fbpx
Wikipedia

Pythagorean theorem

In mathematics, the Pythagorean theorem or Pythagoras' theorem is a fundamental relation in Euclidean geometry between the three sides of a right triangle. It states that the area of the square whose side is the hypotenuse (the side opposite the right angle) is equal to the sum of the areas of the squares on the other two sides. This theorem can be written as an equation relating the lengths of the sides a, b and the hypotenuse c, often called the Pythagorean equation:[1]

Pythagorean theorem
TypeTheorem
FieldEuclidean geometry
StatementThe sum of the areas of the two squares on the legs (a and b) equals the area of the square on the hypotenuse (c).
Symbolic statement
Generalizations
Consequences

The theorem is named for the Greek philosopher Pythagoras, born around 570 BC. The theorem has been proved numerous times by many different methods – possibly the most for any mathematical theorem. The proofs are diverse, including both geometric proofs and algebraic proofs, with some dating back thousands of years.

When Euclidean space is represented by a Cartesian coordinate system in analytic geometry, Euclidean distance satisfies the Pythagorean relation: the squared distance between two points equals the sum of squares of the difference in each coordinate between the points.

The theorem can be generalized in various ways: to higher-dimensional spaces, to spaces that are not Euclidean, to objects that are not right triangles, and to objects that are not triangles at all but n-dimensional solids. The Pythagorean theorem has attracted interest outside mathematics as a symbol of mathematical abstruseness, mystique, or intellectual power; popular references in literature, plays, musicals, songs, stamps, and cartoons abound.

Proofs using constructed squares

 
Rearrangement proof of the Pythagorean theorem.
(The area of the white space remains constant throughout the translation rearrangement of the triangles. At all moments in time, the area is always . And likewise, at all moments in time, the area is always a²+b².)

Rearrangement proofs

In one rearrangement proof, two squares are used whose sides have a measure of   and which contain four right triangles whose sides are a, b and c, with the hypotenuse being c. In the square on the right side, the triangles are placed such that the corners of the square correspond to the corners of the right angle in the triangles, forming a square in the center whose sides are length c. Each outer square has an area of   as well as  , with   representing the total area of the four triangles. Within the big square on the left side, the four triangles are moved to form two similar rectangles with sides of length a and b. These rectangles in their new position have now delineated two new squares, one having side length a is formed in the bottom-left corner, and another square of side length b formed in the top-right corner. In this new position, this left side now has a square of area   as well as  . Since both squares have the area of   it follows that the other measure of the square area also equal each other such that   =  . With the area of the four triangles removed from both side of the equation what remains is   [2]

In another proof rectangles in the second box can also be placed such that both have one corner that correspond to consecutive corners of the square. In this way they also form two boxes, this time in consecutive corners, with areas   and  which will again lead to a second square of with the area  .

English mathematician Sir Thomas Heath gives this proof in his commentary on Proposition I.47 in Euclid's Elements, and mentions the proposals of German mathematicians Carl Anton Bretschneider and Hermann Hankel that Pythagoras may have known this proof. Heath himself favors a different proposal for a Pythagorean proof, but acknowledges from the outset of his discussion "that the Greek literature which we possess belonging to the first five centuries after Pythagoras contains no statement specifying this or any other particular great geometric discovery to him."[3] Recent scholarship has cast increasing doubt on any sort of role for Pythagoras as a creator of mathematics, although debate about this continues.[4]

Algebraic proofs

 
Diagram of the two algebraic proofs

The theorem can be proved algebraically using four copies of the same triangle arranged symmetrically around a square with side c, as shown in the lower part of the diagram.[5] This results in a larger square, with side a + b and area (a + b)2. The four triangles and the square side c must have the same area as the larger square,

 

giving

 

A similar proof uses four copies of a right triangle with sides a, b and c, arranged inside a square with side c as in the top half of the diagram.[6] The triangles are similar with area  , while the small square has side ba and area (ba)2. The area of the large square is therefore

 

But this is a square with side c and area c2, so

 

Other proofs of the theorem

This theorem may have more known proofs than any other (the law of quadratic reciprocity being another contender for that distinction); the book The Pythagorean Proposition contains 370 proofs.[7]

Proof using similar triangles

 
Proof using similar triangles

This proof is based on the proportionality of the sides of three similar triangles, that is, upon the fact that the ratio of any two corresponding sides of similar triangles is the same regardless of the size of the triangles.

Let ABC represent a right triangle, with the right angle located at C, as shown on the figure. Draw the altitude from point C, and call H its intersection with the side AB. Point H divides the length of the hypotenuse c into parts d and e. The new triangle, ACH, is similar to triangle ABC, because they both have a right angle (by definition of the altitude), and they share the angle at A, meaning that the third angle will be the same in both triangles as well, marked as θ in the figure. By a similar reasoning, the triangle CBH is also similar to ABC. The proof of similarity of the triangles requires the triangle postulate: The sum of the angles in a triangle is two right angles, and is equivalent to the parallel postulate. Similarity of the triangles leads to the equality of ratios of corresponding sides:

 

The first result equates the cosines of the angles θ, whereas the second result equates their sines.

These ratios can be written as

 

Summing these two equalities results in

 

which, after simplification, demonstrates the Pythagorean theorem:

 

The role of this proof in history is the subject of much speculation. The underlying question is why Euclid did not use this proof, but invented another. One conjecture is that the proof by similar triangles involved a theory of proportions, a topic not discussed until later in the Elements, and that the theory of proportions needed further development at that time.[8]

Einstein's proof by dissection without rearrangement

Albert Einstein gave a proof by dissection in which the pieces do not need to be moved.[9] Instead of using a square on the hypotenuse and two squares on the legs, one can use any other shape that includes the hypotenuse, and two similar shapes that each include one of two legs instead of the hypotenuse (see Similar figures on the three sides). In Einstein's proof, the shape that includes the hypotenuse is the right triangle itself. The dissection consists of dropping a perpendicular from the vertex of the right angle of the triangle to the hypotenuse, thus splitting the whole triangle into two parts. Those two parts have the same shape as the original right triangle, and have the legs of the original triangle as their hypotenuses, and the sum of their areas is that of the original triangle. Because the ratio of the area of a right triangle to the square of its hypotenuse is the same for similar triangles, the relationship between the areas of the three triangles holds for the squares of the sides of the large triangle as well.

Trigonometric proof using Einstein's construction

 
Right triangle on the hypotenuse dissected into two similar right triangles on the legs, according to Einstein's proof.

Both the proof using similar triangles and Einstein's proof rely on constructing the height to the hypotenuse of the right triangle  . The same construction provides a trigonometric proof of the Pythagorean theorem using the definition of the sine as a ratio inside a right triangle:

 
 
 

and thus

 

This proof is essentially the same as the above proof using similar triangles, where some ratios of lengths are replaced by sines.

Euclid's proof

 
Proof in Euclid's Elements

In outline, here is how the proof in Euclid's Elements proceeds. The large square is divided into a left and right rectangle. A triangle is constructed that has half the area of the left rectangle. Then another triangle is constructed that has half the area of the square on the left-most side. These two triangles are shown to be congruent, proving this square has the same area as the left rectangle. This argument is followed by a similar version for the right rectangle and the remaining square. Putting the two rectangles together to reform the square on the hypotenuse, its area is the same as the sum of the area of the other two squares. The details follow.

Let A, B, C be the vertices of a right triangle, with a right angle at A. Drop a perpendicular from A to the side opposite the hypotenuse in the square on the hypotenuse. That line divides the square on the hypotenuse into two rectangles, each having the same area as one of the two squares on the legs.

For the formal proof, we require four elementary lemmata:

  1. If two triangles have two sides of the one equal to two sides of the other, each to each, and the angles included by those sides equal, then the triangles are congruent (side-angle-side).
  2. The area of a triangle is half the area of any parallelogram on the same base and having the same altitude.
  3. The area of a rectangle is equal to the product of two adjacent sides.
  4. The area of a square is equal to the product of two of its sides (follows from 3).

Next, each top square is related to a triangle congruent with another triangle related in turn to one of two rectangles making up the lower square.[10]

 
Illustration including the new lines
 
Showing the two congruent triangles of half the area of rectangle BDLK and square BAGF

The proof is as follows:

  1. Let ACB be a right-angled triangle with right angle CAB.
  2. On each of the sides BC, AB, and CA, squares are drawn, CBDE, BAGF, and ACIH, in that order. The construction of squares requires the immediately preceding theorems in Euclid, and depends upon the parallel postulate.[11]
  3. From A, draw a line parallel to BD and CE. It will perpendicularly intersect BC and DE at K and L, respectively.
  4. Join CF and AD, to form the triangles BCF and BDA.
  5. Angles CAB and BAG are both right angles; therefore C, A, and G are collinear.
  6. Angles CBD and FBA are both right angles; therefore angle ABD equals angle FBC, since both are the sum of a right angle and angle ABC.
  7. Since AB is equal to FB, BD is equal to BC and angle ABD equals angle FBC, triangle ABD must be congruent to triangle FBC.
  8. Since A-K-L is a straight line, parallel to BD, then rectangle BDLK has twice the area of triangle ABD because they share the base BD and have the same altitude BK, i.e., a line normal to their common base, connecting the parallel lines BD and AL. (lemma 2)
  9. Since C is collinear with A and G, and this line is parallel to FB, then square BAGF must be twice in area to triangle FBC.
  10. Therefore, rectangle BDLK must have the same area as square BAGF = AB2.
  11. By applying steps 3 to 10 to the other side of the figure, it can be similarly shown that rectangle CKLE must have the same area as square ACIH = AC2.
  12. Adding these two results, AB2 + AC2 = BD × BK + KL × KC
  13. Since BD = KL, BD × BK + KL × KC = BD(BK + KC) = BD × BC
  14. Therefore, AB2 + AC2 = BC2, since CBDE is a square.

This proof, which appears in Euclid's Elements as that of Proposition 47 in Book 1, demonstrates that the area of the square on the hypotenuse is the sum of the areas of the other two squares.[12][13] This is quite distinct from the proof by similarity of triangles, which is conjectured to be the proof that Pythagoras used.[14][15]

Proofs by dissection and rearrangement

Another by rearrangement is given by the middle animation. A large square is formed with area c2, from four identical right triangles with sides a, b and c, fitted around a small central square. Then two rectangles are formed with sides a and b by moving the triangles. Combining the smaller square with these rectangles produces two squares of areas a2 and b2, which must have the same area as the initial large square.[16]

The third, rightmost image also gives a proof. The upper two squares are divided as shown by the blue and green shading, into pieces that when rearranged can be made to fit in the lower square on the hypotenuse – or conversely the large square can be divided as shown into pieces that fill the other two. This way of cutting one figure into pieces and rearranging them to get another figure is called dissection. This shows the area of the large square equals that of the two smaller ones.[17]

 
Animation showing proof by rearrangement of four identical right triangles
 
Animation showing another proof by rearrangement
 
Proof using an elaborate rearrangement

Proof by area-preserving shearing

 
Visual proof of the Pythagorean theorem by area-preserving shearing

As shown in the accompanying animation, area-preserving shear mappings and translations can transform the squares on the sides adjacent to the right-angle onto the square on the hypotenuse, together covering it exactly.[18] Each shear leaves the base and height unchanged, thus leaving the area unchanged too. The translations also leave the area unchanged, as they do not alter the shapes at all. Each square is first sheared into a parallelogram, and then into a rectangle which can be translated onto one section of the square on the hypotenuse.

Other algebraic proofs

A related proof was published by future U.S. President James A. Garfield (then a U.S. Representative) (see diagram).[19][20][21] Instead of a square it uses a trapezoid, which can be constructed from the square in the second of the above proofs by bisecting along a diagonal of the inner square, to give the trapezoid as shown in the diagram. The area of the trapezoid can be calculated to be half the area of the square, that is

 

The inner square is similarly halved, and there are only two triangles so the proof proceeds as above except for a factor of  , which is removed by multiplying by two to give the result.

Proof using differentials

One can arrive at the Pythagorean theorem by studying how changes in a side produce a change in the hypotenuse and employing calculus.[22][23][24]

The triangle ABC is a right triangle, as shown in the upper part of the diagram, with BC the hypotenuse. At the same time the triangle lengths are measured as shown, with the hypotenuse of length y, the side AC of length x and the side AB of length a, as seen in the lower diagram part.

 
Diagram for differential proof

If x is increased by a small amount dx by extending the side AC slightly to D, then y also increases by dy. These form two sides of a triangle, CDE, which (with E chosen so CE is perpendicular to the hypotenuse) is a right triangle approximately similar to ABC. Therefore, the ratios of their sides must be the same, that is:

 

This can be rewritten as   , which is a differential equation that can be solved by direct integration:

 

giving

 

The constant can be deduced from x = 0, y = a to give the equation

 

This is more of an intuitive proof than a formal one: it can be made more rigorous if proper limits are used in place of dx and dy.

Converse

The converse of the theorem is also true:[25]

Given a triangle with sides of length a, b, and c, if a2 + b2 = c2, then the angle between sides a and b is a right angle.

For any three positive real numbers a, b, and c such that a2 + b2 = c2, there exists a triangle with sides a, b and c as a consequence of the converse of the triangle inequality.

This converse appears in Euclid's Elements (Book I, 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."[26]

It can be proved using the law of cosines or as follows:

Let ABC be a triangle with side lengths a, b, and c, with a2 + b2 = c2. Construct a second triangle with sides of length a and b containing a right angle. By the Pythagorean theorem, it follows that the hypotenuse of this triangle has length c = a2 + b2, the same as the hypotenuse of the first triangle. Since both triangles' sides are the same lengths a, b and c, the triangles are congruent and must have the same angles. Therefore, the angle between the side of lengths a and b in the original triangle is a right angle.

The above proof of the converse makes use of the Pythagorean theorem itself. The converse can also be proved without assuming the Pythagorean theorem.[27][28]

A corollary of the Pythagorean theorem's converse is a simple means of determining whether a triangle is right, obtuse, or acute, as follows. Let c be chosen to be the longest of the three sides and a + b > c (otherwise there is no triangle according to the triangle inequality). The following statements apply:[29]

Edsger W. Dijkstra has stated this proposition about acute, right, and obtuse triangles in this language:

sgn(α + βγ) = sgn(a2 + b2c2),

where α is the angle opposite to side a, β is the angle opposite to side b, γ is the angle opposite to side c, and sgn is the sign function.[30]

Consequences and uses of the theorem

Pythagorean triples

A Pythagorean triple has three positive integers a, b, and c, such that a2 + b2 = c2. In other words, a Pythagorean triple represents the lengths of the sides of a right triangle where all three sides have integer lengths.[1] Such a triple is commonly written (a, b, c). Some well-known examples are (3, 4, 5) and (5, 12, 13).

A primitive Pythagorean triple is one in which a, b and c are coprime (the greatest common divisor of a, b and c is 1).

The following is a list of primitive Pythagorean triples with values less than 100:

(3, 4, 5), (5, 12, 13), (7, 24, 25), (8, 15, 17), (9, 40, 41), (11, 60, 61), (12, 35, 37), (13, 84, 85), (16, 63, 65), (20, 21, 29), (28, 45, 53), (33, 56, 65), (36, 77, 85), (39, 80, 89), (48, 55, 73), (65, 72, 97)

Inverse Pythagorean theorem

Given a right triangle with sides   and altitude   (a line from the right angle and perpendicular to the hypotenuse  ). The Pythagorean theorem has,

 

while the inverse Pythagorean theorem relates the two legs   to the altitude  ,[31]

 

The equation can be transformed to,

 

where   for any non-zero real  . If the   are to be integers, the smallest solution   is then

 

using the smallest Pythagorean triple  . The reciprocal Pythagorean theorem is a special case of the optic equation

 

where the denominators are squares and also for a heptagonal triangle whose sides   are square numbers.

