e so that a:b = d:e. Alternately, a:d = b:e. Since
the first ratio a:d is in lowest terms, therefore a divides b. Then each term of the continued proportion
divides the next, hence a divides d. But that's impossible since a and d are relatively prime. Therefore,
the continued proportion cannot be extended.
Book IX introduction
© 1996, 2002
D.E.Joyce
Clark University
Proposition 18
Given two numbers, to investigate whether it is possible to find a third proportional to them.
Let A and B be the given two numbers. It is required to investigate whether it is possible to
find a third proportional to them.
Now A and B are either relatively prime or not. And, if they are relatively prime, it was proved IX.16
that it is impossible to find a third proportional to them.
If possible, let D be such third proportional. Then the product of A and D equals the square on
B. But the square on B is C, therefore the product of A and D equals C.
Therefore it is not possible to find a third proportional number to A and B when A does not
measure C.
Q.E.D.
Note that a third proportional d to a and b has to satisfy a:b = b:d, so d would have to be b2/a. So the
third proportional exists when a divides b2. This conclusion is just what Euclid discovers in this
proposition.
Book IX introduction
© 1996
D.E.Joyce
Clark University
Proposition 19
Given three numbers, to investigate when it is possible to find a fourth proportional to them.
Q.E.D.
Note that a fourth proportional d to a, b and c has to satisfy a:b = c:d, so d would have to be bc/a. So the
third proportional exists when a divides bc. No doubt that is what Euclid concludes in the missing part
of this proposition.
Book IX introduction
© 1996, 2002
D.E.Joyce
Clark University
http://aleph0.clarku.edu/~djoyce/java/elements/bookIX/propIX19.html5/26/2008 6:25:18 PM
Euclid's Elements, Book IX, Proposition 20
Proposition 20
Prime numbers are more than any assigned multitude of prime numbers.
Let A, B, and C be the assigned prime numbers.
Next, let EF not be prime. Therefore it is measured by some prime number. Let it be measured VII.31
by the prime number G.
I say that G is not the same with any of the numbers A, B, and C.
Now A, B, and C measure DE, therefore G also measures DE. But it also measures EF.
Therefore G, being a number, measures the remainder, the unit DF, which is absurd.
Therefore G is not the same with any one of the numbers A, B, and C. And by hypothesis it is
prime. Therefore the prime numbers A, B, C, and G have been found which are more than the
assigned multitude of A, B, and C.
Therefore, prime numbers are more than any assigned multitude of prime numbers.
Q.E.D.
The statement says that there are more than any finite number n of prime numbers. Suppose that a1,
a2, ..., an are prime numbers. Euclid, as usual takes an specific small number, = 3, of primes to illustrate the general case. Let
n
m be the least common multiple of all of them. (This least common multiple was also considered in proposition IX.14. It wasn't
noted in the proof of that proposition that the least common multiple of primes is their product, and it isn't noted in this proof,
either.)
Consider the number m + 1. If it's prime, then there are at least n + 1 primes.
So suppose m + 1 is not prime. Then according to VII.31, some prime g divides it. But g cannot be any of the primes a1, a2, ..., a ,
n
since they all divide m and do not divide m + 1. Therefore, there are at least n + 1 primes.
Book IX introduction
D.E.Joyce
Clark University
Proposition 21
If as many even numbers as we please are added together, then the sum is even.
Let as many even numbers as we please, AB, BC,
CD, and DE, be added together.
Since each of the numbers AB, BC, CD, and DE is even, therefore each has a half part, so
that the sum AE also has a half part. But an even number is that which is divisible into two VII.Def.6
equal parts, therefore AE is even.
Therefore, if as many even numbers as we please are added together, then the sum is even.
Q.E.D.
With this proposition, Euclid commences the study of even and odd numbers. The study continues
through proposition IX.34. The statements of these propositions probably constitute the oldest part of the
Elements and date back to the Pythagoreans. Indeed, their proofs depend on no other propositions
(except IX.31 which discusses prime numbers and may have been inserted among these propositions
because it involves odd numbers), so that the statements together with the proofs may be the oldest part
of the Elements.
The proof of this proposition implicitly relies on a principle that the order that numbers are summed is
irrelevant. For example, when showing that the sum of the two even numbers a and b is even, first a is
divided into two equal parts, a = c + c, and b is divided into two equal parts, b = d + d, therefore
a + b = (c + c) + (d + d).
But to reach the conclusion that a + b is divisible into two equal parts, we need
a + b = (c + d) + (c + d),
Of course the order that the terms are added has no effect on the sum. That is an implicit assumption
made by Euclid and most everyone after him until the 19th century.
The modern way to deal with this question is to recognize two properties of addition of numbers. One of
them is commutativity. Addition is commutative since for any two numbers a and b,
a + b = b + a.
Commutativity says we can two numbers in any order and get the same result.
The other property, associativity, is more subtle. When computing the sum a + b + c of three numbers,
there is still a choice of which numbers to add first. You can either add a + b first to get d, then add d + c
to get the sum, or you can add b + c first to get e, then add a + e to get the sum. The same sum should
result. As an equation, associativity says you can move the parentheses around:
(a + b) + c = a + (b + c).
These two properties, commutativity and associativity, are enough to guarantee that when you add any
number of terms together, the order that they're added is irrelvant.
These properties should either be taken as postulates about numbers, or else proven from more basic
assumptions. Besides these properties of addition, Euclid missed some other basic properties of the
arithmetic operations.
Book IX introduction
© 1996, 2002
D.E.Joyce
Clark University
Proposition 22
If as many odd numbers as we please are added together, and their multitude is even, then the
sum is even.
Let as many odd numbers as we please, AB, BC,
CD, and DE, even in multitude, be added together.
Since each of the numbers AB, BC, CD, and DE is odd, if a unit is subtracted from each, (VII.
then each of the remainders is even, so that the sum of them is even. But the multitude of Def.7)
the units is also even. Therefore the sum AE is also even. IX.21
Therefore, if as many odd numbers as we please are added together, and their multitude is even, then
the sum is even.
Q.E.D.
A critical step in the proof is the claim that if 1 is subracted from an odd number, then the remainder is
even. This was mentioned in VII.Def.7, but never proved. See the Guide for that defintion for details.
Unless that gap is filled, this proposition, along with many that depend upon it, are unjustified.
Book IX introduction
© 1996, 2002.
D.E.Joyce
Clark University
Proposition 23
If as many odd numbers as we please are added together, and their multitude is odd, then the
sum is also odd.
Let as many odd numbers as we please, AB, BC, and CD,
the multitude of which is odd, be added together.
Subtract the unit DE from CD, therefore the remainder CE is even. VII.Def.7
IX.22
But CA is also even, therefore the sum AE is also even. IX.21
Therefore, if as many odd numbers as we please are added together, and their multitude is odd, then
the sum is also odd.
Q.E.D.
Book IX introduction
© 1996
D.E.Joyce
Clark University
Proposition 24
If an even number is subtracted from an even number, then the remainder is even.
Let the even number BC be subtracted from the even number AB.
Since AB is even, therefore it has a half part. For the same reason BC also has a half part, so VII.Def.6
that the remainder CA also has a half part, and CA is therefore even.
Therefore, if an even number is subtracted from an even number, then the remainder is even.
Q.E.D.
Book IX introduction
© 1996
D.E.Joyce
Clark University
http://aleph0.clarku.edu/~djoyce/java/elements/bookIX/propIX24.html5/26/2008 6:28:10 PM
Euclid's Elements, Book IX, Proposition 25
Proposition 25
If an odd number is subtracted from an even number, then the remainder is odd.
Let the odd number BC be subtracted from the even
number.
But AB is also even, therefore the remainder AD is also even. And CD is a unit, therefore IX.24
CA is odd. VII.Def.7
Therefore, if an odd number is subtracted from an even number, then the remainder is odd.
Q.E.D.
This is the second of four propositions that examine the result of subtracting even and odd numbers from
each other.
Book IX introduction
© 1996
D.E.Joyce
Clark University
Proposition 26
If an odd number is subtracted from an odd number, then the remainder is even.
Let the odd number BC be subtracted from the odd
number AB.
Therefore, if an odd number is subtracted from an odd number, then the remainder is even.
Q.E.D.
Book IX introduction
© 1996
D.E.Joyce
Clark University
Proposition 27
If an even number is subtracted from an odd number, then the remainder is odd.
