You are on page 1of 8

HIQUIAL, EMMANUEL T.

BS-CE 3

MECH223 5:30E-6:30 M-F

1. Brachistochrone Problem

Find the shape of the curve down which a bead sliding from rest and accelerated by gravity will slip
(without friction) from one point to another in the least time. The term derives from the Greek
(brachistos) "the shortest" and (chronos) "time, delay."

The brachistochrone problem was one of the earliest problems posed in the calculus of variations.
Newton was challenged to solve the problem in 1696, and did so the very next day (Boyer and
Merzbach 1991, p. 405). In fact, the solution, which is a segment of a cycloid, was found by Leibniz,
L'Hospital, Newton, and the two Bernoullis. Johann Bernoulli solved the problem using the analogous
one of considering the path of light refracted by transparent layers of varying density (Mach 1893,
Gardner 1984, Courant and Robbins 1996). Actually, Johann Bernoulli had originally found an
incorrect proof that the curve is a cycloid, and challenged his brother Jakob to find the required
curve. When Jakob correctly did so, Johann tried to substitute the proof for his own (Boyer and
Merzbach 1991, p. 417).

In the solution, the bead may actually travel uphill along the cycloid for a distance, but the path is
nonetheless faster than a straight line (or any other line).

The time to travel from a point to another point is given by the integral

where is the arc length and is the speed. The speed at any point is given by a simple application of
conservation of energy equating kinetic energy to gravitational potential energy,

giving

Plugging this into (◇) together with the identity

then gives

The function to be varied is thus

To proceed, one would normally have to apply the full-blown Euler-Lagrange differential equation

However, the function is particularly nice since does not appear explicitly. Therefore,
, and we can immediately use the Beltrami identity
Computing

subtracting from , and simplifying then gives

Squaring both sides and rearranging slightly results in

where the square of the old constant has been expressed in terms of a new (positive) constant .
This equation is solved by the parametric equations

which are--lo and behold--the equations of a cycloid.


If kinetic friction is included, the problem can also be solved analytically, although the solution is
significantly messier. In that case, terms corresponding to the normal component of weight and the
normal component of the acceleration (present because of path curvature) must be included.
Including both terms requires a constrained variational technique (Ashby et al. 1975), but including
the normal component of weight only gives an approximate solution. The tangent and normal
vectors are

gravity and friction are then

and the components along the curve are

so Newton's Second Law gives

But

so
Using the Euler-Lagrange differential equation gives

This can be reduced to

Now letting

the solution is
2. Dido's Problem

HISTORY

Dido was the legendary founder of Carthage (Tunisia). When she arrived in 814BC on the coast of
Tunisia, she asked for a piece of land. Her request was satisfied provided that the land could be
encompassed by an ox-hide. With a remarkable mathematical intuition, she cut the ox-hide into a
long thin strip and used it to encircle the land. This land became Carthage and Dido became the
Queen.

In mathematics, the isoperimetric inequality is a geometric inequality involving the square of the
circumference of a closed curve in the plane and the area of a plane region it encloses, as well as its
various generalizations. Isoperimetric literally means "having the same perimeter". Specifically, the
isoperimetric inequality states, for the length L of a closed curve and the area A of the planar region
that it encloses, that
4 π A ≤ L^2 and that equality holds if and only if the curve is a circle.

The isoperimetric problem is to determine a plane figure of the largest possible area whose
boundary has a specified length.[1] The closely related Dido's problem asks for a region of the
maximal area bounded by a straight line and a curvilinear arc whose endpoints belong to that line. It
is named after Dido, the legendary founder and first queen of Carthage. The solution to the
isoperimetric problem is given by a circle and was known already in Ancient Greece. However, the
first mathematically rigorous proof of this fact was obtained only in the 19th century. Since then,
many other proofs have been found.

TYPES OF PROBLEM

The isoperimetric problem in the plane

The classical isoperimetric problem dates back to antiquity. The problem can be stated as follows:
Among all closed curves in the plane of fixed perimeter, which curve (if any) maximizes the area of its
enclosed region? This question can be shown to be equivalent to the following problem: Among all
closed curves in the plane enclosing a fixed area, which curve (if any) minimizes the perimeter?

This problem is conceptually related to the principle of least action in physics, in that it can be
restated: what is the principle of action which encloses the greatest area, with the greatest economy
of effort? The 15th-century philosopher and scientist, Cardinal Nicholas of Cusa, considered
rotational action, the process by which a circle is generated, to be the most direct reflection, in the
realm of sensory impressions, of the process by which the universe is created. German astronomer
and astrologer Johannes Kepler invoked the isoperimetric principle in discussing the morphology of
the solar system, in Mysterium Cosmographicum (The Sacred Mystery of the Cosmos, 1596).
METHODS

I. The isoperimetric inequality

The solution to the isoperimetric problem is usually expressed in the form of an inequality that
relates the length L of a closed curve and the area A of the planar region that it encloses. The
isoperimetric inequality states that

and that the equality holds if and only if the curve is a circle. Indeed, the area of a disk of radius R is
πR2 and the circumference of the circle is 2πR, so both sides of the inequality are equal to 4π2R2 in
this case.