Incommensurable lengths

 
The spiral of Theodorus: A construction for line segments with lengths whose ratios are the square root of a positive integer

One of the consequences of the Pythagorean theorem is that line segments whose lengths are incommensurable (so the ratio of which is not a rational number) can be constructed using a straightedge and compass. Pythagoras' theorem enables construction of incommensurable lengths because the hypotenuse of a triangle is related to the sides by the square root operation.

The figure on the right shows how to construct line segments whose lengths are in the ratio of the square root of any positive integer.[32] Each triangle has a side (labeled "1") that is the chosen unit for measurement. In each right triangle, Pythagoras' theorem establishes the length of the hypotenuse in terms of this unit. If a hypotenuse is related to the unit by the square root of a positive integer that is not a perfect square, it is a realization of a length incommensurable with the unit, such as 2, 3, 5 . For more detail, see Quadratic irrational.

Incommensurable lengths conflicted with the Pythagorean school's concept of numbers as only whole numbers. The Pythagorean school dealt with proportions by comparison of integer multiples of a common subunit.[33] According to one legend, Hippasus of Metapontum (ca. 470 B.C.) was drowned at sea for making known the existence of the irrational or incommensurable.[34][35]

Complex numbers

 
The absolute value of a complex number z is the distance r from z to the origin.

For any complex number

 

the absolute value or modulus is given by

 

So the three quantities, r, x and y are related by the Pythagorean equation,

 

Note that r is defined to be a positive number or zero but x and y can be negative as well as positive. Geometrically r is the distance of the z from zero or the origin O in the complex plane.

This can be generalised to find the distance between two points, z1 and z2 say. The required distance is given by

 

so again they are related by a version of the Pythagorean equation,

 

Euclidean distance

The distance formula in Cartesian coordinates is derived from the Pythagorean theorem.[36] If (x1, y1) and (x2, y2) are points in the plane, then the distance between them, also called the Euclidean distance, is given by

 

More generally, in Euclidean n-space, the Euclidean distance between two points,   and  , is defined, by generalization of the Pythagorean theorem, as:

 

If instead of Euclidean distance, the square of this value (the squared Euclidean distance, or SED) is used, the resulting equation avoids square roots and is simply a sum of the SED of the coordinates:

 

The squared form is a smooth, convex function of both points, and is widely used in optimization theory and statistics, forming the basis of least squares.

Euclidean distance in other coordinate systems

If Cartesian coordinates are not used, for example, if polar coordinates are used in two dimensions or, in more general terms, if curvilinear coordinates are used, the formulas expressing the Euclidean distance are more complicated than the Pythagorean theorem, but can be derived from it. A typical example where the straight-line distance between two points is converted to curvilinear coordinates can be found in the applications of Legendre polynomials in physics. The formulas can be discovered by using Pythagoras' theorem with the equations relating the curvilinear coordinates to Cartesian coordinates. For example, the polar coordinates (r, θ) can be introduced as:

 

Then two points with locations (r1, θ1) and (r2, θ2) are separated by a distance s:

 

Performing the squares and combining terms, the Pythagorean formula for distance in Cartesian coordinates produces the separation in polar coordinates as:

 

using the trigonometric product-to-sum formulas. This formula is the law of cosines, sometimes called the generalized Pythagorean theorem.[37] From this result, for the case where the radii to the two locations are at right angles, the enclosed angle Δθ = π/2, and the form corresponding to Pythagoras' theorem is regained:   The Pythagorean theorem, valid for right triangles, therefore is a special case of the more general law of cosines, valid for arbitrary triangles.

Pythagorean trigonometric identity

 
Similar right triangles showing sine and cosine of angle θ

In a right triangle with sides a, b and hypotenuse c, trigonometry determines the sine and cosine of the angle θ between side a and the hypotenuse as:

 

From that it follows:

 

where the last step applies Pythagoras' theorem. This relation between sine and cosine is sometimes called the fundamental Pythagorean trigonometric identity.[38] In similar triangles, the ratios of the sides are the same regardless of the size of the triangles, and depend upon the angles. Consequently, in the figure, the triangle with hypotenuse of unit size has opposite side of size sin θ and adjacent side of size cos θ in units of the hypotenuse.

Relation to the cross product

 
The area of a parallelogram as a cross product; vectors a and b identify a plane and a × b is normal to this plane.

The Pythagorean theorem relates the cross product and dot product in a similar way:[39]

 

This can be seen from the definitions of the cross product and dot product, as

 

with n a unit vector normal to both a and b. The relationship follows from these definitions and the Pythagorean trigonometric identity.

This can also be used to define the cross product. By rearranging the following equation is obtained

 

This can be considered as a condition on the cross product and so part of its definition, for example in seven dimensions.[40][41]

Generalizations

Similar figures on the three sides

The Pythagorean theorem generalizes beyond the areas of squares on the three sides to any similar figures. This was known by Hippocrates of Chios in the 5th century BC,[42] and was included by Euclid in his Elements:[43]

If one erects similar figures (see Euclidean geometry) with corresponding sides on the sides of a right triangle, then the sum of the areas of the ones on the two smaller sides equals the area of the one on the larger side.

This extension assumes that the sides of the original triangle are the corresponding sides of the three congruent figures (so the common ratios of sides between the similar figures are a:b:c).[44] While Euclid's proof only applied to convex polygons, the theorem also applies to concave polygons and even to similar figures that have curved boundaries (but still with part of a figure's boundary being the side of the original triangle).[44]

The basic idea behind this generalization is that the area of a plane figure is proportional to the square of any linear dimension, and in particular is proportional to the square of the length of any side. Thus, if similar figures with areas A, B and C are erected on sides with corresponding lengths a, b and c then:

 
 

But, by the Pythagorean theorem, a2 + b2 = c2, so A + B = C.

Conversely, if we can prove that A + B = C for three similar figures without using the Pythagorean theorem, then we can work backwards to construct a proof of the theorem. For example, the starting center triangle can be replicated and used as a triangle C on its hypotenuse, and two similar right triangles (A and B ) constructed on the other two sides, formed by dividing the central triangle by its altitude. The sum of the areas of the two smaller triangles therefore is that of the third, thus A + B = C and reversing the above logic leads to the Pythagorean theorem a2 + b2 = c2. (See also Einstein's proof by dissection without rearrangement)

 
Generalization for similar triangles,
green area A + B = blue area C
 
Pythagoras' theorem using similar right triangles
 
Generalization for regular pentagons

Law of cosines

 
The separation s of two points (r1, θ1) and (r2, θ2) in polar coordinates is given by the law of cosines. Interior angle Δθ = θ1−θ2.

The Pythagorean theorem is a special case of the more general theorem relating the lengths of sides in any triangle, the law of cosines, which states that

 
where   is the angle between sides   and  .[45]

When   is   radians or 90°, then  , and the formula reduces to the usual Pythagorean theorem.

Arbitrary triangle

 
Generalization of Pythagoras' theorem by Tâbit ibn Qorra.[46] Lower panel: reflection of triangle CAD (top) to form triangle DAC, similar to triangle ABC (top).

At any selected angle of a general triangle of sides a, b, c, inscribe an isosceles triangle such that the equal angles at its base θ are the same as the selected angle. Suppose the selected angle θ is opposite the side labeled c. Inscribing the isosceles triangle forms triangle CAD with angle θ opposite side b and with side r along c. A second triangle is formed with angle θ opposite side a and a side with length s along c, as shown in the figure. Thābit ibn Qurra stated that the sides of the three triangles were related as:[47][48]

 

As the angle θ approaches π/2, the base of the isosceles triangle narrows, and lengths r and s overlap less and less. When θ = π/2, ADB becomes a right triangle, r + s = c, and the original Pythagorean theorem is regained.

One proof observes that triangle ABC has the same angles as triangle CAD, but in opposite order. (The two triangles share the angle at vertex A, both contain the angle θ, and so also have the same third angle by the triangle postulate.) Consequently, ABC is similar to the reflection of CAD, the triangle DAC in the lower panel. Taking the ratio of sides opposite and adjacent to θ,

 

Likewise, for the reflection of the other triangle,

 

Clearing fractions and adding these two relations:

 

the required result.

The theorem remains valid if the angle   is obtuse so the lengths r and s are non-overlapping.

General triangles using parallelograms

 
Generalization for arbitrary triangles,
green area = blue area
 
Construction for proof of parallelogram generalization

Pappus's area theorem is a further generalization, that applies to triangles that are not right triangles, using parallelograms on the three sides in place of squares (squares are a special case, of course). The upper figure shows that for a scalene triangle, the area of the parallelogram on the longest side is the sum of the areas of the parallelograms on the other two sides, provided the parallelogram on the long side is constructed as indicated (the dimensions labeled with arrows are the same, and determine the sides of the bottom parallelogram). This replacement of squares with parallelograms bears a clear resemblance to the original Pythagoras' theorem, and was considered a generalization by Pappus of Alexandria in 4 AD[49][50]

The lower figure shows the elements of the proof. Focus on the left side of the figure. The left green parallelogram has the same area as the left, blue portion of the bottom parallelogram because both have the same base b and height h. However, the left green parallelogram also has the same area as the left green parallelogram of the upper figure, because they have the same base (the upper left side of the triangle) and the same height normal to that side of the triangle. Repeating the argument for the right side of the figure, the bottom parallelogram has the same area as the sum of the two green parallelograms.

Solid geometry

 
Pythagoras' theorem in three dimensions relates the diagonal AD to the three sides.
 
A tetrahedron with outward facing right-angle corner

In terms of solid geometry, Pythagoras' theorem can be applied to three dimensions as follows. Consider a rectangular solid as shown in the figure. The length of diagonal BD is found from Pythagoras' theorem as:

 

where these three sides form a right triangle. Using horizontal diagonal BD and the vertical edge AB, the length of diagonal AD then is found by a second application of Pythagoras' theorem as:

 

or, doing it all in one step:

 

This result is the three-dimensional expression for the magnitude of a vector v (the diagonal AD) in terms of its orthogonal components {vk} (the three mutually perpendicular sides):

 

This one-step formulation may be viewed as a generalization of Pythagoras' theorem to higher dimensions. However, this result is really just the repeated application of the original Pythagoras' theorem to a succession of right triangles in a sequence of orthogonal planes.

A substantial generalization of the Pythagorean theorem to three dimensions is de Gua's theorem, named for Jean Paul de Gua de Malves: If a tetrahedron has a right angle corner (like a corner of a cube), then the square of the area of the face opposite the right angle corner is the sum of the squares of the areas of the other three faces. This result can be generalized as in the "n-dimensional Pythagorean theorem":[51]

Let   be orthogonal vectors in Rn. Consider the n-dimensional simplex S with vertices  . (Think of the (n − 1)-dimensional simplex with vertices   not including the origin as the "hypotenuse" of S and the remaining (n − 1)-dimensional faces of S as its "legs".) Then the square of the volume of the hypotenuse of S is the sum of the squares of the volumes of the n legs.

This statement is illustrated in three dimensions by the tetrahedron in the figure. The "hypotenuse" is the base of the tetrahedron at the back of the figure, and the "legs" are the three sides emanating from the vertex in the foreground. As the depth of the base from the vertex increases, the area of the "legs" increases, while that of the base is fixed. The theorem suggests that when this depth is at the value creating a right vertex, the generalization of Pythagoras' theorem applies. In a different wording:[52]

Given an n-rectangular n-dimensional simplex, the square of the (n − 1)-content of the facet opposing the right vertex will equal the sum of the squares of the (n − 1)-contents of the remaining facets.

Inner product spaces

 
Vectors involved in the parallelogram law

The Pythagorean theorem can be generalized to inner product spaces,[53] which are generalizations of the familiar 2-dimensional and 3-dimensional Euclidean spaces. For example, a function may be considered as a vector with infinitely many components in an inner product space, as in functional analysis.[54]

In an inner product space, the concept of perpendicularity is replaced by the concept of orthogonality: two vectors v and w are orthogonal if their inner product   is zero. The inner product is a generalization of the dot product of vectors. The dot product is called the standard inner product or the Euclidean inner product. However, other inner products are possible.[55]

The concept of length is replaced by the concept of the normv‖ of a vector v, defined as:[56]

 

In an inner-product space, the Pythagorean theorem states that for any two orthogonal vectors v and w we have

 

Here the vectors v and w are akin to the sides of a right triangle with hypotenuse given by the vector sum v + w. This form of the Pythagorean theorem is a consequence of the properties of the inner product:

 

where   because of orthogonality.

A further generalization of the Pythagorean theorem in an inner product space to non-orthogonal vectors is the parallelogram law :[56]

 

which says that twice the sum of the squares of the lengths of the sides of a parallelogram is the sum of the squares of the lengths of the diagonals. Any norm that satisfies this equality is ipso facto a norm corresponding to an inner product.[56]

The Pythagorean identity can be extended to sums of more than two orthogonal vectors. If v1, v2, ..., vn are pairwise-orthogonal vectors in an inner-product space, then application of the Pythagorean theorem to successive pairs of these vectors (as described for 3-dimensions in the section on solid geometry) results in the equation[57]

 

Sets of m-dimensional objects in n-dimensional space

Another generalization of the Pythagorean theorem applies to Lebesgue-measurable sets of objects in any number of dimensions. Specifically, the square of the measure of an m-dimensional set of objects in one or more parallel m-dimensional flats in n-dimensional Euclidean space is equal to the sum of the squares of the measures of the orthogonal projections of the object(s) onto all m-dimensional coordinate subspaces.[58]

In mathematical terms:

 

where:

  •   is a measure in m-dimensions (a length in one dimension, an area in two dimensions, a volume in three dimensions, etc.).
  •   is a set of one or more non-overlapping m-dimensional objects in one or more parallel m-dimensional flats in n-dimensional Euclidean space.
  •   is the total measure (sum) of the set of m-dimensional objects.
  •   represents an m-dimensional projection of the original set onto an orthogonal coordinate subspace.
  •   is the measure of the m-dimensional set projection onto m-dimensional coordinate subspace  . Because object projections can overlap on a coordinate subspace, the measure of each object projection in the set must be calculated individually, then measures of all projections added together to provide the total measure for the set of projections on the given coordinate subspace.
  •   is the number of orthogonal, m-dimensional coordinate subspaces in n-dimensional space (Rn) onto which the m-dimensional objects are projected (mn):
     

Non-Euclidean geometry

The Pythagorean theorem is derived from the axioms of Euclidean geometry, and in fact, were the Pythagorean theorem to fail for some right triangle, then the plane in which this triangle is contained cannot be Euclidean. More precisely, the Pythagorean theorem implies, and is implied by, Euclid's Parallel (Fifth) Postulate.[59][60] Thus, right triangles in a non-Euclidean geometry[61] do not satisfy the Pythagorean theorem. For example, in spherical geometry, all three sides of the right triangle (say a, b, and c) bounding an octant of the unit sphere have length equal to π/2, and all its angles are right angles, which violates the Pythagorean theorem because  .

Here two cases of non-Euclidean geometry are considered—spherical geometry and hyperbolic plane geometry; in each case, as in the Euclidean case for non-right triangles, the result replacing the Pythagorean theorem follows from the appropriate law of cosines.

However, the Pythagorean theorem remains true in hyperbolic geometry and elliptic geometry if the condition that the triangle be right is replaced with the condition that two of the angles sum to the third, say A+B = C. The sides are then related as follows: the sum of the areas of the circles with diameters a and b equals the area of the circle with diameter c.[62]

Spherical geometry

 
Spherical triangle

For any right triangle on a sphere of radius R (for example, if γ in the figure is a right angle), with sides a, b, c, the relation between the sides takes the form:[63]

 

This equation can be derived as a special case of the spherical law of cosines that applies to all spherical triangles:

 

For infinitesimal triangles on the sphere (or equivalently, for finite spherical triangles on a sphere of infinite radius), the spherical relation between the sides of a right triangle reduces to the Euclidean form of the Pythagorean theorem. To see how, assume we have a spherical triangle of fixed side lengths a, b, and c on a sphere with expanding radius R. As R approaches infinity the quantities a/R, b/R, and c/R tend to zero and the spherical Pythagorean identity reduces to   so we must look at its asymptotic expansion.