Let the even number BC be subtracted from the odd
number AB.
IX.24
But BC is also even, therefore the remainder CD is even. Therefore CA is odd. VII.Def.7
Therefore, if an even number is subtracted from an odd number, then the remainder is odd.
Q.E.D.
This is the last of four propositions that examine the result of subtracting even and odd numbers from
each other.
Book IX introduction
© 1996
D.E.Joyce
Clark University
Proposition 28
But, if as many even numbers as we please be added together, the whole is even. IX.21
Therefore C is even.
Therefore, if an odd number is multiplied by an even number, then the product is even.
Q.E.D.
This is one of two propositions that examine the result of multiplying even and odd numbers by each
other. The third proposition, the product of two even numbers, is omitted.
Note that the proof for this theorem makes no use of the assumption that A is an odd number. The
statement of this theorem might just as well be "if any number is multiplied by an even number, then the
product is even."
The proof of proposition IX.31 concludes at one point that since a number divides a odd number, it must
be even. Such a statement follows from this proposition.
Book IX introduction
© 1996, 2002
D.E.Joyce
Clark University
Proposition 29
Therefore, if an odd number is multiplied by an odd number, then the product is odd.
Q.E.D.
With the completion of this proposition, the study of addition, subtraction, and multiplication of even
and odd numbers is also completed. There remain a few more propositions about even and odd numbers.
Book IX introduction
© 1996
D.E.Joyce
Clark University
Proposition 30
If an odd number measures an even number, then it also measures half of it.
Let the odd number A measure the even number B.
Thus A measures B an even number of times. For this reason then it also measures the half of it.
Therefore, if an odd number measures an even number, then it also measures half of it.
Q.E.D.
Book IX introduction
© 1996
D.E.Joyce
Clark University
Proposition 31
If an odd number is relatively prime to any number, then it is also relatively prime to double it.
Let the odd number A be relatively prime to any number B, and let C be double of B.
Therefore A cannot but be relatively prime to C. Therefore A and C are relatively prime.
Therefore, if an odd number is relatively prime to any number, then it is also relatively prime to
double it.
Q.E.D.
A generalization of this proposition would be "If two numbers (2 and B in this proposition) are relatively
prime to to any number (A), then their product (2B) is also relatively prime to it (A)." That is proposition
VII.24.
Book IX introduction
© 1996
D.E.Joyce
Clark University
Proposition 32
Each of the numbers which are continually doubled beginning from a dyad is even-times even
only.
Let as many numbers as we please, B, C, and D, be continually doubled beginning from the
dyad A.
Set out a unit. Since then as many numbers as we please beginning from a unit are in
continued proportion, and the number A after the unit is prime, therefore D, the greatest of IX.13
the numbers A, B, C, and D, is not measured by any other number except A, B, and C. And VII.Def.8
each of the numbers A, B, and C is even, therefore D is even-times even only.
Similarly we can prove that each of the numbers B and C is even-times even only.
Therefore, each of the numbers which are continually doubled beginning from a dyad is even-times
even only.
Q.E.D.
Numbers which are even-times even only are just the powers of 2: 4, 8, 16, 32, etc.
An alternate proof of this proposition uses IX.30 rather than IX.13. If an odd number could divide D,
then by IX.30, it would divide half of it, namely C. Then it would divide B, then A, the diad, which is
absurd. Note that the reduction step mentioned in this alternate proof is much simpler than the reduction
step used in the proof of IX.30.
Book IX introduction
© 1996, 2002
D.E.Joyce
Clark University
Proposition 33
Now that it is even-times odd is manifest, for the half of it, being odd, measures it an even VII.Def.9
number of times.
Therefore, if a number has its half odd, then it is even-times odd only.
Q.E.D.
To say that a number is even-times odd only means that it is even-times odd, but it is not even-times
even. As this proposition states, such numbers are the numbers which are twice odd numbers.
Book IX introduction
© 1996
D.E.Joyce
Clark University
Proposition 34
If an [even] number neither is one of those which is continually doubled from a dyad, nor has
its half odd, then it is both even-times even and even-times odd.
Let the [even] number A neither be one of those
doubled from a dyad, nor have its half odd.
Now that A is even-times even is manifest, for it has not its half odd. VII.Def.8
If we bisect A, then bisect its half, and do this continually, we shall come
upon some odd number which measures A according to an even number. If
not, we shall come upon a dyad, and A will be among those which are
doubled from a dyad, which is contrary to the hypothesis.
Therefore, if an [even] number neither is one of those which is continually doubled from a dyad, nor
has its half odd, then it is both even-times even and even-times odd.
Q.E.D.
This completes the series of propostions on even and odd numbers that started with IX.21.
Book IX introduction
© 1996
D.E.Joyce
Clark University
Proposition 35
If as many numbers as we please are in continued proportion, and there is subtracted from the
second and the last numbers equal to the first, then the excess of the second is to the first as the
excess of the last is to the sum of all those before it.
Let there be as many numbers as we please in continued proportion, A, BC, D, and EF,
beginning from A as least, and let there be subtracted from BC and EF the numbers BG and
FH, each equal to A.
And since EF is to D as D is to BC, and as BC is to A, while D equals FL, BC equals FK, and
VII.11
A equals FH, therefore EF is to FL as LF is to FK, and as FK is to FH. Taken separately EL is VII.13
to LF as LK is to FK, and as KH is to FH.
Since one of the antecedents is to one of the consequents as the sum of the antecedents is to
the sum of the consequents, therefore KH is to FH as the sum of EL, LK, and KH is to the VII.12
sum of LF, FK, and HF.
But KH equals CG, FH equals A, and the sum of LF, FK, and HF equals the sum of D, BC,
and A, therefore CG is to A as EH is to the sum of D, BC, and A.
Therefore the excess of the second is to the first as the excess of the last is to the sum of those
before it.
Therefore, if as many numbers as we please are in continued proportion, and there is subtracted from
the second and the last numbers equal to the first, then the excess of the second is to the first as the
excess of the last is to the sum of all those before it.
Q.E.D.
This proposition says if a sequence of numbers a1, a2, a3, ..., a , a is in continued proportion
n n+1
then
This conclusion gives a way of computing the sum of the terms in the continued proportion as
a –
n+1
a1 + a2 + ... + a = a1 a1
n .
a2 – a1
If we denote the first term by a and the ratio of the terms by r, then this gives the familiar formula
n
r –
n 1 .
a + ar + ar2 + ... + ar -1 = a
r–1
The proof is much more comprehisible when it's translated in to algebraic notation. The correspondence
is as follows
A = a1 BG = FH = a1
BC = a2 GC = a2 – a1
... EH = a +1 – a1 a2:
n
D=a a1.
n
EF = a
n
+1
a :a = a :a ,
n+1 n n n–1
(a – a ):(a – a ) = a :a ,
n+1 n n n–1 n n–1
then alternately
(a – a ):a = (a – a ):a .
n+1 n n n n–1 n–1
Next, using proposition VII.12, the sum of the antecedents is to the sum of the consequences equals this
same ratio. Therefore
But a –a +a –a + ... + a2 – a1 equals a – a1. That gives us the conclusion of the proposition
n+1 n n n–1 n+1
Book IX introduction
© 1996, 2002
D.E.Joyce
Clark University
Proposition 36
If as many numbers as we please beginning from a unit are set out continuously in double
proportion until the sum of all becomes prime, and if the sum multiplied into the last makes
some number, then the product is perfect.
Let as many numbers as we please, A, B, C, and D, beginning from a unit be set out in
double proportion, until the sum of all becomes prime, let E equal the sum, and let E
multiplied by D make FG.
For, however many A, B, C, and D are in multitude, take so many E, HK, L, and M in
double proportion beginning from E.
But M, L, HK, and E are continuously double of each other, therefore E, HK, L, M, and FG
are continuously proportional in double proportion.
Subtract from the second HK and the last FG the numbers HN and FO, each equal to the
first E. Therefore the excess of the second is to the first as the excess of the last is to the IX.35
sum of those before it. Therefore NK is to E as OG is to the sum of M, L, KH, and E.
And NK equals E, therefore OG also equals M, L, HK, E. But FO also equals E, and E
equals the sum of A, B, C, D and the unit. Therefore the whole FG equals the sum of E,
HK, L, M, A, B, C, D, and the unit, and it is measured by them.
I say also that FG is not measured by any other number except A, B, C, D, E, HK, L, M,
and the unit.
