You are on page 1of 63

Area of a circle

In geometry, the area enclosed by a


circle of radius r is πr2. Here the Greek
letter π represents the constant ratio of
the circumference of any circle to its
diameter, approximately equal to
3.14159.

One method of deriving this formula,


which originated with Archimedes,
involves viewing the circle as the limit of
a sequence of regular polygons with an
increasing number of sides. The area of
a regular polygon is half its perimeter
multiplied by the distance from its
center to its sides, and because the
sequence tends to a circle, the
corresponding formula–that the area is
half the circumference times the radius–
1
namely, A = 2 × 2πr × r, holds for a
circle.

Terminology
Although often referred to as the area of
a circle in informal contexts, strictly
speaking the term disk refers to the
interior region of the circle, while circle is
reserved for the boundary only, which is
a curve and covers no area itself.
Therefore, the area of a disk is the more
precise phrase for the area enclosed by
a circle.

History
Modern mathematics can obtain the
area using the methods of integral
calculus or its more sophisticated
offspring, real analysis. However, the
area of a disk was studied by the
Ancient Greeks. Eudoxus of Cnidus in
the fifth century B.C. had found that the
area of a disk is proportional to its
radius squared.[1] Archimedes used the
tools of Euclidean geometry to show
that the area inside a circle is equal to
that of a right triangle whose base has
the length of the circle's circumference
and whose height equals the circle's
radius in his book Measurement of a
Circle. The circumference is 2πr, and the
area of a triangle is half the base times
the height, yielding the area π r2 for the
disk. Prior to Archimedes, Hippocrates
of Chios was the first to show that the
area of a disk is proportional to the
square of its diameter, as part of his
quadrature of the lune of Hippocrates,[2]
but did not identify the constant of
proportionality.
Historical arguments
A variety of arguments have been
advanced historically to establish the
equation to varying degrees of
mathematical rigor. The most famous of
these is Archimedes' method of
exhaustion, one of the earliest uses of
the mathematical concept of a limit, as
well as the origin of Archimedes' axiom
which remains part of the standard
analytical treatment of the real number
system. The original proof of
Archimedes is not rigorous by modern
standards, because it assumes that we
can compare the length of arc of a circle
to the length of a secant and a tangent
line, and similar statements about the
area, as geometrically evident.

Using polygons

The area of a regular polygon is half its


perimeter times the apothem. As the
number of sides of the regular polygon
increases, the polygon tends to a circle,
and the apothem tends to the radius.
This suggests that the area of a disk is
half the circumference of its bounding
circle times the radius.[3]
Archimedes's proof

Following Archimedes' argument in The


Measurement of a Circle (c. 260 BCE),
compare the area enclosed by a circle to
a right triangle whose base has the
length of the circle's circumference and
whose height equals the circle's radius.
If the area of the circle is not equal to
that of the triangle, then it must be either
greater or less. We eliminate each of
these by contradiction, leaving equality
as the only possibility. We use regular
polygons in the same way.
Not greater

Circle with square and octagon


inscribed, showing area gap

Suppose that the area C enclosed by the


circle is greater than the area T = 1⁄2cr of
the triangle. Let E denote the excess
amount. Inscribe a square in the circle,
so that its four corners lie on the circle.
Between the square and the circle are
four segments. If the total area of those
gaps, G4, is greater than E, split each arc
in half. This makes the inscribed square
into an inscribed octagon, and produces
eight segments with a smaller total gap,
G8. Continue splitting until the total gap
area, Gn, is less than E. Now the area of
the inscribed polygon, Pn = C − Gn, must
be greater than that of the triangle.

But this forces a contradiction, as


follows. Draw a perpendicular from the
center to the midpoint of a side of the
polygon; its length, h, is less than the
circle radius. Also, let each side of the
polygon have length s; then the sum of
the sides, ns, is less than the circle
circumference. The polygon area
consists of n equal triangles with height
h and base s, thus equals 1⁄2nhs. But
since h < r and ns < c, the polygon area
must be less than the triangle area,
1⁄ cr, a contradiction. Therefore, our
2
supposition that C might be greater than
T must be wrong.