The Maclaurin series for the cosine function can be written as   with the remainder term in big O notation. Letting   be a side of the triangle, and treating the expression as an asymptotic expansion in terms of R for a fixed c,

 

and likewise for a and b. Substituting the asymptotic expansion for each of the cosines into the spherical relation for a right triangle yields

 

Subtracting 1 and then negating each side,

 

Multiplying through by 2R2, the asymptotic expansion for c in terms of fixed a, b and variable R is

 

The Euclidean Pythagorean relationship   is recovered in the limit, as the remainder vanishes when the radius R approaches infinity.

For practical computation in spherical trigonometry with small right triangles, cosines can be replaced with sines using the double-angle identity   to avoid loss of significance. Then the spherical Pythagorean theorem can alternately be written as

 

Hyperbolic geometry

In a hyperbolic space with uniform Gaussian curvature −1/R2, for a right triangle with legs a, b, and hypotenuse c, the relation between the sides takes the form:[64]

 

where cosh is the hyperbolic cosine. This formula is a special form of the hyperbolic law of cosines that applies to all hyperbolic triangles:[65]

 

with γ the angle at the vertex opposite the side c.

By using the Maclaurin series for the hyperbolic cosine, cosh x ≈ 1 + x2/2, it can be shown that as a hyperbolic triangle becomes very small (that is, as a, b, and c all approach zero), the hyperbolic relation for a right triangle approaches the form of Pythagoras' theorem.

For small right triangles (a, b << R), the hyperbolic cosines can be eliminated to avoid loss of significance, giving

 

Very small triangles

For any uniform curvature K (positive, zero, or negative), in very small right triangles (|K|a2, |K|b2 << 1) with hypotenuse c, it can be shown that

 

Differential geometry

 
Distance between infinitesimally separated points in Cartesian coordinates (top) and polar coordinates (bottom), as given by Pythagoras' theorem

The Pythagorean theorem applies to infinitesimal triangles seen in differential geometry. In three dimensional space, the distance between two infinitesimally separated points satisfies

 

with ds the element of distance and (dx, dy, dz) the components of the vector separating the two points. Such a space is called a Euclidean space. However, in Riemannian geometry, a generalization of this expression useful for general coordinates (not just Cartesian) and general spaces (not just Euclidean) takes the form:[66]

 

which is called the metric tensor. (Sometimes, by abuse of language, the same term is applied to the set of coefficients gij.) It may be a function of position, and often describes curved space. A simple example is Euclidean (flat) space expressed in curvilinear coordinates. For example, in polar coordinates:

 

History

There is debate whether the Pythagorean theorem was discovered once, or many times in many places, and the date of first discovery is uncertain, as is the date of the first proof. Historians of Mesopotamian mathematics have concluded that the Pythagorean rule was in widespread use during the Old Babylonian period (20th to 16th centuries BC), over a thousand years before Pythagoras was born.[68][69][70][71] The history of the theorem can be divided into four parts: knowledge of Pythagorean triples, knowledge of the relationship among the sides of a right triangle, knowledge of the relationships among adjacent angles, and proofs of the theorem within some deductive system.

Written c. 1800 BC, the Egyptian Middle Kingdom Berlin Papyrus 6619 includes a problem whose solution is the Pythagorean triple 6:8:10, but the problem does not mention a triangle. The Mesopotamian tablet Plimpton 322, also written c. 1800 BC near Larsa, contains many entries closely related to Pythagorean triples.[72]

In India, the Baudhayana Shulba Sutra, the dates of which are given variously as between the 8th and 5th century BC,[73] contains a list of Pythagorean triples and a statement of the Pythagorean theorem, both in the special case of the isosceles right triangle and in the general case, as does the Apastamba Shulba Sutra (c. 600 BC).[a]

Byzantine Neoplatonic philosopher and mathematician Proclus, writing in the fifth century AD, states two arithmetic rules, "one of them attributed to Plato, the other to Pythagoras",[75] for generating special Pythagorean triples. The rule attributed to Pythagoras (c. 570 – c. 495 BC) starts from an odd number and produces a triple with leg and hypotenuse differing by one unit; the rule attributed to Plato (428/427 or 424/423 – 348/347 BC) starts from an even number and produces a triple with leg and hypotenuse differing by two units. According to Thomas L. Heath (1861–1940), no specific attribution of the theorem to Pythagoras exists in the surviving Greek literature from the five centuries after Pythagoras lived.[76] However, when authors such as Plutarch and Cicero attributed the theorem to Pythagoras, they did so in a way which suggests that the attribution was widely known and undoubted.[77][78] Classicist Kurt von Fritz wrote, "Whether this formula is rightly attributed to Pythagoras personally, but one can safely assume that it belongs to the very oldest period of Pythagorean mathematics."[35] Around 300 BC, in Euclid's Elements, the oldest extant axiomatic proof of the theorem is presented.[79]

 
Geometric proof of the Pythagorean theorem from the Zhoubi Suanjing

With contents known much earlier, but in surviving texts dating from roughly the 1st century BC, the Chinese text Zhoubi Suanjing (周髀算经), (The Arithmetical Classic of the Gnomon and the Circular Paths of Heaven) gives a reasoning for the Pythagorean theorem for the (3, 4, 5) triangle — in China it is called the "Gougu theorem" (勾股定理).[80][81] During the Han Dynasty (202 BC to 220 AD), Pythagorean triples appear in The Nine Chapters on the Mathematical Art,[82] together with a mention of right triangles.[83] Some believe the theorem arose first in China,[84] where it is alternatively known as the "Shang Gao theorem" (商高定理),[85] named after the Duke of Zhou's astronomer and mathematician, whose reasoning composed most of what was in the Zhoubi Suanjing.[86]

See also

Notes and references

Notes

  1. ^ Van der Waerden believed that this material "was certainly based on earlier traditions". Carl Boyer states that the Pythagorean theorem in the Śulba-sũtram may have been influenced by ancient Mesopotamian math, but there is no conclusive evidence in favor or opposition of this possibility.[74]

References

  1. ^ a b Judith D. Sally; Paul Sally (2007). "Chapter 3: Pythagorean triples". Roots to research: a vertical development of mathematical problems. American Mathematical Society Bookstore. p. 63. ISBN 978-0-8218-4403-8.
  2. ^ Benson, Donald. The Moment of Proof : Mathematical Epiphanies, pp. 172–173 (Oxford University Press, 1999).
  3. ^ Euclid (1956), pp. 351–352
  4. ^ Huffman, Carl (23 February 2005). "Pythagoras". In Zalta, Edward N. (ed.). The Stanford Encyclopedia of Philosophy (Winter 2018 Edition)., "It should now be clear that decisions about sources are crucial in addressing the question of whether Pythagoras was a mathematician and scientist. The view of Pythagoras's cosmos sketched in the first five paragraphs of this section, according to which he was neither a mathematician nor a scientist, remains the consensus."
  5. ^ Alexander Bogomolny. "Cut-the-knot.org: Pythagorean theorem and its many proofs, Proof #4". Cut the Knot. Retrieved 4 November 2010.
  6. ^ Alexander Bogomolny. "Cut-the-knot.org: Pythagorean theorem and its many proofs, Proof #3". Cut the Knot. Retrieved 4 November 2010.
  7. ^ (Loomis 1940)
  8. ^ (Maor 2007, p. 39)
  9. ^ Schroeder, Manfred Robert (2012). Fractals, Chaos, Power Laws: Minutes from an Infinite Paradise. Courier Corporation. pp. 3–4. ISBN 978-0486134789.
  10. ^ See for example Pythagorean theorem by shear mapping 2016-10-14 at the Wayback Machine, Saint Louis University website Java applet
  11. ^ Jan Gullberg (1997). Mathematics: from the birth of numbers. W. W. Norton & Company. p. 435. ISBN 0-393-04002-X.
  12. ^ Heiberg, J.L. "Euclid's Elements of Geometry" (PDF). pp. 46–47.
  13. ^ "Euclid's Elements, Book I, Proposition 47". See also a web page version using Java applets by Prof. David E. Joyce, Clark University.
  14. ^ Stephen W. Hawking (2005). God created the integers: the mathematical breakthroughs that changed history. Philadelphia: Running Press Book Publishers. p. 12. ISBN 0-7624-1922-9. This proof first appeared after a computer program was set to check Euclidean proofs.
  15. ^ The proof by Pythagoras probably was not a general one, as the theory of proportions was developed only two centuries after Pythagoras; see (Maor 2007, p. 25)
  16. ^ Alexander Bogomolny. "Pythagorean theorem, proof number 10". Cut the Knot. Retrieved 27 February 2010.
  17. ^ (Loomis 1940, p. 113, Geometric proof 22 and Figure 123)
  18. ^ Polster, Burkard (2004). Q.E.D.: Beauty in Mathematical Proof. Walker Publishing Company. p. 49.
  19. ^ Published in a weekly mathematics column: James A Garfield (1876). "Pons Asinorum". The New England Journal of Education. 3 (14): 161. as noted in William Dunham (1997). The mathematical universe: An alphabetical journey through the great proofs, problems, and personalities. Wiley. p. 96. ISBN 0-471-17661-3. and in A calendar of mathematical dates: April 1, 1876 July 14, 2010, at the Wayback Machine by V. Frederick Rickey
  20. ^ Lantz, David. . Math.Colgate.edu. Archived from the original on 2013-08-28. Retrieved 2018-01-14.
  21. ^ Maor, Eli, The Pythagorean Theorem, Princeton University Press, 2007: pp. 106-107.
  22. ^ Mike Staring (1996). "The Pythagorean proposition: A proof by means of calculus". Mathematics Magazine. Mathematical Association of America. 69 (1): 45–46. doi:10.2307/2691395. JSTOR 2691395.
  23. ^ Bogomolny, Alexander. . Interactive Mathematics Miscellany and Puzzles. Alexander Bogomolny. Archived from the original on 2010-07-06. Retrieved 2010-05-09.
  24. ^ Bruce C. Berndt (1988). "Ramanujan – 100 years old (fashioned) or 100 years new (fangled)?". The Mathematical Intelligencer. 10 (3): 24–31. doi:10.1007/BF03026638. S2CID 123311054.
  25. ^ Judith D. Sally; Paul J. Sally Jr. (2007-12-21). "Theorem 2.4 (Converse of the Pythagorean theorem).". Roots to Research. American Mathematical Society. pp. 54–55. ISBN 978-0-8218-4403-8.
  26. ^ Euclid's Elements, Book I, Proposition 48 From D.E. Joyce's web page at Clark University
  27. ^ Casey, Stephen, "The converse of the theorem of Pythagoras", Mathematical Gazette 92, July 2008, 309–313.
  28. ^ Mitchell, Douglas W., "Feedback on 92.47", Mathematical Gazette 93, March 2009, 156.
  29. ^ Ernest Julius Wilczynski; Herbert Ellsworth Slaught (1914). "Theorem 1 and Theorem 2". Plane trigonometry and applications. Allyn and Bacon. p. 85.
  30. ^ Dijkstra, Edsger W. (September 7, 1986). "On the theorem of Pythagoras". EWD975. E. W. Dijkstra Archive.
  31. ^ Alexander Bogomolny, Pythagorean Theorem for the Reciprocals,https://www.cut-the-knot.org/pythagoras/PTForReciprocals.shtml
  32. ^ Law, Henry (1853). "Corollary 5 of Proposition XLVII (Pythagoras's Theorem)". The Elements of Euclid: with many additional propositions, and explanatory notes, to which is prefixed an introductory essay on logic. John Weale. p. 49.
  33. ^ Shaughan Lavine (1994). Understanding the infinite. Harvard University Press. p. 13. ISBN 0-674-92096-1.
  34. ^ (Heath 1921, Vol I, pp. 65); Hippasus was on a voyage at the time, and his fellows cast him overboard. See James R. Choike (1980). "The pentagram and the discovery of an irrational number". The College Mathematics Journal. 11: 312–316.
  35. ^ a b A careful discussion of Hippasus's contributions is found in Kurt Von Fritz (Apr 1945). "The Discovery of Incommensurability by Hippasus of Metapontum". Annals of Mathematics. Second Series. 46 (2): 242–264. doi:10.2307/1969021. JSTOR 1969021.
  36. ^ Jon Orwant; Jarkko Hietaniemi; John Macdonald (1999). "Euclidean distance". Mastering algorithms with Perl. O'Reilly Media, Inc. p. 426. ISBN 1-56592-398-7.
  37. ^ Wentworth, George (2009). Plane Trigonometry and Tables. BiblioBazaar, LLC. p. 116. ISBN 978-1-103-07998-8., Exercises, page 116
  38. ^ Lawrence S. Leff (2005). PreCalculus the Easy Way (7th ed.). Barron's Educational Series. p. 296. ISBN 0-7641-2892-2.
  39. ^ WS Massey (Dec 1983). (PDF). The American Mathematical Monthly. Mathematical Association of America. 90 (10): 697–701. doi:10.2307/2323537. JSTOR 2323537. S2CID 43318100. Archived from the original (PDF) on 2021-02-26.
  40. ^ Pertti Lounesto (2001). "§7.4 Cross product of two vectors". Clifford algebras and spinors (2nd ed.). Cambridge University Press. p. 96. ISBN 0-521-00551-5.
  41. ^ Francis Begnaud Hildebrand (1992). Methods of applied mathematics (Reprint of Prentice-Hall 1965 2nd ed.). Courier Dover Publications. p. 24. ISBN 0-486-67002-3.
  42. ^ Heath, T. L., A History of Greek Mathematics, Oxford University Press, 1921; reprinted by Dover, 1981.
  43. ^ Euclid's Elements: Book VI, Proposition VI 31: "In right-angled triangles the figure on the side subtending the right angle is equal to the similar and similarly described figures on the sides containing the right angle."
  44. ^ a b Putz, John F. and Sipka, Timothy A. "On generalizing the Pythagorean theorem", The College Mathematics Journal 34 (4), September 2003, pp. 291–295.
  45. ^ Lawrence S. Leff (2005-05-01). cited work. Barron's Educational Series. p. 326. ISBN 0-7641-2892-2.
  46. ^ Howard Whitley Eves (1983). "§4.8:...generalization of Pythagorean theorem". Great moments in mathematics (before 1650). Mathematical Association of America. p. 41. ISBN 0-88385-310-8.
  47. ^ Aydin Sayili (Mar 1960). "Thâbit ibn Qurra's Generalization of the Pythagorean Theorem". Isis. 51 (1): 35–37. doi:10.1086/348837. JSTOR 227603. S2CID 119868978.
  48. ^ Judith D. Sally; Paul Sally (2007-12-21). "Exercise 2.10 (ii)". Roots to Research: A Vertical Development of Mathematical Problems. p. 62. ISBN 978-0-8218-4403-8.
  49. ^ For the details of such a construction, see Jennings, George (1997). "Figure 1.32: The generalized Pythagorean theorem". Modern geometry with applications: with 150 figures (3rd ed.). Springer. p. 23. ISBN 0-387-94222-X.
  50. ^ Claudi Alsina, Roger B. Nelsen: Charming Proofs: A Journey Into Elegant Mathematics. MAA, 2010, ISBN 9780883853481, pp. 77–78 (excerpt, p. 77, at Google Books)
  51. ^ Rajendra Bhatia (1997). Matrix analysis. Springer. p. 21. ISBN 0-387-94846-5.
  52. ^ For an extended discussion of this generalization, see, for example, Willie W. Wong 2009-12-29 at the Wayback Machine 2002, A generalized n-dimensional Pythagorean theorem.
  53. ^ Ferdinand van der Heijden; Dick de Ridder (2004). Classification, parameter estimation, and state estimation. Wiley. p. 357. ISBN 0-470-09013-8.
  54. ^ Qun Lin; Jiafu Lin (2006). Finite element methods: accuracy and improvement. Elsevier. p. 23. ISBN 7-03-016656-6.
  55. ^ Howard Anton; Chris Rorres (2010). Elementary Linear Algebra: Applications Version (10th ed.). Wiley. p. 336. ISBN 978-0-470-43205-1.
  56. ^ a b c Karen Saxe (2002). "Theorem 1.2". Beginning functional analysis. Springer. p. 7. ISBN 0-387-95224-1.
  57. ^ Douglas, Ronald G. (1998). Banach Algebra Techniques in Operator Theory (2nd ed.). New York, New York: Springer-Verlag New York, Inc. pp. 60–61. ISBN 978-0-387-98377-6.
  58. ^ Donald R Conant & William A Beyer (Mar 1974). "Generalized Pythagorean Theorem". The American Mathematical Monthly. Mathematical Association of America. 81 (3): 262–265. doi:10.2307/2319528. JSTOR 2319528.
  59. ^ Eric W. Weisstein (2003). CRC concise encyclopedia of mathematics (2nd ed.). p. 2147. ISBN 1-58488-347-2. The parallel postulate is equivalent to the Equidistance postulate, Playfair axiom, Proclus axiom, the Triangle postulate and the Pythagorean theorem.
  60. ^ Alexander R. Pruss (2006). The principle of sufficient reason: a reassessment. Cambridge University Press. p. 11. ISBN 0-521-85959-X. We could include...the parallel postulate and derive the Pythagorean theorem. Or we could instead make the Pythagorean theorem among the other axioms and derive the parallel postulate.
  61. ^ Stephen W. Hawking (2005). cited work. p. 4. ISBN 0-7624-1922-9.
  62. ^ Victor Pambuccian (December 2010). "Maria Teresa Calapso's Hyperbolic Pythagorean Theorem". The Mathematical Intelligencer. 32 (4): 2. doi:10.1007/s00283-010-9169-0.
  63. ^ Barrett O'Neill (2006). "Exercise 4". Elementary Differential Geometry (2nd ed.). Academic Press. p. 441. ISBN 0-12-088735-5.
  64. ^ Saul Stahl (1993). "Theorem 8.3". The Poincaré half-plane: a gateway to modern geometry. Jones & Bartlett Learning. p. 122. ISBN 0-86720-298-X.
  65. ^ Jane Gilman (1995). "Hyperbolic triangles". Two-generator discrete subgroups of PSL(2,R). American Mathematical Society Bookstore. ISBN 0-8218-0361-1.
  66. ^ Tai L. Chow (2000). Mathematical methods for physicists: a concise introduction. Cambridge University Press. p. 52. ISBN 0-521-65544-7.
  67. ^ Neugebauer 1969, p. 36.
  68. ^ Neugebauer 1969: p. 36 "In other words it was known during the whole duration of Babylonian mathematics that the sum of the squares on the lengths of the sides of a right triangle equals the square of the length of the hypotenuse."
  69. ^ Friberg, Jöran (1981). "Methods and traditions of Babylonian mathematics: Plimpton 322, Pythagorean triples, and the Babylonian triangle parameter equations". Historia Mathematica. 8: 277–318. doi:10.1016/0315-0860(81)90069-0.: p. 306 "Although Plimpton 322 is a unique text of its kind, there are several other known texts testifying that the Pythagorean theorem was well known to the mathematicians of the Old Babylonian period."
  70. ^ Høyrup, Jens. "Pythagorean 'Rule' and 'Theorem' – Mirror of the Relation Between Babylonian and Greek Mathematics". In Renger, Johannes (ed.). Babylon: Focus mesopotamischer Geschichte, Wiege früher Gelehrsamkeit, Mythos in der Moderne. 2. Internationales Colloquium der Deutschen Orient-Gesellschaft 24.–26. März 1998 in Berlin (PDF). Berlin: Deutsche Orient-Gesellschaft / Saarbrücken: SDV Saarbrücker Druckerei und Verlag. pp. 393–407., p. 406, "To judge from this evidence alone it is therefore likely that the Pythagorean rule was discovered within the lay surveyors’ environment, possibly as a spin-off from the problem treated in Db2-146, somewhere between 2300 and 1825 BC." (Db2-146 is an Old Babylonian clay tablet from Eshnunna concerning the computation of the sides of a rectangle given its area and diagonal.)
  71. ^ Robson, E. (2008). Mathematics in Ancient Iraq: A Social History. Princeton University Press.: p. 109 "Many Old Babylonian mathematical practitioners … knew that the square on the diagonal of a right triangle had the same area as the sum of the squares on the length and width: that relationship is used in the worked solutions to word problems on cut-and-paste ‘algebra’ on seven different tablets, from Ešnuna, Sippar, Susa, and an unknown location in southern Babylonia."
  72. ^ Robson, Eleanor (2001). "Neither Sherlock Holmes nor Babylon: a reassessment of Plimpton 322". Historia Mathematica. 28 (3): 167–206. doi:10.1006/hmat.2001.2317.
  73. ^ Kim Plofker (2009). Mathematics in India. Princeton University Press. pp. 17–18. ISBN 978-0-691-12067-6.
  74. ^ Carl Benjamin Boyer; Uta C. Merzbach (2011). "China and India". A history of mathematics (3rd ed.). Wiley. p. 229. ISBN 978-0470525487. Quote: [In Sulba-sutras,] we find rules for the construction of right angles by means of triples of cords the lengths of which form Pythagorean triages, such as 3, 4, and 5, or 5, 12, and 13, or 8, 15, and 17, or 12, 35, and 37. Although Mesopotamian influence in the Sulvasũtras is not unlikely, we know of no conclusive evidence for or against this. Aspastamba knew that the square on the diagonal of a rectangle is equal to the sum of the squares on the two adjacent sides. Less easily explained is another rule given by Apastamba – one that strongly resembles some of the geometric algebra in Book II of Euclid's Elements. (...)
  75. ^ Proclus (1970). A Commentary of the First Book of Euclid's Elements. Translated by Morrow, Glenn R. Princeton University Press. 428.6.
  76. ^ "Introduction and books 1,2". The University Press. March 25, 1908 – via Google Books.
  77. ^ (Heath 1921, Vol I, p. 144): "Though this is the proposition universally associated by tradition with the name of Pythagoras, no really trustworthy evidence exists that it was actually discovered by him. The comparatively late writers who attribute it to him add the story that he sacrificed an ox to celebrate his discovery."
  78. ^ An extensive discussion of the historical evidence is provided in (Euclid 1956, p. 351) page=351
  79. ^ Asger Aaboe (1997). Episodes from the early history of mathematics. Mathematical Association of America. p. 51. ISBN 0-88385-613-1. ...it is not until Euclid that we find a logical sequence of general theorems with proper proofs.
  80. ^ Robert P. Crease (2008). The great equations: breakthroughs in science from Pythagoras to Heisenberg. W W Norton & Co. p. 25. ISBN 978-0-393-06204-5.
  81. ^ A rather extensive discussion of the origins of the various texts in the Zhou Bi is provided by Christopher Cullen (2007). Astronomy and Mathematics in Ancient China: The 'Zhou Bi Suan Jing'. Cambridge University Press. pp. 139 ff. ISBN 978-0-521-03537-8.
  82. ^ This work is a compilation of 246 problems, some of which survived the book burning of 213 BC, and was put in final form before 100 AD. It was extensively commented upon by Liu Hui in 263 AD. Philip D. Straffin Jr. (2004). "Liu Hui and the first golden age of Chinese mathematics". In Marlow Anderson; Victor J. Katz; Robin J. Wilson (eds.). Sherlock Holmes in Babylon: and other tales of mathematical history. Mathematical Association of America. pp. 69 ff. ISBN 0-88385-546-1. See particularly §3: Nine chapters on the mathematical art, pp. 71 ff.
  83. ^ Kangshen Shen; John N. Crossley; Anthony Wah-Cheung Lun (1999). The nine chapters on the mathematical art: companion and commentary. Oxford University Press. p. 488. ISBN 0-19-853936-3.
  84. ^ In particular, Li Jimin; see Centaurus, Volume 39. Copenhagen: Munksgaard. 1997. pp. 193, 205.
  85. ^ Chen, Cheng-Yih (1996). "§3.3.4 Chén Zǐ's formula and the Chóng-Chã method; Figure 40". Early Chinese work in natural science: a re-examination of the physics of motion, acoustics, astronomy and scientific thoughts. Hong Kong University Press. p. 142. ISBN 962-209-385-X.
  86. ^ Wen-tsün Wu (2008). "The Gougu theorem". Selected works of Wen-tsün Wu. World Scientific. p. 158. ISBN 978-981-279-107-8.