If possible, let some number P measure FG, and let P not be the same with any of the
numbers A, B, C, D, E, HK, L, or M.
And, as many times as P measures FG, so many units let there be in Q, therefore Q
multiplied by P makes FG.
And, since A, B, C, and D are continuously proportional beginning from a unit, therefore D IX.13
is not measured by any other number except A, B, or C.
And, by hypothesis, P is not the same with any of the numbers A, B, or C, therefore P does VII.
not measure D. But P is to D as E is to Q, therefore neither does E measure Q. Def.20
And E is prime, and any prime number is prime to any number which it does not measure. VII.29
Therefore E and Q are relatively prime.
But primes are also least, and the least numbers measure those which have the same ratio
the same number of times, the antecedent the antecedent and the consequent the VII.21
consequent, and E is to Q as P is to D, therefore E measures P the same number of times VII.20
that Q measures D.
But D is not measured by any other number except A, B, or C, therefore Q is the same with
one of the numbers A, B, or C. Let it be the same with B.
And, however many B, C, and D are in multitude, take so many E, HK, and L beginning
from E.
Now E, HK, and L are in the same ratio with B, C, and D, therefore, ex aequali B is to D as VII.14
E is to L.
Therefore the product of B and L equals the product of D and E. But the product of D and
E equals the product of Q and P, therefore the product of Q and P also equals the product VII.19
of B and L.
And FG was proved equal to the sum of A, B, C, D, E, HK, L, M, and the unit, and a VII.
perfect number is that which equals its own parts, therefore FG is perfect. Def.22
Therefore, if as many numbers as we please beginning from a unit are set out continuously in double
proportion until the sum of all becomes prime, and if the sum multiplied into the last makes some
number, then the product is perfect.
Q.E.D.
Euclid begins by assuming that the sum of a number of powers of 2 (the sum beginning with 1) is a
prime number. Let p be the number of powers of 2, and let s be their sum which is prime.
p-1
s = 1 + 2 + 22 + ... + 2
p-1
Note that the last power of 2 is 2 since the sum starts with 1, which is 20.
In Euclid's proof, A represents 2, B represents 22, C represents 23, and D is supposed to be the last power
p-1
of 2, so it represents 2 . Also, E represents their sum s, and FG is the product of E and D, so it
p-1
represents s2 . Let's denote that last by n.
p-1
n = s2
In the first part of this proof, Euclid finds some proper divisors of n that sum to n. These come in two
sequences:
p-1
1, 2, 22, ..., 2
and
n-2
s, 2s, 22s, ..., 2 s
It is clear that each of these is a proper divisor of n, and later in the proof Euclid shows that they are the
only proper divisors of n.
Using the previous proposition, IX.35, Euclid finds the sum of the continued proportion,
n-2
s + 2s + 22s + ... + 2 s,
n-1 p-1
to be 2 s – s. But s was the sum 1 + 2 + 22 + ... + 2 , hence,
n-1 p-1
n=2 s = 1 + 2 + 22 + ... + 2
n-2
+ s + 2s + 22s + ... + 2 s
All that is left to do is to show that they are the only proper divisors of n, for then n will be the sum of all
of its proper divisors, whence a perfect number.
The remainder of the proof is detailed and difficult to follow. It hinges on IX.13 which implies that the
p p
only factors of 2 -1 are powers of 2, so all the factors of 2 -1 have been found. Here's a not-too-faithful
version of Euclid's argument. Suppose n factors as ab where a is not a proper divisor of n in the list
above. In Euclid's proof, P represents a and Q represents b.
p
Since a divides s 2 -1, but is not a power of 2, and s is prime, therefore s divides a. Then b has to be a
power of 2. But then a has to be a power of 2 times s. But all the powers of 2 times s are on the list of
known proper divisors. Therefore, the list includes all the proper divisors.
p p
Note that the sum, s = 1 + 2 + 22 + ... + 2 -1, equals 2 – 1, by IX.35. As this fact is not needed in the
proof, Euclid omits to mention it. Thus, we can restate the proposition as follows:
p p p-1
If 2 – 1 is a prime number, then (2 – 1) 2 is a perfect number.
p
Prime numbers of the form 2 – 1 have come to be called Mersenne primes named in honor of Marin
Mersenne (1588-1648), one of many people who have studied these numbers. The four smallest perfect
numbers, 6, 28, 496, and 8128, were known to the ancient Greek mathematicians. The Mersenne primes
p
2 – 1 corresponding to these four perfect numbers are 3, 7, 31, and 127, respectively, where the
exponents p are 2, 3, 5, and 7, respectively.
The observation that these four exponents are all prime suggests the following two questions:
p
1. In order for 2 – 1 to be prime, is it sufficient for p to be prime?
p
2. In order for 2 – 1 to be prime, is it necessary for p to be prime?
Naturally, the next number to check for primality is 211 – 1, 2047, which, by a simple search for prime
factors is found not to be prime. The number 2047 factors as 23 times 89. Therefore, primality of p is not
sufficient.
In 1640 Pierre de Fermat (1601-1665) wrote to Mersenne with his investigation of these primes. Fermat
p
found three conditions on p that were necessary for 2 – 1 to be prime. One of these conditions answers
the second question above— p does have to be prime. Here's a quick argument for that. If p did factor,
p ab
say as ab, then 2 – 1, which is 2 – 1, would also factor, namely as
ab a a(b-1) a(b-2) a
2 – 1 = (2 – 1) (2 +2 + ... + 2 ).
Many mathematicians have studied Mersenne primes since then. A fairly practical testing algorithm was
p
constructed by Lucas in 1876. He showed that the the number 2 – 1 is prime if and only if it divides the
number S(p-1), where S(p-1) is defined recursively: S(1) = 4, and S(n+1) = S(n)2 – 2.
The search for more Mersenne primes, and therefore more perfect numbers, continues. It is not known if
there are infinitely many or finitely many even perfect numbers. Mersenne primes are scarce, but more
continue to be found. There are at least 39 of them, the largest known (as of December 2005) is
230402457 – 1. It has 9152052 digits. For more information, see the The Great Internet Mersenne Prime
Search, GIMPS.
There is a also a question about odd perfect numbers: Are there any? It has been shown that there are no
small odd perfect numbers; it is known that odd numbers with fewer then 300 digits are not perfect. It
may well be that there are no odd perfect numbers, but to date there is no proof.
Book IX introduction
© 1996, 2005
D.E.Joyce
Clark University
Table of contents
● Definitions I (4)
● Propositions 1-47
● Definitions II (6)
● Propositions 48-84
● Definitions III (6)
● Propositions 85-115
Definitions I
Definition 1.
Those magnitudes are said to be commensurable which are measured by the same measure, and
those incommensurable which cannot have any common measure.
Definition 2.
Straight lines are commensurable in square when the squares on them are measured by the same
area, and incommensurable in square when the squares on them cannot possibly have any area as
a common measure.
Definition 3.
With these hypotheses, it is proved that there exist straight lines infinite in multitude which are
commensurable and incommensurable respectively, some in length only, and others in square
also, with an assigned straight line. Let then the assigned straight line be called rational, and
those straight lines which are commensurable with it, whether in length and in square, or in
square only, rational, but those that are incommensurable with it irrational.
Definition 4.
And the let the square on the assigned straight line be called rational, and those areas which are
commensurable with it rational, but those which are incommensurable with it irrational, and the
straight lines which produce them irrational, that is, in case the areas are squares, the sides
themselves, but in case they are any other rectilineal figures, the straight lines on which are
described squares equal to them.
Propositions 1-47
Proposition 1.
Two unequal magnitudes being set out, if from the greater there is subtracted a magnitude greater
than its half, and from that which is left a magnitude greater than its half, and if this process is
repeated continually, then there will be left some magnitude less than the lesser magnitude set
out. And the theorem can similarly be proven even if the parts subtracted are halves.
Proposition 2.
If, when the less of two unequal magnitudes is continually subtracted in turn from the greater that
which is left never measures the one before it, then the two magnitudes are incommensurable.
Proposition 3.
To find the greatest common measure of two given commensurable magnitudes.
Corollary. If a magnitude measures two magnitudes, then it also measures their greatest common
measure.
Proposition 4.
To find the greatest common measure of three given commensurable magnitudes.
Corollary. If a magnitude measures three magnitudes, then it also measures their greatest
common measure. The greatest common measure can be found similarly for more magnitudes,
and the corollary extended.