Not less

Circle with square and octagon


circumscribed, showing area gap
Suppose that the area enclosed by the
circle is less than the area T of the
triangle. Let D denote the deficit amount.
Circumscribe a square, so that the
midpoint of each edge lies on the circle.
If the total area gap between the square
and the circle, G4, is greater than D, slice
off the corners with circle tangents to
make a circumscribed octagon, and
continue slicing until the gap area is less
than D. The area of the polygon, Pn, must
be less than T.
This, too, forces a contradiction. For, a
perpendicular to the midpoint of each
polygon side is a radius, of length r. And
since the total side length is greater than
the circumference, the polygon consists
of n identical triangles with total area
greater than T. Again we have a
contradiction, so our supposition that C
might be less than T must be wrong as
well.
Therefore, it must be the case that the
area enclosed by the circle is precisely
the same as the area of the triangle.
This concludes the proof.

Rearrangement proof

Circle area by rearrangement


Graphs of side, s; apothem, a; and
area, A of regular polygons of n
sides and circumradius 1, with the
base, b of a rectangle with the
same area. The green line shows
the case n = 6.

Following Satō Moshun (Smith & Mikami


1914, pp. 130–132), Nicholas of Cusa[4]
and Leonardo da Vinci (Beckmann 1976,
p. 19), we can use inscribed regular
polygons in a different way. Suppose we
inscribe a hexagon. Cut the hexagon into
six triangles by splitting it from the
center. Two opposite triangles both
touch two common diameters; slide
them along one so the radial edges are
adjacent. They now form a
parallelogram, with the hexagon sides
making two opposite edges, one of
which is the base, s. Two radial edges
form slanted sides, and the height, h is
equal to its apothem (as in the
Archimedes proof). In fact, we can also
assemble all the triangles into one big
parallelogram by putting successive
pairs next to each other. The same is
true if we increase it to eight sides and
so on. For a polygon with 2n sides, the
parallelogram will have a base of length
ns, and a height h. As the number of
sides increases, the length of the
parallelogram base approaches half the
circle circumference, and its height
approaches the circle radius. In the limit,
the parallelogram becomes a rectangle
with width πr and height r.
Unit disk area by rearranging n polygons.

polygon parallelogram

n side base height area

4 1.4142136 2.8284271 0.7071068 2.0000000

6 1.0000000 3.0000000 0.8660254 2.5980762

8 0.7653669 3.0614675 0.9238795 2.8284271

10 0.6180340 3.0901699 0.9510565 2.9389263

12 0.5176381 3.1058285 0.9659258 3.0000000

14 0.4450419 3.1152931 0.9749279 3.0371862

16 0.3901806 3.1214452 0.9807853 3.0614675

96 0.0654382 3.1410320 0.9994646 3.1393502

∞ 1/∞ π 1 π

Modern proofs
There are various equivalent definitions
of the constant π. The conventional
definition in pre-calculus geometry is the
ratio of the circumference of a circle to
its diameter:

However, because the circumference of


a circle is not a primitive analytical
concept, this definition is not suitable in
modern rigorous treatments. A standard
modern definition is that π is equal to
twice the least positive root of the
cosine function or, equivalently, the half-
period of the sine (or cosine) function.
The cosine function can be defined
either as a power series, or as the
solution of a certain differential
equation. This avoids any reference to
circles in the definition of π, so that
statements about the relation of π to the
circumference and area of circles are
actually theorems, rather than
definitions, that follow from the
analytical definitions of concepts like
"area" and "circumference".

The analytical definitions are seen to be


equivalent, if it is agreed that the
circumference of the circle is measured
as a rectifiable curve by means of the
integral
The integral appearing on the right is an
abelian integral whose value is a half-
period of the sine function, equal to π.
Thus is seen to be
true as a theorem.

Several of the arguments that follow use


only concepts from elementary calculus
to reproduce the formula , but
in many cases to regard these as actual
proofs, they rely implicitly on the fact
that one can develop trigonometric
functions and the fundamental constant
π in a way that is totally independent of
their relation to geometry. We have
indicated where appropriate how each of
these proofs can be made totally
independent of all trigonometry, but in
some cases that requires more
sophisticated mathematical ideas than
those afforded by elementary calculus.

Onion proof

Area of the disk via ring


integration

Using calculus, we can sum the area


incrementally, partitioning the disk into
thin concentric rings like the layers of an
onion. This is the method of shell
integration in two dimensions. For an
infinitesimally thin ring of the "onion" of
radius t, the accumulated area is 2πt dt,
the circumferential length of the ring
times its infinitesimal width (one can
approximate this ring by a rectangle with
width=2πt and height=dt). This gives an
elementary integral for a disk of radius r.