Works cited

  • Bell, John L. (1999). The Art of the Intelligible: An Elementary Survey of Mathematics in its Conceptual Development. Kluwer. ISBN 0-7923-5972-0.
  • Euclid (1956). The Thirteen Books of Euclid's Elements, Translated from the Text of Heiberg, with Introduction and Commentary. Vol. 1 (Books I and II). Translated by Heath, Thomas L. (Reprint of 2nd (1925) ed.). Dover. On-line text at archive.org
  • Heath, Sir Thomas (1921). "The 'Theorem of Pythagoras'". A History of Greek Mathematics (2 Vols.) (Dover Publications, Inc. (1981) ed.). Clarendon Press, Oxford. pp. 144 ff. ISBN 0-486-24073-8.
  • Libeskind, Shlomo (2008). Euclidean and transformational geometry: a deductive inquiry. Jones & Bartlett Learning. ISBN 978-0-7637-4366-6. This high-school geometry text covers many of the topics in this WP article.
  • Loomis, Elisha Scott (1940). The Pythagorean Proposition (2nd ed.). Ann Arbor, Michigan: Edwards Brothers. Reissued 1968 by the National Council of Teachers of Mathematics. A lower-quality scan was published online by the Education Resources Information Center, ERIC ED037335.
  • Maor, Eli (2007). The Pythagorean Theorem: A 4,000-Year History. Princeton, New Jersey: Princeton University Press. ISBN 978-0-691-12526-8.
  • Neugebauer, Otto (1969). The exact sciences in antiquity. Acta Historica Scientiarum Naturalium et Medicinalium. Vol. 9 (Republication of 1957 Brown University Press 2nd ed.). Courier Dover Publications. pp. 1–191. ISBN 0-486-22332-9. PMID 14884919.
  • Robson, Eleanor and Jacqueline Stedall, eds., The Oxford Handbook of the History of Mathematics, Oxford: Oxford University Press, 2009. pp. vii + 918. ISBN 978-0-19-921312-2.
  • Stillwell, John (1989). Mathematics and Its History. Springer-Verlag. ISBN 0-387-96981-0. Also ISBN 3-540-96981-0.
  • Swetz, Frank; Kao, T. I. (1977). Was Pythagoras Chinese?: An Examination of Right Triangle Theory in Ancient China. Pennsylvania State University Press. ISBN 0-271-01238-2.
  • van der Waerden, Bartel Leendert (1983). Geometry and Algebra in Ancient Civilizations. Springer. ISBN 3-540-12159-5. Pythagorean triples Babylonian scribes van der Waerden.

External links

  • Euclid (1997) [c. 300 BC]. David E. Joyce (ed.). Elements. Retrieved 2006-08-30. In HTML with Java-based interactive figures.
  • "Pythagorean theorem". Encyclopedia of Mathematics. EMS Press. 2001 [1994].
  • History topic: Pythagoras's theorem in Babylonian mathematics
  • Interactive links:
    • Interactive proof in Java of the Pythagorean theorem
    • Another interactive proof in Java of the Pythagorean theorem
    • Pythagorean theorem with interactive animation
    • Animated, non-algebraic, and user-paced Pythagorean theorem
  • Pythagorean theorem water demo on YouTube
  • Pythagorean theorem (more than 70 proofs from cut-the-knot)
  • Weisstein, Eric W. "Pythagorean theorem". MathWorld.