Proposition 5.
Commensurable magnitudes have to one another the ratio which a number has to a number.
Proposition 6.
If two magnitudes have to one another the ratio which a number has to a number, then the
magnitudes are commensurable.
Corollary.
Proposition 7.
Incommensurable magnitudes do not have to one another the ratio which a number has to a
number.
Proposition 8.
If two magnitudes do not have to one another the ratio which a number has to a number, then the
magnitudes are incommensurable.
Proposition 9.
The squares on straight lines commensurable in length have to one another the ratio which a
square number has to a square number; and squares which have to one another the ratio which a
square number has to a square number also have their sides commensurable in length. But the
squares on straight lines incommensurable in length do not have to one another the ratio which a
square number has to a square number; and squares which do not have to one another the ratio
which a square number has to a square number also do not have their sides commensurable in
length either.
Corollary. Straight lines commensurable in length are always commensurable in square also, but
those commensurable in square are not always also commensurable in length.
Lemma. Similar plane numbers have to one another the ratio which a square number has to a
square number, and if two numbers have to one another the ratio which a square number has to a
square number, then they are similar plane numbers.
Corollary 2. Numbers which are not similar plane numbers, that is, those which do not have their
sides proportional, do not have to one another the ratio which a square number has to a square
number
Proposition 10.
To find two straight lines incommensurable, the one in length only, and the other in square also,
with an assigned straight line.
Proposition 11.
If four magnitudes are proportional, and the first is commensurable with the second, then the
third also is commensurable with the fourth; but, if the first is incommensurable with the second,
then the third also is incommensurable with the fourth.
Proposition 12.
Magnitudes commensurable with the same magnitude are also commensurable with one another.
Proposition 13.
If two magnitudes are commensurable, and one of them is incommensurable with any magnitude,
then the remaining one is also incommensurable with the same.
Proposition 14.
Lemma. Given two unequal straight lines, to find by what square the square on the greater is
greater than the square on the less. And, given two straight lines, to find the straight line the
square on which equals the sum of the squares on them.
Proposition 14. If four straight lines are proportional, and the square on the first is greater than
the square on the second by the square on a straight line commensurable with the first, then the
square on the third is also greater than the square on the fourth by the square on a third line
commensurable with the third. And, if the square on the first is greater than the square on the
second by the square on a straight line incommensurable with the first, then the square on the
third is also greater than the square on the fourth by the square on a third line incommensurable
with the third.
Proposition 15.
If two commensurable magnitudes are added together, then the whole is also commensurable
with each of them; and, if the whole is commensurable with one of them, then the original
magnitudes are also commensurable.
Proposition 16.
If two incommensurable magnitudes are added together, the sum is also incommensurable with
each of them; but, if the sum is incommensurable with one of them, then the original magnitudes
are also incommensurable.
Proposition 17.
Lemma. If to any straight line there is applied a parallelogram but falling short by a square, then
the applied parallelogram equals the rectangle contained by the segments of the straight line
resulting from the application.
Proposition 17. If there are two unequal straight lines, and to the greater there is applied a
parallelogram equal to the fourth part of the square on the less but falling short by a square, and if
it divides it into parts commensurable in length, then the square on the greater is greater than the
square on the less by the square on a straight line commensurable with the greater. And if the
square on the greater is greater than the square on the less by the square on a straight line
commensurable with the greater, and if there is applied to the greater a parallelogram equal to the
fourth part of the square on the less falling short by a square, then it divides it into parts
commensurable in length.
Proposition 18.
If there are two unequal straight lines, and to the greater there is applied a parallelogram equal to
the fourth part of the square on the less but falling short by a square, and if it divides it into
incommensurable parts, then the square on the greater is greater than the square on the less by the
square on a straight line incommensurable with the greater. And if the square on the greater is
greater than the square on the less by the square on a straight line incommensurable with the
greater, and if there is applied to the greater a parallelogram equal to the fourth part of the square
on the less but falling short by a square, then it divides it into incommensurable parts.
Proposition 19.
Lemma.
Proposition 19. The rectangle contained by rational straight lines commensurable in length is
rational.
Proposition 20.
If a rational area is applied to a rational straight line, then it produces as breadth a straight line
rational and commensurable in length with the straight line to which it is applied.
Proposition 21.
The rectangle contained by rational straight lines commensurable in square only is irrational, and
the side of the square equal to it is irrational. Let the latter be called medial.
Proposition 22.
Lemma. If there are two straight lines, then the first is to the second as the square on the first is
to the rectangle contained by the two straight lines.
Proposition 22. The square on a medial straight line, if applied to a rational straight line,
produces as breadth a straight line rational and incommensurable in length with that to which it is
applied.
Proposition 23.
A straight line commensurable with a medial straight line is medial.
Proposition 24.
The rectangle contained by medial straight lines commensurable in length is medial.
Proposition 25.
The rectangle contained by medial straight lines commensurable in square only is either rational
or medial.
Proposition 26.
A medial area does not exceed a medial area by a rational area.
Proposition 27.
To find medial straight lines commensurable in square only which contain a rational rectangle.
Proposition 28.
To find medial straight lines commensurable in square only which contain a medial rectangle.
Proposition 29.
Lemma 1. To find two square numbers such that their sum is also square.
Lemma 2. To find two square numbers such that their sum is not square.
Proposition 29. To find two rational straight lines commensurable in square only such that the
square on the greater is greater than the square on the less by the square on a straight line
commensurable in length with the greater.
Proposition 30.
To find two rational straight lines commensurable in square only such that the square on the
greater is greater than the square on the less by the square on a straight line incommensurable in
length with the greater.
Proposition 31.
To find two medial straight lines commensurable in square only, containing a rational rectangle,
such that the square on the greater is greater than the square on the less by the square on a straight
line commensurable in length with the greater.
Proposition 32.
To find two medial straight lines commensurable in square only, containing a medial rectangle,
such that the square on the greater is greater than the square on the less by the square on a straight
Proposition 33.
Lemma.
Proposition 33. To find two straight lines incommensurable in square which make the sum of the
squares on them rational but the rectangle contained by them medial.
Proposition 34.
To find two straight lines incommensurable in square which make the sum of the squares on them
medial but the rectangle contained by them rational.
Proposition 35.
To find two straight lines incommensurable in square which make the sum of the squares on them
medial and the rectangle contained by them medial and moreover incommensurable with the sum
of the squares on them.
Proposition 36.
If two rational straight lines commensurable in square only are added together, then the whole is
irrational; let it be called binomial.
Proposition 37.
If two medial straight lines commensurable in square only and containing a rational rectangle are
added together, the whole is irrational; let it be called the first bimedial straight line.
Proposition 38.
If two medial straight lines commensurable in square only and containing a medial rectangle are
added together, then the whole is irrational; let it be called the second bimedial straight line.
Proposition 39.
If two straight lines incommensurable in square which make the sum of the squares on them
rational but the rectangle contained by them medial are added together, then the whole straight
line is irrational; let it be called major.
Proposition 40.
If two straight lines incommensurable in square which make the sum of the squares on them
medial but the rectangle contained by them rational are added together, then the whole straight
line is irrational; let it be called the side of a rational plus a medial area.
Proposition 41.
If two straight lines incommensurable in square which make the sum of the squares on them
medial and the rectangle contained by them medial and also incommensurable with the sum of
the squares on them are added together, then the whole straight line is irrational; let it be called
the side of the sum of two medial areas.
Lemma.
Proposition 42.
A binomial straight line is divided into its terms at one point only.
Proposition 43.
A first bimedial straight line is divided at one and the same point only.
Proposition 44.
A second bimedial straight line is divided at one point only.
Proposition 45.
A major straight line is divided at one point only.
Proposition 46.
The side of a rational plus a medial area is divided at one point only.
Proposition 47.
The side of the sum of two medial areas is divided at one point only.
Definitions II
Definition 1.
Given a rational straight line and a binomial, divided into its terms, such that the square on the
greater term is greater than the square on the lesser by the square on a straight line
commensurable in length with the greater, then, if the greater term is commensurable in length
with the rational straight line set out, let the whole be called a first binomial straight line;
Definition 2.
But if the lesser term is commensurable in length with the rational straight line set out, let the
whole be called a second binomial;
Definition 3.
And if neither of the terms is commensurable in length with the rational straight line set out, let
the whole be called a third binomial.