Dozens of proofs of the isoperimetric inequality have been found. In 1902, Hurwitz published a short
proof using the Fourier series that applies to arbitrary rectifiable curves (not assumed to be smooth).
An elegant direct proof based on comparison of a smooth simple closed curve with an appropriate
circle was given by E. Schmidt in 1938. It uses only the arc length formula, expression for the area of
a plane region from Green's theorem, and the Cauchy–Schwarz inequality.

For a given closed curve, the isoperimetric quotient is defined as the ratio of its area and that of the
circle having the same perimeter. This is equal to

and the isoperimetric inequality says that Q ≤ 1. Equivalently, the isoperimetric ratio L2/A is at least
4π for every curve.
The isoperimetric quotient of a regular n-gon is

Let C be a smooth regular convex closed curve. Then the improved isoperimetric inequality states the
following

where L , A , A ~ 0.5 denote the length of C, the area of the region bounded by C and the oriented
area of the Wigner caustic of C, respectively, and the equality holds if and only if is a curve of
constant width.

II. The isoperimetric inequality on the sphere

Let C be a simple closed curve on a sphere of radius 1. Denote by L the length of C and by A the area
enclosed by C. The spherical isoperimetric inequality states that

and that the equality holds if and only if the curve is a circle. There are, in fact, two ways to measure
the spherical area enclosed by a simple closed curve, but the inequality is symmetric with the
respect to taking the complement. This inequality was discovered by Paul Lévy (1919) who also
extended it to higher dimensions and general surfaces.
In the more general case of arbitrary radius R, it is known that
.
III. Isoperimetric inequality in higher dimensions

The isoperimetric theorem generalizes to surfaces in the three-dimensional Euclidean space. Among
all simple closed surfaces with given surface area, the sphere encloses a region of maximal volume.
An analogous statement holds in Euclidean spaces of any dimension.

In full generality (Federer 1969, §3.2.43), the isoperimetric inequality states that for any set S ⊂ Rn
whose closure has finite Lebesgue measure

where M*n-1 is the (n-1)-dimensional Minkowski content, Ln is the n-dimensional Lebesgue measure,
and ωn is the volume of the unit ball in Rn. If the boundary of S is rectifiable, then the Minkowski
content is the (n-1)-dimensional Hausdorff measure.

The isoperimetric inequality in n-dimensions can be quickly proven by the Brunn–Minkowski


inequality (Osserman (1978); Federer (1969, §3.2.43)).

The n-dimensional isoperimetric inequality is equivalent (for sufficiently smooth domains) to the
Sobolev inequality on Rn with optimal constant:

for all u ∈ W1,1(Rn).

IV. Isoperimetric inequalities in a metric measure space

Most of the work on isoperimetric problem has been done in the context of smooth regions in
Euclidean spaces, or more generally, in Riemannian manifolds. However, the isoperimetric problem
can be formulated in much greater generality, using the notion of Minkowski content. Let be
a metric measure space: X is a metric space with metric d, and μ is a Borel measure on X. The
boundary measure, or Minkowski content, of a measurable subset A of X is defined as the lim inf

where

is the ε-extension of A.
The isoperimetric problem in X asks how small can be for a given μ(A). If X is the Euclidean
plane with the usual distance and the Lebesgue measure then this question generalizes the classical
isoperimetric problem to planar regions whose boundary is not necessarily smooth, although the
answer turns out to be the same.
The function
is called the isoperimetric profile of the metric measure space . Isoperimetric profiles have
been studied for Cayley graphs of discrete groups and for special classes of Riemannian manifolds
(where usually only regions A with regular boundary are considered).

V. Isoperimetric inequality for triangles

The isoperimetric inequality for triangles in terms of perimeter p and area T states that

with equality for the equilateral triangle. This is implied, via the geometric-algebraic mean inequality,
by a stronger inequality which has also been called the isoperimetric inequality for triangles:

PROBLEM

What is the closed curve which has the maximum area for a given perimeter ?

The curve can be described by position and velocity . We need to compute


the length of the curve and the area of the enclosed surface.
The length is obtained by integration of the line element along the curve. The line element can be
intuitively computed as an infinitesimal Pythagorean formula or
.

The area is less intuitive to compute but it is simple if we use the Green’s theorem.
Green’s Theorem : Over a region in the plane with boundary we have

With and we get . Thus the enclosed area is given by


the elegant formula :

Now, the problem can be stated as an optimization problem with constraints :


Determine the closed curve that maximizes subject to the constraint

SOLUTION
There is a well established solution to this problem, namely an isoperimetric problem in the calculus
of variations.
Isoperimetric problem : Optimize subject to the constraint

To solve this problem, define the Lagrangian function

where is a constant called a Lagrange multiplier. Then the solutions to the problem can be shown to
satisfy the Euler-Lagrange equations :

The Lagrangian of the Dido problem is :

There are two Euler-Lagrange equations corresponding to x and y respectively :

Applying the differentiation :


Integrating both sides :

and

and
Squaring and summing :

We recognize the equation of a circle of center and radius . Since the imposed perimeter is p
we can deduce . Thus, the answer to the problem is a circle of radius and area .

Sources:

Weisstein, Eric W. "Brachistochrone Problem." From MathWorld--A Wolfram Web Resource.


http://mathworld.wolfram.com/BrachistochroneProblem.html
https://mathematicalgarden.wordpress.com/2008/12/21/the-problem-of-dido/
https://en.wikipedia.org/wiki/Isoperimetric_inequality

You might also like