pythagorean, theorem, mathematics, pythagoras, theorem, fundamental, relation, euclidean, geometry, between, three, sides, right, triangle, states, that, area, square, whose, side, hypotenuse, side, opposite, right, angle, equal, areas, squares, other, sides, . In mathematics the Pythagorean theorem or Pythagoras theorem is a fundamental relation in Euclidean geometry between the three sides of a right triangle It states that the area of the square whose side is the hypotenuse the side opposite the right angle is equal to the sum of the areas of the squares on the other two sides This theorem can be written as an equation relating the lengths of the sides a b and the hypotenuse c often called the Pythagorean equation 1 Pythagorean theoremTypeTheoremFieldEuclidean geometryStatementThe sum of the areas of the two squares on the legs a and b equals the area of the square on the hypotenuse c Symbolic statementa 2 b 2 c 2 displaystyle a 2 b 2 c 2 GeneralizationsLaw of cosines Solid geometry Non Euclidean geometry Differential geometryConsequencesPythagorean triple Reciprocal Pythagorean theorem Complex number Euclidean distance Pythagorean trigonometric identity a 2 b 2 c 2 displaystyle a 2 b 2 c 2 The theorem is named for the Greek philosopher Pythagoras born around 570 BC The theorem has been proved numerous times by many different methods possibly the most for any mathematical theorem The proofs are diverse including both geometric proofs and algebraic proofs with some dating back thousands of years When Euclidean space is represented by a Cartesian coordinate system in analytic geometry Euclidean distance satisfies the Pythagorean relation the squared distance between two points equals the sum of squares of the difference in each coordinate between the points The theorem can be generalized in various ways to higher dimensional spaces to spaces that are not Euclidean to objects that are not right triangles and to objects that are not triangles at all but n dimensional solids The Pythagorean theorem has attracted interest outside mathematics as a symbol of mathematical abstruseness mystique or intellectual power popular references in literature plays musicals songs stamps and cartoons abound Contents 1 Proofs using constructed squares 1 1 Rearrangement proofs 1 2 Algebraic proofs 2 Other proofs of the theorem 2 1 Proof using similar triangles 2 2 Einstein s proof by dissection without rearrangement 2 3 Trigonometric proof using Einstein s construction 2 4 Euclid s proof 2 5 Proofs by dissection and rearrangement 2 6 Proof by area preserving shearing 2 7 Other algebraic proofs 2 8 Proof using differentials 3 Converse 4 Consequences and uses of the theorem 4 1 Pythagorean triples 4 2 Inverse Pythagorean theorem 4 3 Incommensurable lengths 4 4 Complex numbers 4 5 Euclidean distance 4 6 Euclidean distance in other coordinate systems 4 7 Pythagorean trigonometric identity 4 8 Relation to the cross product 5 Generalizations 5 1 Similar figures on the three sides 5 2 Law of cosines 5 3 Arbitrary triangle 5 4 General triangles using parallelograms 5 5 Solid geometry 5 6 Inner product spaces 5 7 Sets of m dimensional objects in n dimensional space 5 8 Non Euclidean geometry 5 8 1 Spherical geometry 5 8 2 Hyperbolic geometry 5 8 3 Very small triangles 5 9 Differential geometry 6 History 7 See also 8 Notes and references 8 1 Notes 8 2 References 8 3 Works cited 9 External linksProofs using constructed squares Rearrangement proof of the Pythagorean theorem The area of the white space remains constant throughout the translation rearrangement of the triangles At all moments in time the area is always c And likewise at all moments in time the area is always a b Rearrangement proofs In one rearrangement proof two squares are used whose sides have a measure of a b displaystyle a b and which contain four right triangles whose sides are a b and c with the hypotenuse being c In the square on the right side the triangles are placed such that the corners of the square correspond to the corners of the right angle in the triangles forming a square in the center whose sides are length c Each outer square has an area of a b 2 displaystyle a b 2 as well as 2 a b c 2 displaystyle 2ab c 2 with 2 a b displaystyle 2ab representing the total area of the four triangles Within the big square on the left side the four triangles are moved to form two similar rectangles with sides of length a and b These rectangles in their new position have now delineated two new squares one having side length a is formed in the bottom left corner and another square of side length b formed in the top right corner In this new position this left side now has a square of area a b 2 displaystyle a b 2 as well as 2 a b a 2 b 2 displaystyle 2ab a 2 b 2 Since both squares have the area of a b 2 displaystyle a b 2 it follows that the other measure of the square area also equal each other such that 2 a b c 2 displaystyle 2ab c 2 2 a b a 2 b 2 displaystyle 2ab a 2 b 2 With the area of the four triangles removed from both side of the equation what remains is a 2 b 2 c 2 displaystyle a 2 b 2 c 2 2 In another proof rectangles in the second box can also be placed such that both have one corner that correspond to consecutive corners of the square In this way they also form two boxes this time in consecutive corners with areas a 2 displaystyle a 2 and b 2 displaystyle b 2 which will again lead to a second square of with the area 2 a b a 2 b 2 displaystyle 2ab a 2 b 2 English mathematician Sir Thomas Heath gives this proof in his commentary on Proposition I 47 in Euclid s Elements and mentions the proposals of German mathematicians Carl Anton Bretschneider and Hermann Hankel that Pythagoras may have known this proof Heath himself favors a different proposal for a Pythagorean proof but acknowledges from the outset of his discussion that the Greek literature which we possess belonging to the first five centuries after Pythagoras contains no statement specifying this or any other particular great geometric discovery to him 3 Recent scholarship has cast increasing doubt on any sort of role for Pythagoras as a creator of mathematics although debate about this continues 4 Algebraic proofs Diagram of the two algebraic proofs The theorem can be proved algebraically using four copies of the same triangle arranged symmetrically around a square with side c as shown in the lower part of the diagram 5 This results in a larger square with side a b and area a b 2 The four triangles and the square side c must have the same area as the larger square b a 2 c 2 4 a b 2 c 2 2 a b displaystyle b a 2 c 2 4 frac ab 2 c 2 2ab giving c 2 b a 2 2 a b b 2 2 a b a 2 2 a b a 2 b 2 displaystyle c 2 b a 2 2ab b 2 2ab a 2 2ab a 2 b 2 A similar proof uses four copies of a right triangle with sides a b and c arranged inside a square with side c as in the top half of the diagram 6 The triangles are similar with area 1 2 a b displaystyle tfrac 1 2 ab while the small square has side b a and area b a 2 The area of the large square is therefore b a 2 4 a b 2 b a 2 2 a b b 2 2 a b a 2 2 a b a 2 b 2 displaystyle b a 2 4 frac ab 2 b a 2 2ab b 2 2ab a 2 2ab a 2 b 2 But this is a square with side c and area c2 so c 2 a 2 b 2 displaystyle c 2 a 2 b 2 Other proofs of the theoremThis theorem may have more known proofs than any other the law of quadratic reciprocity being another contender for that distinction the book The Pythagorean Proposition contains 370 proofs 7 Proof using similar triangles In this section and as usual in geometry a word of two capital letters such as AB denotes the length of the line segment defined by the points labeled with the letters and not a multiplication So AB2 denotes the square of the length AB and not the product A B 2 displaystyle A times B 2 Proof using similar triangles This proof is based on the proportionality of the sides of three similar triangles that is upon the fact that the ratio of any two corresponding sides of similar triangles is the same regardless of the size of the triangles Let ABC represent a right triangle with the right angle located at C as shown on the figure Draw the altitude from point C and call H its intersection with the side AB Point H divides the length of the hypotenuse c into parts d and e The new triangle ACH is similar to triangle ABC because they both have a right angle by definition of the altitude and they share the angle at A meaning that the third angle will be the same in both triangles as well marked as 8 in the figure By a similar reasoning the triangle CBH is also similar to ABC The proof of similarity of the triangles requires the triangle postulate The sum of the angles in a triangle is two right angles and is equivalent to the parallel postulate Similarity of the triangles leads to the equality of ratios of corresponding sides B C A B B H B C and A C A B A H A C displaystyle frac BC AB frac BH BC text and frac AC AB frac AH AC The first result equates the cosines of the angles 8 whereas the second result equates their sines These ratios can be written as B C 2 A B B H and A C 2 A B A H displaystyle BC 2 AB times BH text and AC 2 AB times AH Summing these two equalities results in B C 2 A C 2 A B B H A B A H A B A H B H A B 2 displaystyle BC 2 AC 2 AB times BH AB times AH AB AH BH AB 2 which after simplification demonstrates the Pythagorean theorem B C 2 A C 2 A B 2 displaystyle BC 2 AC 2 AB 2 The role of this proof in history is the subject of much speculation The underlying question is why Euclid did not use this proof but invented another One conjecture is that the proof by similar triangles involved a theory of proportions a topic not discussed until later in the Elements and that the theory of proportions needed further development at that time 8 Einstein s proof by dissection without rearrangement Albert Einstein gave a proof by dissection in which the pieces do not need to be moved 9 Instead of using a square on the hypotenuse and two squares on the legs one can use any other shape that includes the hypotenuse and two similar shapes that each include one of two legs instead of the hypotenuse see Similar figures on the three sides In Einstein s proof the shape that includes the hypotenuse is the right triangle itself The dissection consists of dropping a perpendicular from the vertex of the right angle of the triangle to the hypotenuse thus splitting the whole triangle into two parts Those two parts have the same shape as the original right triangle and have the legs of the original triangle as their hypotenuses and the sum of their areas is that of the original triangle Because the ratio of the area of a right triangle to the square of its hypotenuse is the same for similar triangles the relationship between the areas of the three triangles holds for the squares of the sides of the large triangle as well Trigonometric proof using Einstein s construction Right triangle on the hypotenuse dissected into two similar right triangles on the legs according to Einstein s proof Both the proof using similar triangles and Einstein s proof rely on constructing the height to the hypotenuse of the right triangle A B C displaystyle triangle ABC The same construction provides a trigonometric proof of the Pythagorean theorem using the definition of the sine as a ratio inside a right triangle sin a a c displaystyle sin alpha frac a c sin b b c displaystyle sin beta frac b c c b sin b a sin a b 2 c a 2 c displaystyle c b sin beta a sin alpha frac b 2 c frac a 2 c and thus c 2 a 2 b 2 displaystyle c 2 a 2 b 2 This proof is essentially the same as the above proof using similar triangles where some ratios of lengths are replaced by sines Euclid s proof Proof in Euclid s Elements In outline here is how the proof in Euclid s Elements proceeds The large square is divided into a left and right rectangle A triangle is constructed that has half the area of the left rectangle Then another triangle is constructed that has half the area of the square on the left most side These two triangles are shown to be congruent proving this square has the same area as the left rectangle This argument is followed by a similar version for the right rectangle and the remaining square Putting the two rectangles together to reform the square on the hypotenuse its area is the same as the sum of the area of the other two squares The details follow Let A B C be the vertices of a right triangle with a right angle at A Drop a perpendicular from A to the side opposite the hypotenuse in the square on the hypotenuse That line divides the square on the hypotenuse into two rectangles each having the same area as one of the two squares on the legs For the formal proof we require four elementary lemmata If two triangles have two sides of the one equal to two sides of the other each to each and the angles included by those sides equal then the triangles are congruent side angle side The area of a triangle is half the area of any parallelogram on the same base and having the same altitude The area of a rectangle is equal to the product of two adjacent sides The area of a square is equal to the product of two of its sides follows from 3 Next each top square is related to a triangle congruent with another triangle related in turn to one of two rectangles making up the lower square 10 Illustration including the new lines Showing the two congruent triangles of half the area of rectangle BDLK and square BAGF The proof is as follows Let ACB be a right angled triangle with right angle CAB On each of the sides BC AB and CA squares are drawn CBDE BAGF and ACIH in that order The construction of squares requires the immediately preceding theorems in Euclid and depends upon the parallel postulate 11 From A draw a line parallel to BD and CE It will perpendicularly intersect BC and DE at K and L respectively Join CF and AD to form the triangles BCF and BDA Angles CAB and BAG are both right angles therefore C A and G are collinear Angles CBD and FBA are both right angles therefore angle ABD equals angle FBC since both are the sum of a right angle and angle ABC Since AB is equal to FB BD is equal to BC and angle ABD equals angle FBC triangle ABD must be congruent to triangle FBC Since A K L is a straight line parallel to BD then rectangle BDLK has twice the area of triangle ABD because they share the base BD and have the same altitude BK i e a line normal to their common base connecting the parallel lines BD and AL lemma 2 Since C is collinear with A and G and this line is parallel to FB then square BAGF must be twice in area to triangle FBC Therefore rectangle BDLK must have the same area as square BAGF AB2 By applying steps 3 to 10 to the other side of the figure it can be similarly shown that rectangle CKLE must have the same area as square ACIH AC2 Adding these two results AB2 AC2 BD BK KL KC Since BD KL BD BK KL KC BD BK KC BD BC Therefore AB2 AC2 BC2 since CBDE is a square This proof which appears in Euclid s Elements as that of Proposition 47 in Book 1 demonstrates that the area of the square on the hypotenuse is the sum of the areas of the other two squares 12 13 This is quite distinct from the proof by similarity of triangles which is conjectured to be the proof that Pythagoras used 14 15 Proofs by dissection and rearrangement Another by rearrangement is given by the middle animation A large square is formed with area c2 from four identical right triangles with sides a b and c fitted around a small central square Then two rectangles are formed with sides a and b by moving the triangles Combining the smaller square with these rectangles produces two squares of areas a2 and b2 which must have the same area as the initial large square 16 The third rightmost image also gives a proof The upper two squares are divided as shown by the blue and green shading into pieces that when rearranged can be made to fit in the lower square on the hypotenuse or conversely the large square can be divided as shown into pieces that fill the other two This way of cutting one figure into pieces and rearranging them to get another figure is called dissection This shows the area of the large square equals that of the two smaller ones 17 Animation showing proof by rearrangement of four identical right triangles Animation showing another proof by rearrangement Proof using an elaborate rearrangementProof by area preserving shearing Visual proof of the Pythagorean theorem by area preserving shearing As shown in the accompanying animation area preserving shear mappings and translations can transform the squares on the sides adjacent to the right angle onto the square on the hypotenuse together covering it exactly 18 Each shear leaves the base and height unchanged thus leaving the area unchanged too The translations also leave the area unchanged as they do not alter the shapes at all Each square is first sheared into a parallelogram and then into a rectangle which can be translated onto one section of the square on the hypotenuse Other algebraic proofs A related proof was published by future U S President James A Garfield then a U S Representative see diagram 19 20 21 Instead of a square it uses a trapezoid which can be constructed from the square in the second of the above proofs by bisecting along a diagonal of the inner square to give the trapezoid as shown in the diagram The area of the trapezoid can be calculated to be half the area of the square that is 1 2 b a 2 displaystyle frac 1 2 b a 2 The inner square is similarly halved and there are only two triangles so the proof proceeds as above except for a factor of 1 2 displaystyle frac 1 2 which is removed by multiplying by two to give the result Proof using differentials One can arrive at the Pythagorean theorem by studying how changes in a side produce a change in the hypotenuse and employing calculus 22 23 24 The triangle ABC is a right triangle as shown in the upper part of the diagram with BC the hypotenuse