Definition 4.
Again, if the square on the greater term is greater than the square on the lesser by the square on a
straight line incommensurable in length with the greater, then, if the greater term is
commensurable in length with the rational straight line set out, let the whole be called a fourth
binomial;
Definition 5.
If the lesser, a fifth binomial;
Definition 6.
And, if neither, a sixth binomial.
Propositions 48-84
Proposition 48.
To find the first binomial line.
Proposition 49.
To find the second binomial line.
Proposition 50.
To find the third binomial line.
Proposition 51.
To find the fourth binomial line.
Proposition 52.
To find the fifth binomial line.
Proposition 53.
To find the sixth binomial line.
Proposition 54.
Lemma.
Proposition 54. If an area is contained by a rational straight line and the first binomial, then the
side of the area is the irrational straight line which is called binomial.
Proposition 55.
If an area is contained by a rational straight line and the second binomial, then the side of the area
Proposition 56.
If an area is contained by a rational straight line and the third binomial, then the side of the area is
the irrational straight line called a second bimedial.
Proposition 57.
If an area is contained by a rational straight line and the fourth binomial, then the side of the area
is the irrational straight line called major.
Proposition 58.
If an area is contained by a rational straight line and the fifth binomial, then the side of the area is
the irrational straight line called the side of a rational plus a medial area.
Proposition 59.
If an area is contained by a rational straight line and the sixth binomial, then the side of the area
is the irrational straight line called the side of the sum of two medial areas.
Proposition 60.
Lemma. If a straight line is cut into unequal parts, then the sum of the squares on the unequal
parts is greater than twice the rectangle contained by the unequal parts.
Proposition 60. The square on the binomial straight line applied to a rational straight line
produces as breadth the first binomial.
Proposition 61.
The square on the first bimedial straight line applied to a rational straight line produces as breadth
the second binomial.
Proposition 62.
The square on the second bimedial straight line applied to a rational straight line produces as
breadth the third binomial.
Proposition 63.
The square on the major straight line applied to a rational straight line produces as breadth the
fourth binomial.
Proposition 64.
The square on the side of a rational plus a medial area applied to a rational straight line produces
as breadth the fifth binomial.
Proposition 65.
The square on the side of the sum of two medial areas applied to a rational straight line produces
as breadth the sixth binomial.
Proposition 66.
A straight line commensurable with a binomial straight line is itself also binomial and the same in
order.
Proposition 67.
A straight line commensurable with a bimedial straight line is itself also bimedial and the same in
order.
Proposition 68.
A straight line commensurable with a major straight line is itself also major.
Proposition 69.
A straight line commensurable with the side of a rational plus a medial area is itself also the side
of a rational plus a medial area.
Proposition 70.
A straight line commensurable with the side of the sum of two medial areas is the side of the sum
of two medial areas.
Proposition 71.
If a rational and a medial are added together, then four irrational straight lines arise, namely a
binomial or a first bimedial or a major or a side of a rational plus a medial area.
Proposition 72.
If two medial areas incommensurable with one another are added together, then the remaining
two irrational straight lines arise, namely either a second bimedial or a side of the sum of two
medial areas.
Proposition. The binomial straight line and the irrational straight lines after it are neither the
same with the medial nor with one another.
Proposition 73.
If from a rational straight line there is subtracted a rational straight line commensurable with the
whole in square only, then the remainder is irrational; let it be called an apotome.
Proposition 74.
If from a medial straight line there is subtracted a medial straight line which is commensurable
with the whole in square only, and which contains with the whole a rational rectangle, then the
remainder is irrational; let it be called first apotome of a medial straight line.
Proposition 75.
If from a medial straight line there is subtracted a medial straight line which is commensurable
with the whole in square only, and which contains with the whole a medial rectangle, then the
remainder is irrational; let it be called second apotome of a medial straight line.
Proposition 76.
If from a straight line there is subtracted a straight line which is incommensurable in square with
the whole and which with the whole makes the sum of the squares on them added together
rational, but the rectangle contained by them medial, then the remainder is irrational; let it be
called minor.
Proposition 77.
If from a straight line there is subtracted a straight line which is incommensurable in square with
the whole, and which with the whole makes the sum of the squares on them medial but twice the
rectangle contained by them rational, then the remainder is irrational; let it be called that which
produces with a rational area a medial whole.
Proposition 78.
If from a straight line there is subtracted a straight line which is incommensurable in square with
the whole and which with the whole makes the sum of the squares on them medial, twice the
rectangle contained by them medial, and further the squares on them incommensurable with
twice the rectangle contained by them, then the remainder is irrational; let it be called that which
produces with a medial area a medial whole.
Proposition 79.
To an apotome only one rational straight line can be annexed which is commensurable with the
whole in square only.
Proposition 80.
To a first apotome of a medial straight line only one medial straight line can be annexed which is
commensurable with the whole in square only and which contains with the whole a rational
rectangle.
Proposition 81.
To a second apotome of a medial straight line only one medial straight line can be annexed which
is commensurable with the whole in square only and which contains with the whole a medial
rectangle.
Proposition 82.
To a minor straight line only one straight line can be annexed which is incommensurable in
square with the whole and which makes, with the whole, the sum of squares on them rational but
twice the rectangle contained by them medial.
Proposition 83.
To a straight line which produces with a rational area a medial whole only one straight line can
be annexed which is incommensurable in square with the whole straight line and which with the
whole straight line makes the sum of squares on them medial but twice the rectangle contained by
them rational.
Proposition 84.
To a straight line which produces with a medial area a medial whole only one straight line can be
annexed which is incommensurable in square with the whole straight line and which with the
whole straight line makes the sum of squares on them medial and twice the rectangle contained
by them both medial and also incommensurable with the sum of the squares on them.
Definitions III
Definition 1.
Given a rational straight line and an apotome, if the square on the whole is greater than the square
on the annex by the square on a straight line commensurable in length with the whole, and the
whole is commensurable in length with the rational line set out, let the apotome be called a first
apotome.
Definition 2.
But if the annex is commensurable with the rational straight line set out, and the square on the
whole is greater than that on the annex by the square on a straight line commensurable with the
whole, let the apotome be called a second apotome.
Definition 3.
But if neither is commensurable in length with the rational straight line set out, and the square on
the whole is greater than the square on the annex by the square on a straight line commensurable
with the whole, let the apotome be called a third apotome.
Definition 4.
Again, if the square on the whole is greater than the square on the annex by the square on a
straight line incommensurable with the whole, then, if the whole is commensurable in length with
the rational straight line set out, let the apotome be called a fourth apotome;
Definition 5.
Definition 6.
And, if neither, a sixth.
Propositions 85-115
Proposition 85.
To find the first apotome.
Proposition 86.
To find the second apotome.
Proposition 87.
To find the third apotome.
Proposition 88.
To find the fourth apotome.
Proposition 89.
To find the fifth apotome.
Proposition 90.
To find the sixth apotome.
Proposition 91.
If an area is contained by a rational straight line and a first apotome, then the side of the area is an
apotome.
Proposition 92.
If an area is contained by a rational straight line and a second apotome, then the side of the area is
a first apotome of a medial straight line.
Proposition 93.
If an area is contained by a rational straight line and a third apotome, then the side of the area is a
second apotome of a medial straight line.
Proposition 94.
If an area is contained by a rational straight line and a fourth apotome, then the side of the area is
minor.
Proposition 95.
If an area is contained by a rational straight line and a fifth apotome, then the side of the area is a
straight line which produces with a rational area a medial whole.
Proposition 96.
If an area is contained by a rational straight line and a sixth apotome, then the side of the area is a
straight line which produces with a medial area a medial whole.
Proposition 97.
The square on an apotome of a medial straight line applied to a rational straight line produces as
breadth a first apotome.
Proposition 98.
The square on a first apotome of a medial straight line applied to a rational straight line produces
as breadth a second apotome.
Proposition 99.
The square on a second apotome of a medial straight line applied to a rational straight line
produces as breadth a third apotome.
Proposition 100.
The square on a minor straight line applied to a rational straight line produces as breadth a fourth
apotome.
Proposition 101.
The square on the straight line which produces with a rational area a medial whole, if applied to a
rational straight line, produces as breadth a fifth apotome.
Proposition 102.
The square on the straight line which produces with a medial area a medial whole, if applied to a
rational straight line, produces as breadth a sixth apotome.
Proposition 103.