It is rigorously justified by the


multivariate substitution rule in polar
coordinates. Namely, the area is given by
a double integral of the constant
function 1 over the disk itself. If D
denotes the disk, then the double
integral can be computed in polar
coordinates as follows:
which is the same result as obtained
above.

An equivalent rigorous justification,


without relying on the special
coordinates of trigonometry, uses the
coarea formula. Define a function
by .
Note ρ is a Lipschitz function whose
gradient is a unit vector
(almost everywhere). Let D be the disc
in . We will show that
, where is the two-
dimensional Lebesgue measure in .
We shall assume that the one-
dimensional Hausdorff measure of the
circle is , the circumference
of the circle of radius r. (This can be
taken as the definition of
circumference.) Then, by the coarea
formula,
Triangle proof

Circle unwrapped to form a


triangle

The circle and the triangle are


equal in area.

Similar to the onion proof outlined


above, we could exploit calculus in a
different way in order to arrive at the
formula for the area of a disk. Consider
unwrapping the concentric circles to
straight strips. This will form a right
angled triangle with r as its height and
2πr (being the outer slice of onion) as its
base.

Finding the area of this triangle will give


the area of the disk

The opposite and adjacent angles for


this triangle are respectively in degrees
9.0430611..., 80.956939... and in radians
0.1578311... OEIS: A233527,
1.4129651...OEIS: A233528.
Explicitly, we imagine dividing up a circle
into triangles, each with a height equal
to the circle's radius and a base that is
infinitesimally small. The area of each of
these triangles is equal to .
By summing up (integrating) all of the
areas of these triangles, we arrive at the
formula for the circle's area:

It too can be justified by a double


integral of the constant function 1 over
the disk by reversing the order of
integration and using a change of
variables in the above iterated integral:

Making the substitution


converts the
integral to
which is the same as the above result.

The triangle proof can be reformulated


as an application of Green's theorem in
flux-divergence form (i.e. a two-
dimensional version of the divergence
theorem), in a way that avoids all
mention of trigonometry and the
constant π. Consider the vector field
in the plane. So the
divergence of r is equal to two, and
hence the area of a disc D is equal to
By Green's theorem, this is the same as
the outward flux of r across the circle
bounding D:

where n is the unit normal and ds is the


arc length measure. For a circle of
radius R centered at the origin, we have
and , so the above
equality is

The integral of ds over the whole circle


is just the arc length, which is its
circumference, so this shows that the
area A enclosed by the circle is equal to
times the circumference of the
circle.

Another proof that uses triangles


considers the area enclosed by a circle
to be made up of an infinite number of
triangles (i.e. the triangles each have an
angle of d𝜃 at the centre of the circle),
each with an area of 12 · r2 · d𝜃 (derived
from the expression for the area of a
triangle:
1 1 1
2 · a · b · sin 𝜃 = 2 · r · r · sin(d𝜃) = 2 · r 2 · d𝜃

). Note that sin(d𝜃) ≈ d𝜃 due to small


angle approximation. Through summing
the areas of the triangles, the expression
for the area of the circle can therefore be
found:

Semicircle proof

Note that the area of a semicircle of


radius r can be computed by the integral
.

A semicircle of radius r
By trigonometric substitution, we
substitute , hence

The last step follows since the


trigonometric identity
implies that
and have equal integrals
over the interval , using
integration by substitution. But on the
other hand, since ,
the sum of the two integrals is the length
of that interval, which is .
Consequently, the integral of is
equal to half the length of that interval,
which is .

Therefore, the area of a circle of radius r,


which is twice the area of the semi-

circle, is equal to .

This particular proof may appear to beg


the question, if the sine and cosine
functions involved in the trigonometric
substitution are regarded as being
defined in relation to circles. However, as
noted earlier, it is possible to define sine,
cosine, and π in a way that is totally
independent of trigonometry, in which
case the proof is valid by the change of
variables formula and Fubini's theorem,
assuming the basic properties of sine
and cosine (which can also be proved
without assuming anything about their
relation to circles).

Isoperimetric inequality
The circle is the closed curve of least
perimeter that encloses the maximum
area. This is known as the isoperimetric
inequality, which states that if a
rectifiable Jordan curve in the Euclidean
plane has perimeter C and encloses an
area A (by the Jordan curve theorem)
then

Moreover, equality holds in this


inequality if and only if the curve is a
circle, in which case and
.