At the same time the triangle lengths are measured as shown with the hypotenuse of length y the side AC of length x and the side AB of length a as seen in the lower diagram part Diagram for differential proof If x is increased by a small amount dx by extending the side AC slightly to D then y also increases by dy These form two sides of a triangle CDE which with E chosen so CE is perpendicular to the hypotenuse is a right triangle approximately similar to ABC Therefore the ratios of their sides must be the same that is d y d x x y displaystyle frac dy dx frac x y This can be rewritten as y d y x d x displaystyle y dy x dx which is a differential equation that can be solved by direct integration y d y x d x displaystyle int y dy int x dx giving y 2 x 2 C displaystyle y 2 x 2 C The constant can be deduced from x 0 y a to give the equation y 2 x 2 a 2 displaystyle y 2 x 2 a 2 This is more of an intuitive proof than a formal one it can be made more rigorous if proper limits are used in place of dx and dy ConverseThe converse of the theorem is also true 25 Given a triangle with sides of length a b and c if a2 b2 c2 then the angle between sides a and b is a right angle For any three positive real numbers a b and c such that a2 b2 c2 there exists a triangle with sides a b and c as a consequence of the converse of the triangle inequality This converse appears in Euclid s Elements Book I 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 26 It can be proved using the law of cosines or as follows Let ABC be a triangle with side lengths a b and c with a2 b2 c2 Construct a second triangle with sides of length a and b containing a right angle By the Pythagorean theorem it follows that the hypotenuse of this triangle has length c a2 b2 the same as the hypotenuse of the first triangle Since both triangles sides are the same lengths a b and c the triangles are congruent and must have the same angles Therefore the angle between the side of lengths a and b in the original triangle is a right angle The above proof of the converse makes use of the Pythagorean theorem itself The converse can also be proved without assuming the Pythagorean theorem 27 28 A corollary of the Pythagorean theorem s converse is a simple means of determining whether a triangle is right obtuse or acute as follows Let c be chosen to be the longest of the three sides and a b gt c otherwise there is no triangle according to the triangle inequality The following statements apply 29 If a2 b2 c2 then the triangle is right If a2 b2 gt c2 then the triangle is acute If a2 b2 lt c2 then the triangle is obtuse Edsger W Dijkstra has stated this proposition about acute right and obtuse triangles in this language sgn a b g sgn a2 b2 c2 where a is the angle opposite to side a b is the angle opposite to side b g is the angle opposite to side c and sgn is the sign function 30 Consequences and uses of the theoremPythagorean triples Main article Pythagorean triple A Pythagorean triple has three positive integers a b and c such that a2 b2 c2 In other words a Pythagorean triple represents the lengths of the sides of a right triangle where all three sides have integer lengths 1 Such a triple is commonly written a b c Some well known examples are 3 4 5 and 5 12 13 A primitive Pythagorean triple is one in which a b and c are coprime the greatest common divisor of a b and c is 1 The following is a list of primitive Pythagorean triples with values less than 100 3 4 5 5 12 13 7 24 25 8 15 17 9 40 41 11 60 61 12 35 37 13 84 85 16 63 65 20 21 29 28 45 53 33 56 65 36 77 85 39 80 89 48 55 73 65 72 97 Inverse Pythagorean theorem Given a right triangle with sides a b c displaystyle a b c and altitude d displaystyle d a line from the right angle and perpendicular to the hypotenuse c displaystyle c The Pythagorean theorem has a 2 b 2 c 2 displaystyle a 2 b 2 c 2 while the inverse Pythagorean theorem relates the two legs a b displaystyle a b to the altitude d displaystyle d 31 1 a 2 1 b 2 1 d 2 displaystyle frac 1 a 2 frac 1 b 2 frac 1 d 2 The equation can be transformed to 1 x z 2 1 y z 2 1 x y 2 displaystyle frac 1 xz 2 frac 1 yz 2 frac 1 xy 2 where x 2 y 2 z 2 displaystyle x 2 y 2 z 2 for any non zero real x y z displaystyle x y z If the a b d displaystyle a b d are to be integers the smallest solution a gt b gt d displaystyle a gt b gt d is then 1 20 2 1 15 2 1 12 2 displaystyle frac 1 20 2 frac 1 15 2 frac 1 12 2 using the smallest Pythagorean triple 3 4 5 displaystyle 3 4 5 The reciprocal Pythagorean theorem is a special case of the optic equation 1 p 1 q 1 r displaystyle frac 1 p frac 1 q frac 1 r where the denominators are squares and also for a heptagonal triangle whose sides p q r displaystyle p q r are square numbers Incommensurable lengths The spiral of Theodorus A construction for line segments with lengths whose ratios are the square root of a positive integer One of the consequences of the Pythagorean theorem is that line segments whose lengths are incommensurable so the ratio of which is not a rational number can be constructed using a straightedge and compass Pythagoras theorem enables construction of incommensurable lengths because the hypotenuse of a triangle is related to the sides by the square root operation The figure on the right shows how to construct line segments whose lengths are in the ratio of the square root of any positive integer 32 Each triangle has a side labeled 1 that is the chosen unit for measurement In each right triangle Pythagoras theorem establishes the length of the hypotenuse in terms of this unit If a hypotenuse is related to the unit by the square root of a positive integer that is not a perfect square it is a realization of a length incommensurable with the unit such as 2 3 5 For more detail see Quadratic irrational Incommensurable lengths conflicted with the Pythagorean school s concept of numbers as only whole numbers The Pythagorean school dealt with proportions by comparison of integer multiples of a common subunit 33 According to one legend Hippasus of Metapontum ca 470 B C was drowned at sea for making known the existence of the irrational or incommensurable 34 35 Complex numbers The absolute value of a complex number z is the distance r from z to the origin For any complex number z x i y displaystyle z x iy the absolute value or modulus is given by r z x 2 y 2 displaystyle r z sqrt x 2 y 2 So the three quantities r x and y are related by the Pythagorean equation r 2 x 2 y 2 displaystyle r 2 x 2 y 2 Note that r is defined to be a positive number or zero but x and y can be negative as well as positive Geometrically r is the distance of the z from zero or the origin O in the complex plane This can be generalised to find the distance between two points z1 and z2 say The required distance is given by z 1 z 2 x 1 x 2 2 y 1 y 2 2 displaystyle z 1 z 2 sqrt x 1 x 2 2 y 1 y 2 2 so again they are related by a version of the Pythagorean equation z 1 z 2 2 x 1 x 2 2 y 1 y 2 2 displaystyle z 1 z 2 2 x 1 x 2 2 y 1 y 2 2 Euclidean distance Main article Euclidean distance The distance formula in Cartesian coordinates is derived from the Pythagorean theorem 36 If x1 y1 and x2 y2 are points in the plane then the distance between them also called the Euclidean distance is given by x 1 x 2 2 y 1 y 2 2 displaystyle sqrt x 1 x 2 2 y 1 y 2 2 More generally in Euclidean n space the Euclidean distance between two points A a 1 a 2 a n displaystyle A a 1 a 2 dots a n and B b 1 b 2 b n displaystyle B b 1 b 2 dots b n is defined by generalization of the Pythagorean theorem as a 1 b 1 2 a 2 b 2 2 a n b n 2 i 1 n a i b i 2 displaystyle sqrt a 1 b 1 2 a 2 b 2 2 cdots a n b n 2 sqrt sum i 1 n a i b i 2 If instead of Euclidean distance the square of this value the squared Euclidean distance or SED is used the resulting equation avoids square roots and is simply a sum of the SED of the coordinates a 1 b 1 2 a 2 b 2 2 a n b n 2 i 1 n a i b i 2 displaystyle a 1 b 1 2 a 2 b 2 2 cdots a n b n 2 sum i 1 n a i b i 2 The squared form is a smooth convex function of both points and is widely used in optimization theory and statistics forming the basis of least squares Euclidean distance in other coordinate systems If Cartesian coordinates are not used for example if polar coordinates are used in two dimensions or in more general terms if curvilinear coordinates are used the formulas expressing the Euclidean distance are more complicated than the Pythagorean theorem but can be derived from it A typical example where the straight line distance between two points is converted to curvilinear coordinates can be found in the applications of Legendre polynomials in physics The formulas can be discovered by using Pythagoras theorem with the equations relating the curvilinear coordinates to Cartesian coordinates For example the polar coordinates r 8 can be introduced as x r cos 8 y r sin 8 displaystyle x r cos theta y r sin theta Then two points with locations r1 81 and r2 82 are separated by a distance s s 2 x 1 x 2 2 y 1 y 2 2 r 1 cos 8 1 r 2 cos 8 2 2 r 1 sin 8 1 r 2 sin 8 2 2 displaystyle s 2 x 1 x 2 2 y 1 y 2 2 r 1 cos theta 1 r 2 cos theta 2 2 r 1 sin theta 1 r 2 sin theta 2 2 Performing the squares and combining terms the Pythagorean formula for distance in Cartesian coordinates produces the separation in polar coordinates as s 2 r 1 2 r 2 2 2 r 1 r 2 cos 8 1 cos 8 2 sin 8 1 sin 8 2 r 1 2 r 2 2 2 r 1 r 2 cos 8 1 8 2 r 1 2 r 2 2 2 r 1 r 2 cos D 8 displaystyle begin aligned s 2 amp r 1 2 r 2 2 2r 1 r 2 left cos theta 1 cos theta 2 sin theta 1 sin theta 2 right amp r 1 2 r 2 2 2r 1 r 2 cos left theta 1 theta 2 right amp r 1 2 r 2 2 2r 1 r 2 cos Delta theta end aligned using the trigonometric product to sum formulas This formula is the law of cosines sometimes called the generalized Pythagorean theorem 37 From this result for the case where the radii to the two locations are at right angles the enclosed angle D8 p 2 and the form corresponding to Pythagoras theorem is regained s 2 r 1 2 r 2 2 displaystyle s 2 r 1 2 r 2 2 The Pythagorean theorem valid for right triangles therefore is a special case of the more general law of cosines valid for arbitrary triangles Pythagorean trigonometric identity Main article Pythagorean trigonometric identity Similar right triangles showing sine and cosine of angle 8 In a right triangle with sides a b and hypotenuse c trigonometry determines the sine and cosine of the angle 8 between side a and the hypotenuse as sin 8 b c cos 8 a c displaystyle sin theta frac b c quad cos theta frac a c From that it follows cos 2 8 sin 2 8 a 2 b 2 c 2 1 displaystyle cos 2 theta sin 2 theta frac a 2 b 2 c 2 1 where the last step applies Pythagoras theorem This relation between sine and cosine is sometimes called the fundamental Pythagorean trigonometric identity 38 In similar triangles the ratios of the sides are the same regardless of the size of the triangles and depend upon the angles Consequently in the figure the triangle with hypotenuse of unit size has opposite side of size sin 8 and adjacent side of size cos 8 in units of the hypotenuse Relation to the cross product The area of a parallelogram as a cross product vectors a and b identify a plane and a b is normal to this plane The Pythagorean theorem relates the cross product and dot product in a similar way 39 a b 2 a b 2 a 2 b 2 displaystyle mathbf a times mathbf b 2 mathbf a cdot mathbf b 2 mathbf a 2 mathbf b 2 This can be seen from the definitions of the cross product and dot product as a b a b n sin 8 a b a b cos 8 displaystyle begin aligned mathbf a times mathbf b amp ab mathbf n sin theta mathbf a cdot mathbf b amp ab cos theta end aligned with n a unit vector normal to both a and b The relationship follows from these definitions and the Pythagorean trigonometric identity This can also be used to define the cross product By rearranging the following equation is obtained a b 2 a 2 b 2 a b 2 displaystyle mathbf a times mathbf b 2 mathbf a 2 mathbf b 2 mathbf a cdot mathbf b 2 This can be considered as a condition on the cross product and so part of its definition for example in seven dimensions 40 41 GeneralizationsSimilar figures on the three sides The Pythagorean theorem generalizes beyond the areas of squares on the three sides to any similar figures This was known by Hippocrates of Chios in the 5th century BC 42 and was included by Euclid in his Elements 43 If one erects similar figures see Euclidean geometry with corresponding sides on the sides of a right triangle then the sum of the areas of the ones on the two smaller sides equals the area of the one on the larger side This extension assumes that the sides of the original triangle are the corresponding sides of the three congruent figures so the common ratios of sides between the similar figures are a b c 44 While Euclid s proof only applied to convex polygons the theorem also applies to concave polygons and even to similar figures that have curved boundaries but still with part of a figure s boundary being the side of the original triangle 44 The basic idea behind this generalization is that the area of a plane figure is proportional to the square of any linear dimension and in particular is proportional to the square of the length of any side Thus if similar figures with areas A B and C are erected on sides with corresponding lengths a b and c then A a 2 B b 2 C c 2 displaystyle frac A a 2 frac B b 2 frac C c 2 A B a 2 c 2 C b 2 c 2 C displaystyle Rightarrow A B frac a 2 c 2 C frac b 2 c 2 C But by the Pythagorean theorem a2 b2 c2 so A B C Conversely if we can prove that A B C for three similar figures without using the Pythagorean theorem then we can work backwards to construct a proof of the theorem For example the starting center triangle can be replicated and used as a triangle C on its hypotenuse and two similar right triangles A and B constructed on the other two sides formed by dividing the central triangle by its altitude The sum of the areas of the two smaller triangles therefore is that of the third thus A B C and reversing the above logic leads to the Pythagorean theorem a2 b2 c2 See also Einstein s proof by dissection without rearrangement Generalization for similar triangles green area A B blue area C Pythagoras theorem using similar right triangles Generalization for regular pentagonsLaw of cosines Main article Law of cosines The separation s of two points r1 81 and r2 82 in polar coordinates is given by the law of cosines Interior angle D8 81 82 The Pythagorean theorem is a special case of the more general theorem relating the lengths of sides in any triangle the law of cosines which states thata 2 b 2 2 a b cos 8 c 2 displaystyle a 2 b 2 2ab cos theta c 2 where 8 displaystyle theta is the angle between sides a displaystyle a and b displaystyle b 45 When 8 displaystyle theta is p 2 displaystyle frac pi 2 radians or 90 then cos 8 0 displaystyle cos theta 0 and the formula reduces to the usual Pythagorean theorem Arbitrary triangle Generalization of Pythagoras theorem by Tabit ibn Qorra 46 Lower panel reflection of triangle CAD top to form triangle DAC similar to triangle ABC top At any selected angle of a general triangle of sides a b c inscribe an isosceles triangle such that the equal angles at its base 8 are the same as the selected angle Suppose the selected angle 8 is opposite the side labeled c Inscribing the isosceles triangle forms triangle CAD with angle 8 opposite side b and with side r along c A second triangle is formed with angle 8 opposite side a and a side with length s along c as shown in the figure Thabit ibn Qurra stated that the sides of the three triangles were related as 47 48 a 2 b 2 c r s displaystyle a 2 b 2 c r s As the angle 8 approaches p 2 the base of the isosceles triangle narrows and lengths r and s overlap less and less When 8 p 2 ADB becomes a right triangle r s c and the original Pythagorean theorem is regained One proof observes that triangle ABC has the same angles as triangle CAD but in opposite order The two triangles share the angle at vertex A both contain the angle 8 and so also have the same third angle by the triangle postulate Consequently ABC is similar to the reflection of CAD the triangle DAC in the lower panel Taking the ratio of sides opposite and adjacent to 8 c b b r displaystyle frac c b frac b r Likewise for the reflection of the other triangle c a a s displaystyle frac c a frac a s Clearing fractions and adding these two relations c s c r a 2 b 2 displaystyle cs cr a 2 b 2 the required result The theorem remains valid if the angle 8 displaystyle theta is obtuse so the lengths r and s are non overlapping General triangles using parallelograms Generalization for arbitrary triangles green area blue area Construction for proof of parallelogram generalization Pappus s area theorem is a further generalization that applies to triangles that are not right triangles using parallelograms on the three sides in place of squares squares are a special case of course The upper figure shows that for a scalene triangle the area of the parallelogram on the longest side is the sum of the areas of the parallelograms on the other two sides provided the parallelogram on the long side is constructed as indicated the dimensions labeled with arrows are the same and determine the sides of the bottom parallelogram This replacement of squares with parallelograms bears a clear resemblance to the original Pythagoras theorem and was considered a generalization by Pappus of Alexandria in 4 AD 49 50 The lower figure shows the elements of the proof Focus on the left side of the figure The left green parallelogram has the same area as the left blue portion of the bottom parallelogram because both have the same base b and height h However the left green parallelogram also has the same area as the left green parallelogram of the upper figure because they have the same base the upper left side of the triangle and the same height normal to that