A straight line commensurable in length with an apotome is an apotome and the same in order.
Proposition 104.
A straight line commensurable with an apotome of a medial straight line is an apotome of a
medial straight line and the same in order.
Proposition 105.
Proposition 106.
A straight line commensurable with that which produces with a rational area a medial whole is a
straight line which produces with a rational area a medial whole.
Proposition 107.
A straight line commensurable with that which produces a medial area and a medial whole is
itself also a straight line which produces with a medial area a medial whole.
Proposition 108.
If from a rational area a medial area is subtracted, the side of the remaining area becomes one of
two irrational straight lines, either an apotome or a minor straight line.
Proposition 109.
If from a medial area a rational area is subtracted, then there arise two other irrational straight
lines, either a first apotome of a medial straight line or a straight line which produces with a
rational area a medial whole.
Proposition 110.
If from a medial area there is subtracted a medial area incommensurable with the whole, then the
two remaining irrational straight lines arise, either a second apotome of a medial straight line or a
straight line which produce with a medial area a medial whole.
Proposition 111.
The apotome is not the same with the binomial straight line.
Proposition. The apotome and the irrational straight lines following it are neither the same with
the medial straight line nor with one another. There are, in order, thirteen irrational straight lines
in all:
Medial
Binomial
First bimedial
Second bimedial
Major
Side of a rational plus a medial area
Side of the sum of two medial areas
Apotome
First apotome of a medial straight line
Second apotome of a medial straight line
Minor
Producing with a rational area a medial whole
Producing with a medial area a medial whole
Proposition 112.
The square on a rational straight line applied to the binomial straight line produces as breadth an
apotome the terms of which are commensurable with the terms of the binomial straight line and
moreover in the same ratio; and further the apotome so arising has the same order as the binomial
straight line.
Proposition 113.
The square on a rational straight line, if applied to an apotome, produces as breadth the binomial
straight line the terms of which are commensurable with the terms of the apotome and in the
same ratio; and further the binomial so arising has the same order as the apotome.
Proposition 114.
If an area is contained by an apotome and the binomial straight line the terms of which are
commensurable with the terms of the apotome and in the same ratio, then the side of the area is
rational.
Proposition 115.
From a medial straight line there arise irrational straight lines infinite in number, and none of
them is the same as any preceding.
© 1996
D.E.Joyce
Clark University
Table of contents
● Definitions (28)
● Propositions (39)
Definitions
Definition 1.
A solid is that which has length, breadth, and depth.
Definition 2.
A face of a solid is a surface.
Definition 3.
A straight line is at right angles to a plane when it makes right angles with all the straight lines
which meet it and are in the plane.
Definition 4.
A plane is at right angles to a plane when the straight lines drawn in one of the planes at right
angles to the intersection of the planes are at right angles to the remaining plane.
Definition 5.
The inclination of a straight line to a plane is, assuming a perpendicular drawn from the end of
the straight line which is elevated above the plane to the plane, and a straight line joined from the
point thus arising to the end of the straight line which is in the plane, the angle contained by the
straight line so drawn and the straight line standing up.
Definition 6.
The inclination of a plane to a plane is the acute angle contained by the straight lines drawn at
right angles to the intersection at the same point, one in each of the planes.
Definition 7.
A plane is said to be similarly inclined to a plane as another is to another when the said angles of
the inclinations equal one another.
Definition 8.
Parallel planes are those which do not meet.
Definition 9.
Similar solid figures are those contained by similar planes equal in multitude.
Definition 10.
Equal and similar solid figures are those contained by similar planes equal in multitude and
magnitude.
Definition 11.
A solid angle is the inclination constituted by more than two lines which meet one another and
are not in the same surface, towards all the lines, that is, a solid angle is that which is contained
by more than two plane angles which are not in the same plane and are constructed to one point.
Definition 12.
A pyramid is a solid figure contained by planes which is constructed from one plane to one point.
Definition 13.
A prism is a solid figure contained by planes two of which, namely those which are opposite, are
equal, similar, and parallel, while the rest are parallelograms.
Definition 14.
When a semicircle with fixed diameter is carried round and restored again to the same position
from which it began to be moved, the figure so comprehended is a sphere.
Definition 15.
The axis of the sphere is the straight line which remains fixed and about which the semicircle is
turned.
Definition 16.
The center of the sphere is the same as that of the semicircle.
Definition 17.
A diameter of the sphere is any straight line drawn through the center and terminated in both
directions by the surface of the sphere.
Definition 18.
When a right triangle with one side of those about the right angle remains fixed is carried round
and restored again to the same position from which it began to be moved, the figure so
comprehended is a cone. And, if the straight line which remains fixed equals the remaining side
about the right angle which is carried round, the cone will be right-angled; if less, obtuse-angled;
and if greater, acute-angled.
Definition 19.
The axis of the cone is the straight line which remains fixed and about which the triangle is
turned.
Definition 20.
And the base is the circle described by the straight in which is carried round.
Definition 21.
When a rectangular parallelogram with one side of those about the right angle remains fixed is
carried round and restored again to the same position from which it began to be moved, the figure
so comprehended is a cylinder.
Definition 22.
The axis of the cylinder is the straight line which remains fixed and about which the
parallelogram is turned.
Definition 23.
And the bases are the circles described by the two sides opposite to one another which are carried
round.
Definition 24.
Similar cones and cylinders are those in which the axes and the diameters of the bases are
proportional.
Definition 25.
A cube is a solid figure contained by six equal squares.
Definition 26.
An octahedron is a solid figure contained by eight equal and equilateral triangles.
Definition 27.
An icosahedron is a solid figure contained by twenty equal and equilateral triangles.
Definition 28.
A dodecahedron is a solid figure contained by twelve equal, equilateral and equiangular
pentagons.
Propositions
Proposition 1.
A part of a straight line cannot be in the plane of reference and a part in plane more elevated.
Proposition 2.
If two straight lines cut one another, then they lie in one plane; and every triangle lies in one
plane.
Proposition 3.
If two planes cut one another, then their intersection is a straight line.
Proposition 4.
If a straight line is set up at right angles to two straight lines which cut one another at their
common point of section, then it is also at right angles to the plane passing through them.
Proposition 5.
If a straight line is set up at right angles to three straight lines which meet one another at their
common point of section, then the three straight lines lie in one plane.
Proposition 6.
If two straight lines are at right angles to the same plane, then the straight lines are parallel.
Proposition 7.
If two straight lines are parallel and points are taken at random on each of them, then the straight
line joining the points is in the same plane with the parallel straight lines.
Proposition 8.
If two straight lines are parallel, and one of them is at right angles to any plane, then the
remaining one is also at right angles to the same plane.
Proposition 9
Straight lines which are parallel to the same straight line but do not lie in the same plane with it
are also parallel to each other.
Proposition 10.
If two straight lines meeting one another are parallel to two straight lines meeting one another not
in the same plane, then they contain equal angles.
Proposition 11.
To draw a straight line perpendicular to a given plane from a given elevated point.
Proposition 12.
To set up a straight line at right angles to a give plane from a given point in it.
Proposition 13.
From the same point two straight lines cannot be set up at right angles to the same plane on the
same side.
Proposition 14.
Planes to which the same straight line is at right angles are parallel.
Proposition 15.
If two straight lines meeting one another are parallel to two straight lines meeting one another not
in the same plane, then the planes through them are parallel.
Proposition 16.
If two parallel planes are cut by any plane, then their intersections are parallel.
Proposition 17.
If two straight lines are cut by parallel planes, then they are cut in the same ratios.
Proposition 18.
If a straight line is at right angles to any plane, then all the planes through it are also at right
angles to the same plane.
Proposition 19.
If two planes which cut one another are at right angles to any plane, then their intersection is also
at right angles to the same plane.
Proposition 20.
If a solid angle is contained by three plane angles, then the sum of any two is greater than the
remaining one.
Proposition 21.
Any solid angle is contained by plane angles whose sum is less than four right angles.
Proposition 22
If there are three plane angles such that the sum of any two is greater than the remaining one, and
they are contained by equal straight lines, then it is possible to construct a triangle out of the
straight lines joining the ends of the equal straight lines.
Proposition 23.
To construct a solid angles out of three plane angles such that the sum of any two is greater than
the remaining one: thus the sum of the three angles must be less than four right angles.
Proposition 24.
If a solid is contained by parallel planes, then the opposite planes in it are equal and
parallelogrammic.