Fast approximation
The calculations Archimedes used to
approximate the area numerically were
laborious, and he stopped with a
polygon of 96 sides. A faster method
uses ideas of Willebrord Snell
(Cyclometricus, 1621), further developed
by Christiaan Huygens (De Circuli
Magnitudine Inventa, 1654), described in
Gerretsen & Verdenduin (1983, pp. 243–
250).

Archimedes' doubling method

Given a circle, let un be the perimeter of


an inscribed regular n-gon, and let Un be
the perimeter of a circumscribed regular
n-gon. Then un and Un are lower and
upper bounds for the circumference of
the circle that become sharper and
sharper as n increases, and their
average (un + Un)/2 is an especially good
approximation to the circumference. To
compute un and Un for large n,
Archimedes derived the following
doubling formulae:

(geometric mean),
and

(harmonic

mean).

Starting from a hexagon, Archimedes


doubled n four times to get a 96-gon,
which gave him a good approximation to
the circumference of the circle.

In modern notation, we can reproduce


his computation (and go further) as
follows. For a unit circle, an inscribed
hexagon has u6 = 6, and a circumscribed
hexagon has U6 = 4√3. Doubling seven
times yields
Archimedes doubling seven times; n = 6×2k.
un + Un
k n un Un
4

0 6 6.0000000 6.9282032 3.2320508

1 12 6.2116571 6.4307806 3.1606094

2 24 6.2652572 6.3193199 3.1461443

3 48 6.2787004 6.2921724 3.1427182

4 96 6.2820639 6.2854292 3.1418733

5 192 6.2829049 6.2837461 3.1416628

6 384 6.2831152 6.2833255 3.1416102

7 768 6.2831678 6.2832204 3.1415970

un + Un
(Here 2 approximates the
circumference of the unit circle, which is
un + Un
2π, so 4 approximates π.)

The last entry of the table has 355⁄113 as


one of its best rational approximations;
i.e., there is no better approximation
among rational numbers with
denominator up to 113. The number
355⁄ is also an excellent
113
approximation to π, attributed to
Chinese mathematician Zu Chongzhi,
who named it Milü.[5] This approximation
is better than any other rational number
with denominator less than 16,604.[6]

The Snell–Huygens refinement

Snell proposed (and Huygens proved) a


tighter bound than Archimedes':

This for n = 48 gives a better


approximation (about 3.14159292) than
Archimedes' method for n = 768.

Derivation of Archimedes' doubling


formulae

Circle with similar triangles:


circumscribed side, inscribed side
and complement, inscribed split
side and complement

Let one side of an inscribed regular n-


gon have length sn and touch the circle
at points A and B. Let A′ be the point
opposite A on the circle, so that A′A is a
diameter, and A′AB is an inscribed
triangle on a diameter. By Thales'
theorem, this is a right triangle with right
angle at B. Let the length of A′B be cn,
which we call the complement of sn;
thus cn2+sn2 = (2r)2. Let C bisect the arc
from A to B, and let C′ be the point
opposite C on the circle. Thus the length
of CA is s2n, the length of C′A is c2n, and
C′CA is itself a right triangle on diameter
C′C. Because C bisects the arc from A to
B, C′C perpendicularly bisects the chord
from A to B, say at P. Triangle C′AP is
thus a right triangle, and is similar to
C′CA since they share the angle at C′.
Thus all three corresponding sides are in
the same proportion; in particular, we
have C′A : C′C = C′P : C′A and AP : C′A =
CA : C′C. The center of the circle, O,
bisects A′A, so we also have triangle
OAP similar to A′AB, with OP half the
length of A′B. In terms of side lengths,
this gives us

In the first equation C′P is C′O+OP, length


r+1⁄2cn, and C′C is the diameter, 2r. For a
unit circle we have the famous doubling
equation of Ludolph van Ceulen,

If we now circumscribe a regular n-gon,


with side A″B″ parallel to AB, then OAB
and OA″B″ are similar triangles, with
A″B″ : AB = OC : OP. Call the
circumscribed side Sn; then this is
Sn : sn = 1 : 1⁄2cn. (We have again used
that OP is half the length of A′B.) Thus
we obtain

Call the inscribed perimeter un = nsn, and


the circumscribed perimeter Un = nSn.
Then combining equations, we have

so that
This gives a geometric mean equation.