side of the triangle Repeating the argument for the right side of the figure the bottom parallelogram has the same area as the sum of the two green parallelograms Solid geometry Pythagoras theorem in three dimensions relates the diagonal AD to the three sides A tetrahedron with outward facing right angle corner In terms of solid geometry Pythagoras theorem can be applied to three dimensions as follows Consider a rectangular solid as shown in the figure The length of diagonal BD is found from Pythagoras theorem as B D 2 B C 2 C D 2 displaystyle overline BD 2 overline BC 2 overline CD 2 where these three sides form a right triangle Using horizontal diagonal BD and the vertical edge AB the length of diagonal AD then is found by a second application of Pythagoras theorem as A D 2 A B 2 B D 2 displaystyle overline AD 2 overline AB 2 overline BD 2 or doing it all in one step A D 2 A B 2 B C 2 C D 2 displaystyle overline AD 2 overline AB 2 overline BC 2 overline CD 2 This result is the three dimensional expression for the magnitude of a vector v the diagonal AD in terms of its orthogonal components vk the three mutually perpendicular sides v 2 k 1 3 v k 2 displaystyle mathbf v 2 sum k 1 3 mathbf v k 2 This one step formulation may be viewed as a generalization of Pythagoras theorem to higher dimensions However this result is really just the repeated application of the original Pythagoras theorem to a succession of right triangles in a sequence of orthogonal planes A substantial generalization of the Pythagorean theorem to three dimensions is de Gua s theorem named for Jean Paul de Gua de Malves If a tetrahedron has a right angle corner like a corner of a cube then the square of the area of the face opposite the right angle corner is the sum of the squares of the areas of the other three faces This result can be generalized as in the n dimensional Pythagorean theorem 51 Let x 1 x 2 x n displaystyle x 1 x 2 ldots x n be orthogonal vectors in Rn Consider the n dimensional simplex S with vertices 0 x 1 x n displaystyle 0 x 1 ldots x n Think of the n 1 dimensional simplex with vertices x 1 x n displaystyle x 1 ldots x n not including the origin as the hypotenuse of S and the remaining n 1 dimensional faces of S as its legs Then the square of the volume of the hypotenuse of S is the sum of the squares of the volumes of the n legs This statement is illustrated in three dimensions by the tetrahedron in the figure The hypotenuse is the base of the tetrahedron at the back of the figure and the legs are the three sides emanating from the vertex in the foreground As the depth of the base from the vertex increases the area of the legs increases while that of the base is fixed The theorem suggests that when this depth is at the value creating a right vertex the generalization of Pythagoras theorem applies In a different wording 52 Given an n rectangular n dimensional simplex the square of the n 1 content of the facet opposing the right vertex will equal the sum of the squares of the n 1 contents of the remaining facets Inner product spaces Vectors involved in the parallelogram law The Pythagorean theorem can be generalized to inner product spaces 53 which are generalizations of the familiar 2 dimensional and 3 dimensional Euclidean spaces For example a function may be considered as a vector with infinitely many components in an inner product space as in functional analysis 54 In an inner product space the concept of perpendicularity is replaced by the concept of orthogonality two vectors v and w are orthogonal if their inner product v w displaystyle langle mathbf v mathbf w rangle is zero The inner product is a generalization of the dot product of vectors The dot product is called the standard inner product or the Euclidean inner product However other inner products are possible 55 The concept of length is replaced by the concept of the norm v of a vector v defined as 56 v v v displaystyle lVert mathbf v rVert equiv sqrt langle mathbf v mathbf v rangle In an inner product space the Pythagorean theorem states that for any two orthogonal vectors v and w we have v w 2 v 2 w 2 displaystyle left mathbf v mathbf w right 2 left mathbf v right 2 left mathbf w right 2 Here the vectors v and w are akin to the sides of a right triangle with hypotenuse given by the vector sum v w This form of the Pythagorean theorem is a consequence of the properties of the inner product v w 2 v w v w v v w w v w w v v 2 w 2 displaystyle begin aligned left mathbf v mathbf w right 2 amp langle mathbf v w mathbf v w rangle 3mu amp langle mathbf v mathbf v rangle langle mathbf w mathbf w rangle langle mathbf v w rangle langle mathbf w v rangle 3mu amp left mathbf v right 2 left mathbf w right 2 end aligned where v w w v 0 displaystyle langle mathbf v w rangle langle mathbf w v rangle 0 because of orthogonality A further generalization of the Pythagorean theorem in an inner product space to non orthogonal vectors is the parallelogram law 56 2 v 2 2 w 2 v w 2 v w 2 displaystyle 2 mathbf v 2 2 mathbf w 2 mathbf v w 2 mathbf v w 2 which says that twice the sum of the squares of the lengths of the sides of a parallelogram is the sum of the squares of the lengths of the diagonals Any norm that satisfies this equality is ipso facto a norm corresponding to an inner product 56 The Pythagorean identity can be extended to sums of more than two orthogonal vectors If v1 v2 vn are pairwise orthogonal vectors in an inner product space then application of the Pythagorean theorem to successive pairs of these vectors as described for 3 dimensions in the section on solid geometry results in the equation 57 k 1 n v k 2 k 1 n v k 2 displaystyle left sum k 1 n mathbf v k right 2 sum k 1 n mathbf v k 2 Sets of m dimensional objects in n dimensional space Another generalization of the Pythagorean theorem applies to Lebesgue measurable sets of objects in any number of dimensions Specifically the square of the measure of an m dimensional set of objects in one or more parallel m dimensional flats in n dimensional Euclidean space is equal to the sum of the squares of the measures of the orthogonal projections of the object s onto all m dimensional coordinate subspaces 58 In mathematical terms m m s 2 i 1 x m 2 m p i displaystyle mu ms 2 sum i 1 x mathbf mu 2 mp i where m m displaystyle mu m is a measure in m dimensions a length in one dimension an area in two dimensions a volume in three dimensions etc s displaystyle s is a set of one or more non overlapping m dimensional objects in one or more parallel m dimensional flats in n dimensional Euclidean space m m s displaystyle mu ms is the total measure sum of the set of m dimensional objects p displaystyle p represents an m dimensional projection of the original set onto an orthogonal coordinate subspace m m p i displaystyle mu mp i is the measure of the m dimensional set projection onto m dimensional coordinate subspace i displaystyle i Because object projections can overlap on a coordinate subspace the measure of each object projection in the set must be calculated individually then measures of all projections added together to provide the total measure for the set of projections on the given coordinate subspace x displaystyle x is the number of orthogonal m dimensional coordinate subspaces in n dimensional space Rn onto which the m dimensional objects are projected m n x n m n m n m displaystyle x binom n m frac n m n m Non Euclidean geometry The Pythagorean theorem is derived from the axioms of Euclidean geometry and in fact were the Pythagorean theorem to fail for some right triangle then the plane in which this triangle is contained cannot be Euclidean More precisely the Pythagorean theorem implies and is implied by Euclid s Parallel Fifth Postulate 59 60 Thus right triangles in a non Euclidean geometry 61 do not satisfy the Pythagorean theorem For example in spherical geometry all three sides of the right triangle say a b and c bounding an octant of the unit sphere have length equal to p 2 and all its angles are right angles which violates the Pythagorean theorem because a 2 b 2 2 c 2 gt c 2 displaystyle a 2 b 2 2c 2 gt c 2 Here two cases of non Euclidean geometry are considered spherical geometry and hyperbolic plane geometry in each case as in the Euclidean case for non right triangles the result replacing the Pythagorean theorem follows from the appropriate law of cosines However the Pythagorean theorem remains true in hyperbolic geometry and elliptic geometry if the condition that the triangle be right is replaced with the condition that two of the angles sum to the third say A B C The sides are then related as follows the sum of the areas of the circles with diameters a and b equals the area of the circle with diameter c 62 Spherical geometry Spherical triangle For any right triangle on a sphere of radius R for example if g in the figure is a right angle with sides a b c the relation between the sides takes the form 63 cos c R cos a R cos b R displaystyle cos frac c R cos frac a R cos frac b R This equation can be derived as a special case of the spherical law of cosines that applies to all spherical triangles cos c R cos a R cos b R sin a R sin b R cos g displaystyle cos frac c R cos frac a R cos frac b R sin frac a R sin frac b R cos gamma For infinitesimal triangles on the sphere or equivalently for finite spherical triangles on a sphere of infinite radius the spherical relation between the sides of a right triangle reduces to the Euclidean form of the Pythagorean theorem To see how assume we have a spherical triangle of fixed side lengths a b and c on a sphere with expanding radius R As R approaches infinity the quantities a R b R and c R tend to zero and the spherical Pythagorean identity reduces to 1 1 displaystyle 1 1 so we must look at its asymptotic expansion The Maclaurin series for the cosine function can be written as cos x 1 1 2 x 2 O x 4 textstyle cos x 1 tfrac 1 2 x 2 O left x 4 right with the remainder term in big O notation Letting x c R displaystyle x c R be a side of the triangle and treating the expression as an asymptotic expansion in terms of R for a fixed c cos c R 1 c 2 2 R 2 O R 4 displaystyle begin aligned cos frac c R 1 frac c 2 2R 2 O left R 4 right end aligned and likewise for a and b Substituting the asymptotic expansion for each of the cosines into the spherical relation for a right triangle yields 1 c 2 2 R 2 O R 4 1 a 2 2 R 2 O R 4 1 b 2 2 R 2 O R 4 1 a 2 2 R 2 b 2 2 R 2 O R 4 displaystyle begin aligned 1 frac c 2 2R 2 O left R 4 right amp left 1 frac a 2 2R 2 O left R 4 right right left 1 frac b 2 2R 2 O left R 4 right right amp 1 frac a 2 2R 2 frac b 2 2R 2 O left R 4 right end aligned Subtracting 1 and then negating each side c 2 2 R 2 a 2 2 R 2 b 2 2 R 2 O R 4 displaystyle frac c 2 2R 2 frac a 2 2R 2 frac b 2 2R 2 O left R 4 right Multiplying through by 2R2 the asymptotic expansion for c in terms of fixed a b and variable R is c 2 a 2 b 2 O R 2 displaystyle c 2 a 2 b 2 O left R 2 right The Euclidean Pythagorean relationship c 2 a 2 b 2 textstyle c 2 a 2 b 2 is recovered in the limit as the remainder vanishes when the radius R approaches infinity For practical computation in spherical trigonometry with small right triangles cosines can be replaced with sines using the double angle identity cos 2 8 1 2 sin 2 8 displaystyle cos 2 theta 1 2 sin 2 theta to avoid loss of significance Then the spherical Pythagorean theorem can alternately be written as sin 2 c 2 R sin 2 a 2 R sin 2 b 2 R 2 sin 2 a 2 R sin 2 b 2 R displaystyle sin 2 frac c 2R sin 2 frac a 2R sin 2 frac b 2R 2 sin 2 frac a 2R sin 2 frac b 2R Hyperbolic geometry Hyperbolic triangle In a hyperbolic space with uniform Gaussian curvature 1 R2 for a right triangle with legs a b and hypotenuse c the relation between the sides takes the form 64 cosh c R cosh a R cosh b R displaystyle cosh frac c R cosh frac a R cosh frac b R where cosh is the hyperbolic cosine This formula is a special form of the hyperbolic law of cosines that applies to all hyperbolic triangles 65 cosh c R cosh a R cosh b R sinh a R sinh b R cos g displaystyle cosh frac c R cosh frac a R cosh frac b R sinh frac a R sinh frac b R cos gamma with g the angle at the vertex opposite the side c By using the Maclaurin series for the hyperbolic cosine cosh x 1 x2 2 it can be shown that as a hyperbolic triangle becomes very small that is as a b and c all approach zero the hyperbolic relation for a right triangle approaches the form of Pythagoras theorem For small right triangles a b lt lt R the hyperbolic cosines can be eliminated to avoid loss of significance giving sinh 2 c 2 R sinh 2 a 2 R sinh 2 b 2 R 2 sinh 2 a 2 R sinh 2 b 2 R displaystyle sinh 2 frac c 2R sinh 2 frac a 2R sinh 2 frac b 2R 2 sinh 2 frac a 2R sinh 2 frac b 2R Very small triangles For any uniform curvature K positive zero or negative in very small right triangles K a2 K b2 lt lt 1 with hypotenuse c it can be shown that c 2 a 2 b 2 K 3 a 2 b 2 K 2 45 a 2 b 2 a 2 b 2 2 K 3 945 a 2 b 2 a 2 b 2 2 O K 4 c 10 displaystyle c 2 a 2 b 2 frac K 3 a 2 b 2 frac K 2 45 a 2 b 2 a 2 b 2 frac 2K 3 945 a 2 b 2 a 2 b 2 2 O K 4 c 10 Differential geometry Distance between infinitesimally separated points in Cartesian coordinates top and polar coordinates bottom as given by Pythagoras theorem The Pythagorean theorem applies to infinitesimal triangles seen in differential geometry In three dimensional space the distance between two infinitesimally separated points satisfies d s 2 d x 2 d y 2 d z 2 displaystyle ds 2 dx 2 dy 2 dz 2 with ds the element of distance and dx dy dz the components of the vector separating the two points Such a space is called a Euclidean space However in Riemannian geometry a generalization of this expression useful for general coordinates not just Cartesian and general spaces not just Euclidean takes the form 66 d s 2 i j n g i j d x i d x j displaystyle ds 2 sum i j n g ij dx i dx j which is called the metric tensor Sometimes by abuse of language the same term is applied to the set of coefficients gij It may be a function of position and often describes curved space A simple example is Euclidean flat space expressed in curvilinear coordinates For example in polar coordinates d s 2 d r 2 r 2 d 8 2 displaystyle ds 2 dr 2 r 2 d theta 2 History The Plimpton 322 tablet records Pythagorean triples from Babylonian times 67 There is debate whether the Pythagorean theorem was discovered once or many times in many places and the date of first discovery is uncertain as is the date of the first proof Historians of Mesopotamian mathematics have concluded that the Pythagorean rule was in widespread use during the Old Babylonian period 20th to 16th centuries BC over a thousand years before Pythagoras was born 68 69 70 71 The history of the theorem can be divided into four parts knowledge of Pythagorean triples knowledge of the relationship among the sides of a right triangle knowledge of the relationships among adjacent angles and proofs of the theorem within some deductive system Written c 1800 BC the Egyptian Middle Kingdom Berlin Papyrus 6619 includes a problem whose solution is the Pythagorean triple 6 8 10 but the problem does not mention a triangle The Mesopotamian tablet Plimpton 322 also written c 1800 BC near Larsa contains many entries closely related to Pythagorean triples 72 In India the Baudhayana Shulba Sutra the dates of which are given variously as between the 8th and 5th century BC 73 contains a list of Pythagorean triples and a statement of the Pythagorean theorem both in the special case of the isosceles right triangle and in the general case as does the Apastamba Shulba Sutra c 600 BC a Byzantine Neoplatonic philosopher and mathematician Proclus writing in the fifth century AD states two arithmetic rules one of them attributed to Plato the other to Pythagoras 75 for generating special Pythagorean triples The rule attributed to Pythagoras c 570 c 495 BC starts from an odd number and produces a triple with leg and hypotenuse differing by one unit the rule attributed to Plato 428 427 or 424 423 348 347 BC starts from an even number and produces a triple with leg and hypotenuse differing by two units According to Thomas L Heath 1861 1940 no specific attribution of the theorem to Pythagoras exists in the surviving Greek literature from the five centuries after Pythagoras lived 76 However when authors such as Plutarch and Cicero attributed the theorem to Pythagoras they did so in a way which suggests that the attribution was widely known and undoubted 77 78 Classicist Kurt von Fritz wrote Whether this formula is rightly attributed to Pythagoras personally but one can safely assume that it belongs to the very oldest period of Pythagorean mathematics 35 Around 300 BC in Euclid s Elements the oldest extant axiomatic proof of the theorem is presented 79 Geometric proof of the Pythagorean theorem from the Zhoubi Suanjing With contents known much earlier but in surviving texts dating from roughly the 1st century BC the Chinese text Zhoubi Suanjing 周髀算经 The Arithmetical Classic of the Gnomon and the Circular Paths of Heaven gives a reasoning for the Pythagorean theorem for the 3 4 5 triangle in China it is called the Gougu theorem 勾股定理 80 81 During the Han Dynasty 202 BC to 220 AD Pythagorean triples appear in The Nine Chapters on the Mathematical Art 82 together with a mention of right triangles 83 Some believe the theorem arose first in China 84 where it is alternatively known as the Shang Gao theorem 商高定理 85 named after the Duke of Zhou s astronomer and mathematician whose reasoning composed most of what was in the Zhoubi Suanjing 86 See also Mathematics portalAddition in quadrature At Dulcarnon British flag theorem Fermat s Last Theorem Inverse Pythagorean theorem Kepler triangle Linear algebra List of triangle topics Lp space Nonhypotenuse number Parallelogram law Parseval s identity Ptolemy s theorem Pythagorean expectation Pythagorean