Proposition 25.
If a parallelepipedal solid is cut by a plane parallel to the opposite planes, then the base is to the
base as the solid is to the solid.
Proposition 26.
To construct a solid angle equal to a given solid angle on a given straight line at a given point on
it.
Proposition 27.
To describe a parallelepipedal solid similar and similarly situated to a given parallelepipedal solid
on a given straight line.
Proposition 28.
If a parallelepipedal solid is cut by a plane through the diagonals of the opposite planes, then the
Proposition 29.
Parallelepipedal solids which are on the same base and of the same height, and in which the ends
of their edges which stand up are on the same straight lines, equal one another.
Proposition 30.
Parallelepipedal solids which are on the same base and of the same height, and in which the ends
of their edges which stand up are not on the same straight lines, equal one another.
Proposition 31.
Parallelepipedal solids which are on equal bases and of the same height equal one another.
Proposition 32.
Parallelepipedal solids which are of the same height are to one another as their bases.
Proposition 33.
Similar parallelepipedal solids are to one another in the triplicate ratio of their corresponding
sides.
Corollary. If four straight lines are continuously proportional, then the first is to the fourth as a
parallelepipedal solid on the first is to the similar and similarly situated parallelepipedal solid on
the second, in as much as the first has to the fourth the ratio triplicate of that which it has to the
second.
Proposition 34.
In equal parallelepipedal solids the bases are reciprocally proportional to the heights; and those
parallelepipedal solids in which the bases are reciprocally proportional to the heights are equal.
Proposition 35.
If there are two equal plane angles, and on their vertices there are set up elevated straight lines
containing equal angles with the original straight lines respectively, if on the elevated straight
lines points are taken at random and perpendiculars are drawn from them to the planes in which
the original angles are, and if from the points so arising in the planes straight lines are joined to
the vertices of the original angles, then they contain with the elevated straight lines equal angles.
Proposition 36.
If three straight lines are proportional, then the parallelepipedal solid formed out of the three
equals the parallelepipedal solid on the mean which is equilateral, but equiangular with the
aforesaid solid.
Proposition 37.
If four straight lines are proportional, then parallelepipedal solids on them which are similar and
similarly described are also proportional; and, if the parallelepipedal solids on them which are
similar and similarly described are proportional, then the straight lines themselves are also
proportional.
Proposition 38.
If the sides of the opposite planes of a cube are bisected, and the planes are carried through the
points of section, then the intersection of the planes and the diameter of the cube bisect one
another.
Proposition 39.
If there are two prisms of equal height, and one has a parallelogram as base and the other a
triangle, and if the parallelogram is double the triangle, then the prisms are equal.
© 1996
D.E.Joyce
Clark University
Table of contents
● Propositions (18)
Propositions
Proposition 1.
Similar polygons inscribed in circles are to one another as the squares on their diameters.
Proposition 2.
Circles are to one another as the squares on their diameters.
Proposition 3.
Any pyramid with a triangular base is divided into two pyramids equal and similar to one
another, similar to the whole, and having triangular bases, and into two equal prisms, and the two
prisms are greater than half of the whole pyramid.
Proposition 4.
If there are two pyramids of the same height with triangular bases, and each of them is divided
into two pyramids equal and similar to one another and similar to the whole, and into two equal
prisms, then the base of the one pyramid is to the base of the other pyramid as all the prisms in
the one pyramid are to all the prisms, being equal in multitude, in the other pyramid.
Proposition 5.
Pyramids of the same height with triangular bases are to one another as their bases.
Proposition 6.
Pyramids of the same height with polygonal bases are to one another as their bases.
Proposition 7.
Any prism with a triangular base is divided into three pyramids equal to one another with
triangular bases.
Corollary. Any pyramid is a third part of the prism with the same base and equal height.
Proposition 8.
Similar pyramids with triangular bases are in triplicate ratio of their corresponding sides.
Corollary. Similar pyramids with polygonal bases are also to one another in triplicate ratio of
their corresponding sides.
Proposition 9.
In equal pyramids with triangular bases the bases are reciprocally proportional to the heights; and
those pyramids are equal in which the bases are reciprocally proportional to the heights.
Proposition 10.
Any cone is a third part of the cylinder with the same base and equal height.
Proposition 11.
Cones and cylinders of the same height are to one another as their bases.
Proposition 12.
Similar cones and cylinders are to one another in triplicate ratio of the diameters of their bases.
Proposition 13.
If a cylinder is cut by a plane parallel to its opposite planes, then the cylinder is to the cylinder as
the axis is to the axis.
Proposition 14.
Cones and cylinders on equal bases are to one another as their heights.
Proposition 15.
In equal cones and cylinders the bases are reciprocally proportional to the heights; and those
cones and cylinders in which the bases are reciprocally proportional to the heights are equal.
Proposition 16.
Given two circles about the same center, to inscribe in the greater circle an equilateral polygon
with an even number of sides which does not touch the lesser circle.
Proposition 17.
Given two spheres about the same center, to inscribe in the greater sphere a polyhedral solid
which does not touch the lesser sphere at its surface.
Corollary to XII.17.
Proposition 18.
Spheres are to one another in triplicate ratio of their respective diameters.
Elements
Introduction
© 1996
D.E.Joyce
Clark University
Table of contents
● Propositions (18)
Propositions
Proposition 1.
If a straight line is cut in extreme and mean ratio, then the square on the greater segment added to
the half of the whole is five times the square on the half.
Proposition 2.
If the square on a straight line is five times the square on a segment on it, then, when the double
of the said segment is cut in extreme and mean ratio, the greater segment is the remaining part of
the original straight line.
Proposition 3.
If a straight line is cut in extreme and mean ratio, then the square on the sum of the lesser
segment and the half of the greater segment is five times the square on the half of the greater
segment.
Proposition 4.
If a straight line is cut in extreme and mean ratio, then the sum of the squares on the whole and
on the lesser segment is triple the square on the greater segment.
Proposition 5.
If a straight line is cut in extreme and mean ratio, and a straight line equal to the greater segment
is added to it, then the whole straight line has been cut in extreme and mean ratio, and the
original straight line is the greater segment.
Proposition 6.
If a rational straight line is cut in extreme and mean ratio, then each of the segments is the
irrational straight line called apotome.
Proposition 7.
If three angles of an equilateral pentagon, taken either in order or not in order, are equal, then the
pentagon is equiangular.
Proposition 8.
If in an equilateral and equiangular pentagon straight lines subtend two angles are taken in order,
then they cut one another in extreme and mean ratio, and their greater segments equal the side of
the pentagon.
Proposition 9.
If the side of the hexagon and that of the decagon inscribed in the same circle are added together,
then the whole straight line has been cut in extreme and mean ratio, and its greater segment is the
side of the hexagon.
Proposition 10.
If an equilateral pentagon is inscribed ina circle, then the square on the side of the pentagon
equals the sum of the squares on the sides of the hexagon and the decagon inscribed in the same
circle.
Proposition 11.
If an equilateral pentagon is inscribed in a circle which has its diameter rational, then the side of
the pentagon is the irrational straight line called minor.
Proposition 12.
If an equilateral triangle is inscribed in a circle, then the square on the side of the triangle is triple
the square on the radius of the circle.
Proposition 13.
To construct a pyramid, to comprehend it in a given sphere; and to prove that the square on the
diameter of the sphere is one and a half times the square on the side of the pyramid.
Proposition 14.
To construct an octahedron and comprehend it in a sphere, as in the preceding case; and to prove
that the square on the diameter of the sphere is double the square on the side of the octahedron.
Proposition 15.
To construct a cube and comprehend it in a sphere, like the pyramid; and to prove that the square
on the diameter of the sphere is triple the square on the side of the cube.
Proposition 16.
To construct an icosahedron and comprehend it in a sphere, like the aforesaid figures; and to
prove that the square on the side of the icosahedron is the irrational straight line called minor.
Corollary. The square on the diameter of the sphere is five times the square on the radius of the
circle from which the icosahedron has been described, and the diameter of the sphere is
composed of the side of the hexagon and two of the sides of the decagon inscribed in the same
circle.
Proposition 17.
To construct a dodecahedron and comprehend it in a sphere, like the aforesaid figures; and to
prove that the square on the side of the dodecahedron is the irrational straight line called
apotome.
Corollary. When the side of the cube is cut in extreme and mean ratio, the greater segment is the
side of the dodecahedron.