We can also deduce

or

This gives a harmonic mean equation.

Dart approximation

Unit circle area Monte Carlo


integration. Estimate by these 900
709
samples is 4× 900 = 3.15111...
When more efficient methods of finding
areas are not available, we can resort to
"throwing darts". This Monte Carlo
method uses the fact that if random
samples are taken uniformly scattered
across the surface of a square in which
a disk resides, the proportion of samples
that hit the disk approximates the ratio
of the area of the disk to the area of the
square. This should be considered a
method of last resort for computing the
area of a disk (or any shape), as it
requires an enormous number of
samples to get useful accuracy; an
estimate good to 10−n requires about
100n random samples (Thijssen 2006,
p. 273).

Finite rearrangement
We have seen that by partitioning the
disk into an infinite number of pieces we
can reassemble the pieces into a
rectangle. A remarkable fact discovered
relatively recently (Laczkovich 1990) is
that we can dissect the disk into a large
but finite number of pieces and then
reassemble the pieces into a square of
equal area. This is called Tarski's circle-
squaring problem. The nature of
Laczkovich's proof is such that it proves
the existence of such a partition (in fact,
of many such partitions) but does not
exhibit any particular partition.

Non-Euclidean circles
Circles can be defined in non-Euclidean
geometry, and in particular in the
hyperbolic and elliptic planes.

For example, the unit sphere is a


model for the two-dimensional elliptic
plane. It carries an intrinsic metric that
arises by measuring geodesic length.
The geodesic circles are the parallels in
a geodesic coordinate system.

More precisely, fix a point


that we place at the zenith. Associated
to that zenith is a geodesic polar
coordinate system , ,
, where z is the point
. In these coordinates, the geodesic
distance from z to any other point
having coordinates is
the value of at x. A spherical circle is
the set of points a geodesic distance R
from the zenith point z. Equivalently, with
a fixed embedding into , the spherical
circle of radius centered at z is
the set of x in such that
.

We can also measure the area of the


spherical disk enclosed within a
spherical circle, using the intrinsic
surface area measure on the sphere.
The area of the disk of radius R is then
given by

More generally, if a sphere has


radius of curvature , then the area of
the disk of radius R is given by

Observe that, as an application of


L'Hôpital's rule, this tends to the
Euclidean area in the flat limit
.
The hyperbolic case is similar, with the
area of a disk of intrinsic radius R in the
(constant curvature ) hyperbolic
plane given by

where cosh is the hyperbolic cosine.


More generally, for the constant
curvature hyperbolic plane, the
answer is

These identities are important for


comparison inequalities in geometry. For
example, the area enclosed by a circle of
radius R in a flat space is always greater
than the area of a spherical circle and
smaller than a hyperbolic circle, provided
all three circles have the same (intrinsic)
radius. That is,

for all . Intuitively, this is because


the sphere tends to curve back on itself,
yielding circles of smaller area than
those in the plane, whilst the hyperbolic
plane, when immersed into space,
develops fringes that produce additional
area. It is more generally true that the
area of the circle of a fixed radius R is a
strictly decreasing function of the
curvature.
In all cases, if is the curvature
(constant, positive or negative), then the
isoperimetric inequality for a domain
with area A and perimeter L is

where equality is achieved precisely for


the circle.[7]

Generalizations
We can stretch a disk to form an ellipse.
Because this stretch is a linear
transformation of the plane, it has a
distortion factor which will change the
area but preserve ratios of areas. This
observation can be used to compute the
area of an arbitrary ellipse from the area
of a unit circle.

Consider the unit circle circumscribed by


a square of side length 2. The
transformation sends the circle to an
ellipse by stretching or shrinking the
horizontal and vertical diameters to the
major and minor axes of the ellipse. The
square gets sent to a rectangle
circumscribing the ellipse. The ratio of
the area of the circle to the square is
π/4, which means the ratio of the ellipse
to the rectangle is also π/4. Suppose a
and b are the lengths of the major and
minor axes of the ellipse. Since the area
of the rectangle is ab, the area of the
ellipse is πab/4.

We can also consider analogous


measurements in higher dimensions.
For example, we may wish to find the
volume inside a sphere. When we have a
formula for the surface area, we can use
the same kind of "onion" approach we
used for the disk.