tiling Rational trigonometry in Pythagoras theorem Thales theoremNotes and referencesNotes Van der Waerden believed that this material was certainly based on earlier traditions Carl Boyer states that the Pythagorean theorem in the Sulba sũtram may have been influenced by ancient Mesopotamian math but there is no conclusive evidence in favor or opposition of this possibility 74 References a b Judith D Sally Paul Sally 2007 Chapter 3 Pythagorean triples Roots to research a vertical development of mathematical problems American Mathematical Society Bookstore p 63 ISBN 978 0 8218 4403 8 Benson Donald The Moment of Proof Mathematical Epiphanies pp 172 173 Oxford University Press 1999 Euclid 1956 pp 351 352 Huffman Carl 23 February 2005 Pythagoras In Zalta Edward N ed The Stanford Encyclopedia of Philosophy Winter 2018 Edition It should now be clear that decisions about sources are crucial in addressing the question of whether Pythagoras was a mathematician and scientist The view of Pythagoras s cosmos sketched in the first five paragraphs of this section according to which he was neither a mathematician nor a scientist remains the consensus Alexander Bogomolny Cut the knot org Pythagorean theorem and its many proofs Proof 4 Cut the Knot Retrieved 4 November 2010 Alexander Bogomolny Cut the knot org Pythagorean theorem and its many proofs Proof 3 Cut the Knot Retrieved 4 November 2010 Loomis 1940 Maor 2007 p 39 Schroeder Manfred Robert 2012 Fractals Chaos Power Laws Minutes from an Infinite Paradise Courier Corporation pp 3 4 ISBN 978 0486134789 See for example Pythagorean theorem by shear mapping Archived 2016 10 14 at the Wayback Machine Saint Louis University website Java applet Jan Gullberg 1997 Mathematics from the birth of numbers W W Norton amp Company p 435 ISBN 0 393 04002 X Heiberg J L Euclid s Elements of Geometry PDF pp 46 47 Euclid s Elements Book I Proposition 47 See also a web page version using Java applets by Prof David E Joyce Clark University Stephen W Hawking 2005 God created the integers the mathematical breakthroughs that changed history Philadelphia Running Press Book Publishers p 12 ISBN 0 7624 1922 9 This proof first appeared after a computer program was set to check Euclidean proofs The proof by Pythagoras probably was not a general one as the theory of proportions was developed only two centuries after Pythagoras see Maor 2007 p 25 Alexander Bogomolny Pythagorean theorem proof number 10 Cut the Knot Retrieved 27 February 2010 Loomis 1940 p 113 Geometric proof 22 and Figure 123 Polster Burkard 2004 Q E D Beauty in Mathematical Proof Walker Publishing Company p 49 Published in a weekly mathematics column James A Garfield 1876 Pons Asinorum The New England Journal of Education 3 14 161 as noted in William Dunham 1997 The mathematical universe An alphabetical journey through the great proofs problems and personalities Wiley p 96 ISBN 0 471 17661 3 and in A calendar of mathematical dates April 1 1876 Archived July 14 2010 at the Wayback Machine by V Frederick Rickey Lantz David Garfield s proof of the Pythagorean Theorem Math Colgate edu Archived from the original on 2013 08 28 Retrieved 2018 01 14 Maor Eli The Pythagorean Theorem Princeton University Press 2007 pp 106 107 Mike Staring 1996 The Pythagorean proposition A proof by means of calculus Mathematics Magazine Mathematical Association of America 69 1 45 46 doi 10 2307 2691395 JSTOR 2691395 Bogomolny Alexander Pythagorean Theorem Interactive Mathematics Miscellany and Puzzles Alexander Bogomolny Archived from the original on 2010 07 06 Retrieved 2010 05 09 Bruce C Berndt 1988 Ramanujan 100 years old fashioned or 100 years new fangled The Mathematical Intelligencer 10 3 24 31 doi 10 1007 BF03026638 S2CID 123311054 Judith D Sally Paul J Sally Jr 2007 12 21 Theorem 2 4 Converse of the Pythagorean theorem Roots to Research American Mathematical Society pp 54 55 ISBN 978 0 8218 4403 8 Euclid s Elements Book I Proposition 48 From D E Joyce s web page at Clark University Casey Stephen The converse of the theorem of Pythagoras Mathematical Gazette 92 July 2008 309 313 Mitchell Douglas W Feedback on 92 47 Mathematical Gazette 93 March 2009 156 Ernest Julius Wilczynski Herbert Ellsworth Slaught 1914 Theorem 1 and Theorem 2 Plane trigonometry and applications Allyn and Bacon p 85 Dijkstra Edsger W September 7 1986 On the theorem of Pythagoras EWD975 E W Dijkstra Archive Alexander Bogomolny Pythagorean Theorem for the Reciprocals https www cut the knot org pythagoras PTForReciprocals shtml Law Henry 1853 Corollary 5 of Proposition XLVII Pythagoras s Theorem The Elements of Euclid with many additional propositions and explanatory notes to which is prefixed an introductory essay on logic John Weale p 49 Shaughan Lavine 1994 Understanding the infinite Harvard University Press p 13 ISBN 0 674 92096 1 Heath 1921 Vol I pp 65 Hippasus was on a voyage at the time and his fellows cast him overboard See James R Choike 1980 The pentagram and the discovery of an irrational number The College Mathematics Journal 11 312 316 a b A careful discussion of Hippasus s contributions is found in Kurt Von Fritz Apr 1945 The Discovery of Incommensurability by Hippasus of Metapontum Annals of Mathematics Second Series 46 2 242 264 doi 10 2307 1969021 JSTOR 1969021 Jon Orwant Jarkko Hietaniemi John Macdonald 1999 Euclidean distance Mastering algorithms with Perl O Reilly Media Inc p 426 ISBN 1 56592 398 7 Wentworth George 2009 Plane Trigonometry and Tables BiblioBazaar LLC p 116 ISBN 978 1 103 07998 8 Exercises page 116 Lawrence S Leff 2005 PreCalculus the Easy Way 7th ed Barron s Educational Series p 296 ISBN 0 7641 2892 2 WS Massey Dec 1983 Cross products of vectors in higher dimensional Euclidean spaces PDF The American Mathematical Monthly Mathematical Association of America 90 10 697 701 doi 10 2307 2323537 JSTOR 2323537 S2CID 43318100 Archived from the original PDF on 2021 02 26 Pertti Lounesto 2001 7 4 Cross product of two vectors Clifford algebras and spinors 2nd ed Cambridge University Press p 96 ISBN 0 521 00551 5 Francis Begnaud Hildebrand 1992 Methods of applied mathematics Reprint of Prentice Hall 1965 2nd ed Courier Dover Publications p 24 ISBN 0 486 67002 3 Heath T L A History of Greek Mathematics Oxford University Press 1921 reprinted by Dover 1981 Euclid s Elements Book VI Proposition VI 31 In right angled triangles the figure on the side subtending the right angle is equal to the similar and similarly described figures on the sides containing the right angle a b Putz John F and Sipka Timothy A On generalizing the Pythagorean theorem The College Mathematics Journal 34 4 September 2003 pp 291 295 Lawrence S Leff 2005 05 01 cited work Barron s Educational Series p 326 ISBN 0 7641 2892 2 Howard Whitley Eves 1983 4 8 generalization of Pythagorean theorem Great moments in mathematics before 1650 Mathematical Association of America p 41 ISBN 0 88385 310 8 Aydin Sayili Mar 1960 Thabit ibn Qurra s Generalization of the Pythagorean Theorem Isis 51 1 35 37 doi 10 1086 348837 JSTOR 227603 S2CID 119868978 Judith D Sally Paul Sally 2007 12 21 Exercise 2 10 ii Roots to Research A Vertical Development of Mathematical Problems p 62 ISBN 978 0 8218 4403 8 For the details of such a construction see Jennings George 1997 Figure 1 32 The generalized Pythagorean theorem Modern geometry with applications with 150 figures 3rd ed Springer p 23 ISBN 0 387 94222 X Claudi Alsina Roger B Nelsen Charming Proofs A Journey Into Elegant Mathematics MAA 2010 ISBN 9780883853481 pp 77 78 excerpt p 77 at Google Books Rajendra Bhatia 1997 Matrix analysis Springer p 21 ISBN 0 387 94846 5 For an extended discussion of this generalization see for example Willie W Wong Archived 2009 12 29 at the Wayback Machine 2002 A generalized n dimensional Pythagorean theorem Ferdinand van der Heijden Dick de Ridder 2004 Classification parameter estimation and state estimation Wiley p 357 ISBN 0 470 09013 8 Qun Lin Jiafu Lin 2006 Finite element methods accuracy and improvement Elsevier p 23 ISBN 7 03 016656 6 Howard Anton Chris Rorres 2010 Elementary Linear Algebra Applications Version 10th ed Wiley p 336 ISBN 978 0 470 43205 1 a b c Karen Saxe 2002 Theorem 1 2 Beginning functional analysis Springer p 7 ISBN 0 387 95224 1 Douglas Ronald G 1998 Banach Algebra Techniques in Operator Theory 2nd ed New York New York Springer Verlag New York Inc pp 60 61 ISBN 978 0 387 98377 6 Donald R Conant amp William A Beyer Mar 1974 Generalized Pythagorean Theorem The American Mathematical Monthly Mathematical Association of America 81 3 262 265 doi 10 2307 2319528 JSTOR 2319528 Eric W Weisstein 2003 CRC concise encyclopedia of mathematics 2nd ed p 2147 ISBN 1 58488 347 2 The parallel postulate is equivalent to the Equidistance postulate Playfair axiom Proclus axiom the Triangle postulate and the Pythagorean theorem Alexander R Pruss 2006 The principle of sufficient reason a reassessment Cambridge University Press p 11 ISBN 0 521 85959 X We could include the parallel postulate and derive the Pythagorean theorem Or we could instead make the Pythagorean theorem among the other axioms and derive the parallel postulate Stephen W Hawking 2005 cited work p 4 ISBN 0 7624 1922 9 Victor Pambuccian December 2010 Maria Teresa Calapso s Hyperbolic Pythagorean Theorem The Mathematical Intelligencer 32 4 2 doi 10 1007 s00283 010 9169 0 Barrett O Neill 2006 Exercise 4 Elementary Differential Geometry 2nd ed Academic Press p 441 ISBN 0 12 088735 5 Saul Stahl 1993 Theorem 8 3 The Poincare half plane a gateway to modern geometry Jones amp Bartlett Learning p 122 ISBN 0 86720 298 X Jane Gilman 1995 Hyperbolic triangles Two generator discrete subgroups of PSL 2 R American Mathematical Society Bookstore ISBN 0 8218 0361 1 Tai L Chow 2000 Mathematical methods for physicists a concise introduction Cambridge University Press p 52 ISBN 0 521 65544 7 Neugebauer 1969 p 36 Neugebauer 1969 p 36 In other words it was known during the whole duration of Babylonian mathematics that the sum of the squares on the lengths of the sides of a right triangle equals the square of the length of the hypotenuse Friberg Joran 1981 Methods and traditions of Babylonian mathematics Plimpton 322 Pythagorean triples and the Babylonian triangle parameter equations Historia Mathematica 8 277 318 doi 10 1016 0315 0860 81 90069 0 p 306 Although Plimpton 322 is a unique text of its kind there are several other known texts testifying that the Pythagorean theorem was well known to the mathematicians of the Old Babylonian period Hoyrup Jens Pythagorean Rule and Theorem Mirror of the Relation Between Babylonian and Greek Mathematics In Renger Johannes ed Babylon Focus mesopotamischer Geschichte Wiege fruher Gelehrsamkeit Mythos in der Moderne 2 Internationales Colloquium der Deutschen Orient Gesellschaft 24 26 Marz 1998 in Berlin PDF Berlin Deutsche Orient Gesellschaft Saarbrucken SDV Saarbrucker Druckerei und Verlag pp 393 407 p 406 To judge from this evidence alone it is therefore likely that the Pythagorean rule was discovered within the lay surveyors environment possibly as a spin off from the problem treated in Db2 146 somewhere between 2300 and 1825 BC Db2 146 is an Old Babylonian clay tablet from Eshnunna concerning the computation of the sides of a rectangle given its area and diagonal Robson E 2008 Mathematics in Ancient Iraq A Social History Princeton University Press p 109 Many Old Babylonian mathematical practitioners knew that the square on the diagonal of a right triangle had the same area as the sum of the squares on the length and width that relationship is used in the worked solutions to word problems on cut and paste algebra on seven different tablets from Esnuna Sippar Susa and an unknown location in southern Babylonia Robson Eleanor 2001 Neither Sherlock Holmes nor Babylon a reassessment of Plimpton 322 Historia Mathematica 28 3 167 206 doi 10 1006 hmat 2001 2317 Kim Plofker 2009 Mathematics in India Princeton University Press pp 17 18 ISBN 978 0 691 12067 6 Carl Benjamin Boyer Uta C Merzbach 2011 China and India A history of mathematics 3rd ed Wiley p 229 ISBN 978 0470525487 Quote In Sulba sutras we find rules for the construction of right angles by means of triples of cords the lengths of which form Pythagorean triages such as 3 4 and 5 or 5 12 and 13 or 8 15 and 17 or 12 35 and 37 Although Mesopotamian influence in the Sulvasũtras is not unlikely we know of no conclusive evidence for or against this Aspastamba knew that the square on the diagonal of a rectangle is equal to the sum of the squares on the two adjacent sides Less easily explained is another rule given by Apastamba one that strongly resembles some of the geometric algebra in Book II of Euclid s Elements Proclus 1970 A Commentary of the First Book of Euclid sElements Translated by Morrow Glenn R Princeton University Press 428 6 Introduction and books 1 2 The University Press March 25 1908 via Google Books Heath 1921 Vol I p 144 Though this is the proposition universally associated by tradition with the name of Pythagoras no really trustworthy evidence exists that it was actually discovered by him The comparatively late writers who attribute it to him add the story that he sacrificed an ox to celebrate his discovery An extensive discussion of the historical evidence is provided in Euclid 1956 p 351 page 351 Asger Aaboe 1997 Episodes from the early history of mathematics Mathematical Association of America p 51 ISBN 0 88385 613 1 it is not until Euclid that we find a logical sequence of general theorems with proper proofs Robert P Crease 2008 The great equations breakthroughs in science from Pythagoras to Heisenberg W W Norton amp Co p 25 ISBN 978 0 393 06204 5 A rather extensive discussion of the origins of the various texts in the Zhou Bi is provided by Christopher Cullen 2007 Astronomy and Mathematics in Ancient China The Zhou Bi Suan Jing Cambridge University Press pp 139 ff ISBN 978 0 521 03537 8 This work is a compilation of 246 problems some of which survived the book burning of 213 BC and was put in final form before 100 AD It was extensively commented upon by Liu Hui in 263 AD Philip D Straffin Jr 2004 Liu Hui and the first golden age of Chinese mathematics In Marlow Anderson Victor J Katz Robin J Wilson eds Sherlock Holmes in Babylon and other tales of mathematical history Mathematical Association of America pp 69 ff ISBN 0 88385 546 1 See particularly 3 Nine chapters on the mathematical art pp 71 ff Kangshen Shen John N Crossley Anthony Wah Cheung Lun 1999 The nine chapters on the mathematical art companion and commentary Oxford University Press p 488 ISBN 0 19 853936 3 In particular Li Jimin see Centaurus Volume 39 Copenhagen Munksgaard 1997 pp 193 205 Chen Cheng Yih 1996 3 3 4 Chen Zǐ s formula and the Chong Cha method Figure 40 Early Chinese work in natural science a re examination of the physics of motion acoustics astronomy and scientific thoughts Hong Kong University Press p 142 ISBN 962 209 385 X Wen tsun Wu 2008 The Gougu theorem Selected works of Wen tsun Wu World Scientific p 158 ISBN 978 981 279 107 8 Works cited Bell John L 1999 The Art of the Intelligible An Elementary Survey of Mathematics in its Conceptual Development Kluwer ISBN 0 7923 5972 0 Euclid 1956 The Thirteen Books of Euclid s Elements Translated from the Text of Heiberg with Introduction and Commentary Vol 1 Books I and II Translated by Heath Thomas L Reprint of 2nd 1925 ed Dover On line text at archive org Heath Sir Thomas 1921 The Theorem of Pythagoras A History of Greek Mathematics 2 Vols Dover Publications Inc 1981 ed Clarendon Press Oxford pp 144 ff ISBN 0 486 24073 8 Libeskind Shlomo 2008 Euclidean and transformational geometry a deductive inquiry Jones amp Bartlett Learning ISBN 978 0 7637 4366 6 This high school geometry text covers many of the topics in this WP article Loomis Elisha Scott 1940 The Pythagorean Proposition 2nd ed Ann Arbor Michigan Edwards Brothers Reissued 1968 by the National Council of Teachers of Mathematics A lower quality scan was published online by the Education Resources Information Center ERIC ED037335 Maor Eli 2007 The Pythagorean Theorem A 4 000 Year History Princeton New Jersey Princeton University Press ISBN 978 0 691 12526 8 Neugebauer Otto 1969 The exact sciences in antiquity Acta Historica Scientiarum Naturalium et Medicinalium Vol 9 Republication of 1957 Brown University Press 2nd ed Courier Dover Publications pp 1 191 ISBN 0 486 22332 9 PMID 14884919 Robson Eleanor and Jacqueline Stedall eds The Oxford Handbook of the History of Mathematics Oxford Oxford University Press 2009 pp vii 918 ISBN 978 0 19 921312 2 Stillwell John 1989 Mathematics and Its History Springer Verlag ISBN 0 387 96981 0 Also ISBN 3 540 96981 0 Swetz Frank Kao T I 1977 Was Pythagoras Chinese An Examination of Right Triangle Theory in Ancient China Pennsylvania State University Press ISBN 0 271 01238 2 van der Waerden Bartel Leendert 1983 Geometry and Algebra in Ancient Civilizations Springer ISBN 3 540 12159 5 Pythagorean triples Babylonian scribes van der Waerden External linksPythagorean theorem at ProofWiki Wikimedia Commons has media related to Pythagorean theorem Euclid 1997 c 300 BC David E Joyce ed Elements Retrieved 2006 08 30 In HTML with Java based interactive figures Pythagorean theorem Encyclopedia of Mathematics EMS Press 2001 1994 History topic Pythagoras s theorem in Babylonian mathematics Interactive links Interactive proof in Java of the Pythagorean theorem Another interactive proof in Java of the Pythagorean theorem Pythagorean theorem with interactive animation Animated non algebraic and user paced Pythagorean theorem Pythagorean theorem water demo on YouTube Pythagorean theorem more than 70 proofs from cut the knot Weisstein Eric W Pythagorean theorem MathWorld Retrieved from https en wikipedia org w index php title Pythagorean theorem amp oldid 1151388316, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.