Proposition 18.
To set out the sides of the five figures and compare them with one another.
Remark.
No other figure, besides the said five figures, can be constructed by equilateral and equiangular
figures equal to one another.
Lemma. The angle of the equilateral and equiangular pentagon is a right angle and a fifth.
© 1996
D.E.Joyce
Clark University
Euclid
From Wikipedia, the free encyclopedia
Euclid
Nationality Greek
Fields Mathematics
Euclid (Greek: Ε•κλε•δης — Eukleid•s), fl. 300 BC, also known as Euclid of Alexandria and
the "Father of Geometry", was a Greek mathematician of the Hellenistic period who was active in
Alexandria, almost certainly during the reign of Ptolemy I (323 BC–283 BC). His Elements is the
most successful textbook in the history of mathematics. In it, the principles of what is now called
Euclidean geometry are deduced from a small set of axioms. Euclid also wrote works on
perspective, conic sections, spherical geometry, and rigor.
Contents
● 1 Biographical knowledge
● 2 The Elements
● 3 Other works
● 4 See also
● 5 References
● 6 External links
Biographical knowledge
Little is known about Euclid other than his writings. What little biographical information we do
have comes largely from commentaries by Proclus and Pappus of Alexandria: Euclid was active at
the great Library of Alexandria and may have studied at Plato's Academy in Greece. The date and
place of Euclid's birth and the date and circumstances of his death are unknown.
Some writers in the Middle Ages confused him with Euclid of Megara, a Greek Socratic
philosopher who lived approximately one century earlier.
The Elements
A fragment of Euclid's Elements found at Oxyrhynchus, which is dated to circa AD 100. The
diagram accompanies Book II, Proposition 5.
Although many of the results in Elements originated with earlier mathematicians, one of Euclid's
accomplishments was to present them in a single, logically coherent framework, making it easy to
use and easy to reference, including a system of rigorous mathematical proofs that remains the
basis of mathematics 23 centuries later.
Although best-known for its geometric results, the Elements also includes number theory. It
considers the connection between perfect numbers and Mersenne primes, the infinitude of prime
numbers, Euclid's lemma on factorization (which leads to the fundamental theorem of arithmetic
on uniqueness of prime factorizations), and the Euclidean algorithm for finding the greatest
common divisor of two numbers.
The geometrical system described in the Elements was long known simply as geometry, and was
considered to be the only geometry possible. Today, however, that system is often referred to as
Euclidean geometry to distinguish it from other so-called Non-Euclidean geometries that
mathematicians discovered in the 19th century.
Other works
Euclid, as imagined by Raphael in this detail from The School of Athens. No likeness or
description of Euclid's physical appearance made during his lifetime survived antiquity. Therefore,
Euclid's depiction in works of art depends on the artist's imagination.
In addition to the Elements, at least five works of Euclid have survived to the present day.
● Data deals with the nature and implications of "given" information in geometrical
problems; the subject matter is closely related to the first four books of the Elements.
● On Divisions of Figures, which survives only partially in Arabic translation, concerns the
division of geometrical figures into two or more equal parts or into parts in given ratios. It
is similar to a third century AD work by Heron of Alexandria.
● Catoptrics, which concerns the mathematical theory of mirrors, particularly the images
formed in plane and spherical concave mirrors. The attribution to Euclid is doubtful. Its
author may have been Theon of Alexandria.
● Phaenomena is a treatise on spherical Astronomy, it survives in Greek and is quite similar
to "On the Moving Sphere", by Autolycus of Pitane, who flourished around 310 BC.
● Optics is the earliest surviving Greek treatise on perspective. In its definitions Euclid
follows the Platonic tradition that vision is caused by discrete rays which emanate from the
eye. One important definition is the fourth: "Things seen under a greater angle appear
greater, and those under a lesser angle less, while those under equal angles appear equal."
In the 36 propositions that follow, Euclid relates the apparent size of an object to its
distance from the eye and investigates the apparent shapes of cylinders and cones when
viewed from different angles. Proposition 45 is interesting, proving that for any two
unequal magnitudes, there is a point from which the two appear equal. Pappus believed
these results to be important in astronomy and included Euclid's Optics, along with his
Phaenomena, in the Little Astronomy, a compendium of smaller works to be studied before
the Syntaxis (Almagest) of Claudius Ptolemy.
All of these works follow the basic logical structure of the Elements, containing definitions and
proved propositions.
There are also works credibly attributed to Euclid which have been lost.
● Conics was a work on conic sections that was later extended by Apollonius of Perga into
his famous work on the subject. It is likely that the first four books of Apollonius's work
come directly from Euclid. According to Pappus, "Apollonius, having completed Euclid's
four books of conics and added four others, handed down eight volumes of conics." The
Conics of Apollonius quickly supplanted the former work, and by the time of Pappus,
Euclid's work was already lost.
● Porisms might have been an outgrowth of Euclid's work with conic sections, but the exact
meaning of the title is controversial.
● Pseudaria, or Book of Fallacies, was an elementary text about errors in reasoning.
● Surface Loci concerned either loci (sets of points) on surfaces or loci which were
themselves surfaces; under the latter interpretation, it has been hypothesized that the work
might have dealt with quadric surfaces.
● Several works on mechanics are attributed to Euclid by Arabic sources. On the Heavy and
the Light contains, in nine definitions and five propositions, Aristotelian notions of moving
bodies and the concept of specific gravity. On the Balance treats the theory of the lever in a
similarly Euclidean manner, containing one definition, two axioms, and four propositions.
A third fragment, on the circles described by the ends of a moving lever, contains four
propositions. These three works complement each other in such a way that it has been
suggested that they are remnants of a single treatise on mechanics written by Euclid.
See also
● Axiomatic method
● Euclidean algorithm
● Euclidean geometry
References
● "Euclid (Greek mathematician)". Encyclopædia Britannica Online. (2008). Chicago:
Encyclopædia Britannica, Inc. Retrieved on 2008-04-18.
● Artmann, Benno (1999). Euclid: The Creation of Mathematics. New York: Springer. ISBN
0-387-98423-2.
● Ball, W.W. Rouse (1960). A Short Account of the History of Mathematics, 4th ed. [Reprint.
Original publication: London: Macmillan & Co., 1908], New York: Dover Publications,
50–62. ISBN 0-486-20630-0. “Of his life we know next to nothing, save that he was of
Greek descent ...” [p. 52].
● Boyer, Carl B. (1991). A History of Mathematics, 2d ed., John Wiley & Sons, Inc.. ISBN 0-
47154397-7.
● Heath, Thomas L. (1956). The Thirteen Books of Euclid's Elements, Vol. 1 (2nd ed.). New
York: Dover Publications. ISBN 0-486-60088-2: includes extensive commentaries on
Euclid and his work in the context of the history of mathematics that preceded him.
● Heath, Thomas L. (1981). A History of Greek Mathematics, 2 Vols. New York: Dover
Publications. ISBN 0-486-24073-8 / ISBN 0-486-24074-6.
● Kline, Morris (1980). Mathematics: The Loss of Certainty. Oxford: Oxford University
Press. ISBN 0-19-502754-X.
● O'Connor, John J. & Robertson, Edmund F., "Euclid", MacTutor History of Mathematics
archive.
External links
[hide]
Greek mathematics
Anaxagoras · Anthemius · Aratus · Archytas · Aristaeus the Elder ·
Aristarchus · Apollonius · Archimedes · Autolycus · Bion · Boethius ·
Bryson · Callippus · Carpus · Chrysippus · Cleomedes · Conon · Ctesibius ·
Democritus · Dicaearchus · Diocles · Diophantus · Dinostratus ·
Dionysodorus · Domninus · Eratosthenes · Eudemus · Euclid · Eudoxus ·
Eutocius · Geminus · Heron · Hipparchus · Hippasus · Hippias · Hippocrates ·
Mathematicians
Hypatia · Hypsicles · Marinus · Menaechmus · Menelaus · Nicomachus ·
Nicomedes · Nicoteles · Oenopides · Pappus · Perseus · Philolaus · Philon ·
Porphyry · Posidonius · Proclus · Ptolemy · Pythagoras · Pytheas · Serenus ·
Simplicius · Sosigenes · Sporus · Thales · Theaetetus · Theano · Theodorus ·
Theodosius · Theon of Alexandria · Theon of Smyrna · Thymaridas ·