See also
Area of a triangle

References
1. Stewart, James (2003). Single
variable calculus early
transcendentals (https://archive.org/
details/singlevariableca00stew/pag
e/3) (5th. ed.). Toronto ON:
Brook/Cole. pp. 3 (https://archive.or
g/details/singlevariableca00stew/pa
ge/3) . ISBN 0-534-39330-6.
"However, by indirect reasoning,
Eudoxus (fifth century B.C.) used
exhaustion to prove the familiar
formula for the area of a disk:
"

2. Heath, Thomas L. (2003), A Manual


of Greek Mathematics (https://book
s.google.com/books?id=_HZNr_mGF
zQC&pg=PA121) , Courier Dover
Publications, pp. 121–132, ISBN 0-
486-43231-9.

3. Hill, George. Lessons in Geometry:


For the Use of Beginners (https://arc
hive.org/details/lessonsingeomet04
hillgoog/page/n136) , page 124
(1894).

4. Clegg, Brian (2012). Introducing


Infinity. Icon Books. p. 69. ISBN 978-
1-84831-406-1.

5. Martzloff, Jean-Claude (2006). A


History of Chinese Mathematics (htt
ps://archive.org/details/historychine
sema00mart_058) . Springer. p. 281
(https://archive.org/details/historych
inesema00mart_058/page/n298) .
ISBN 9783540337829.

6. Not all best rational approximations


are the convergents of the continued
fraction! (http://shreevatsa.wordpres
s.com/2011/01/10/not-all-best-ratio
nal-approximations-are-the-converge
nts-of-the-continued-fraction/)

7. Isaac Chavel (2001), Isoperimetric


inequalities, Cambridge University
Press

Bibliography
Archimedes (1897), "Measurement of
a circle" (https://archive.org/stream/w
orksofarchimede029517mbp#page/n
279/mode/2up) , in Heath, T. L. (ed.),
The Works of Archimedes (https://ww
w.archive.org/details/worksofarchime
de029517mbp) , Cambridge University
Press
(Originally published by Cambridge
University Press, 1897, based on J. L.
Heiberg's Greek version.)
Beckmann, Petr (1976), A History of Pi,
St. Martin's Griffin, ISBN 978-0-312-
38185-1
Gerretsen, J.; Verdenduin, P. (1983),
"Chapter 8: Polygons and Polyhedra",
in H. Behnke; F. Bachmann; K. Fladt; H.
Kunle (eds.), Fundamentals of
Mathematics, Volume II: Geometry,
translated by S. H. Gould, MIT Press,
pp. 243–250, ISBN 978-0-262-52094-2
(Originally Grundzüge der Mathematik,
Vandenhoeck & Ruprecht, Göttingen,
1971.)
Laczkovich, Miklós (1990),
"Equidecomposability and
discrepancy: A solution to Tarski's
circle squaring problem" (http://dz-srv
1.sub.uni-goettingen.de/sub/digbib/lo
ader?ht=VIEW&did=D262326) ,
Journal für die reine und angewandte
Mathematik, 1990 (404): 77–117,
doi:10.1515/crll.1990.404.77 (https://
doi.org/10.1515%2Fcrll.1990.404.77) ,
MR 1037431 (https://mathscinet.ams.
org/mathscinet-getitem?mr=103743
1) , S2CID 117762563 (https://api.sem
anticscholar.org/CorpusID:11776256
3)
Lang, Serge (1985), "The length of the
circle", Math! : Encounters with High
School Students, Springer-Verlag,
ISBN 978-0-387-96129-3
Smith, David Eugene; Mikami, Yoshio
(1914), A history of Japanese
mathematics (https://archive.org/detai
ls/historyofjapanes00smituoft) ,
Chicago: Open Court Publishing,
pp. 130–132, ISBN 978-0-87548-170-8
Thijssen, J. M. (2006), Computational
Physics, Cambridge University Press,
p. 273, ISBN 978-0-521-57588-1

External links
Science News on Tarski problem (htt
p://www.sciencenews.org/articles/20
041030/mathtrek.asp) Archived (http
s://web.archive.org/web/2008041311
4409/http://www.sciencenews.org/art
icles/20041030/mathtrek.asp) 2008-
04-13 at the Wayback Machine

Retrieved from
"https://en.wikipedia.org/w/index.php?
title=Area_of_a_circle&oldid=1169811636"

This page was last edited on 11 August 2023, at


12:15 (UTC). •
Content is available under CC BY-SA 4.0 unless
otherwise noted.

You might also like