Chapter 8 - Optimization For Engineering Systems - Ralph W. Pike
Chapter 8 - Optimization For Engineering Systems - Ralph W. Pike
CALCULUS OF VARIATIONS
INTRODUCTION
The calculus of variations and its extensions are devoted to finding the
optimum function that gives the best value of the economic model and
constraints of a system. The need for an optimum function, rather than an
optimal point, arises in numerous problems from a wide range of fields in
engineering and physics, which include optimal control, transport
phenomena, optics, elasticity, vibrations, statics and dynamics of solid
bodies, and navigation. Two examples are determining the optimal
temperatures profile in a catalytic reactor to maximize the conversion and
the optimal trajectory for a missile to maximize the satellite payload placed
in orbit. The first calculus of variations problem, the Brachistochrone
problem, was posed and solved by Johannes Bernoulli in 1696 (1). In this
problem the optimum curve was determined to minimize the time traveled
by a particle sliding without friction between two points.
This chapter is devoted to a relatively brief discussion of some of the key
concepts of this topic. These include the Euler equation and the Euler-
Poisson equations for the case of several functions and several independent
variables with and without constraints. It begins with a derivation of the
Euler equation and extends these concepts to more detailed cases. Examples
are given to illustrate this theory.
The purpose of this chapter is to develop an appreciation for what is
required to determine the optimum function for a variational problem. The
extensions and applications to optimal control, Pontryagin’s maximum
principle, and continuous dynamic programming are left to books devoted to
those topics.
𝒙𝟏
𝑰[𝒚(𝒙)] = ∫ [𝟏 + (𝒚′ )𝟐 ]𝟏⁄𝟐 (8-1)
𝒙𝟎
EULER EQUATION
In addition, the values of 𝑦(𝑥0 ) and 𝑦(𝑥1 ) are known, and an example of the
function 𝐹(𝑥, 𝑦, 𝑦 ′ ) was given in equation (8-1) as:
𝒙𝟏
̅(𝒙)] = ∫ 𝑭(𝒙, 𝒚
𝑰[𝒚 ̅′ )𝒅𝒙
̅, 𝒚 (8-6)
𝒙𝟎
The above equation can be put in terms of the optimal function 𝑦(𝑥) and the
arbitrary function 𝑛(𝑥) using equation (8-5).
𝒙𝟏
̅(𝒙)] = ∫ 𝑭(𝒙, 𝒚 + 𝜶𝒏, 𝒚′ + 𝜶𝒏′ )𝒅𝒙
𝑰[𝒚 (8-7)
𝒙𝟎
The mathematical argument (3) is made that all the possible functions 𝑦̅
lie in an arbitrarily small neighborhood of y, because 𝛼 can be made
arbitrarily small. As such, the integral of equation (8-7) may be regarded as
an ordinary function of 𝛼, Φ(𝛼), because 𝛼 would specify the value of the
integral knowing Φ(𝛼 = 0) at the minimum from 𝑦(𝑥).
𝒙𝟏
̅(𝒙)] = Φ(𝜶) = ∫ 𝑭(𝒙, 𝒚 + 𝜶𝒏, 𝒚′ + 𝜶𝒏′ )𝒅𝒙
𝑰[𝒚 (8-8)
𝒙𝟎
𝒅𝚽(𝜶) 𝒅 𝒙𝟏
= ∫ 𝑭(𝒙, 𝒚 ̅′ )𝒅𝒙
̅, 𝒚 (8-9)
𝒅𝜶 𝒅𝜶 𝒙𝟎
𝒅 𝜶𝟐 𝒅𝒇 𝒅𝜶𝟐 𝒅𝜶𝟏
∫ 𝒅𝒙 + 𝒇(𝜶𝟐 , 𝒕) − 𝒇(𝜶𝟏 , 𝒕) (8-10)
𝒅𝒕 𝜶𝟏 𝒅𝒕 𝒅𝒕 𝒅𝒕
𝒅𝒙𝟎 𝒅𝒙𝟏
= =𝟎 (8-11)
𝒅𝜶 𝒅𝜶
𝒙𝟏
𝒅𝚽(𝜶) 𝒅
=∫ ̅′ )𝒅𝒙
̅, 𝒚
𝑭(𝒙, 𝒚 (8-12)
𝒅𝜶 𝒙𝟎 𝒅𝜶
̅ 𝝏𝑭 𝒅𝒚
𝒅𝑭 𝝏𝑭 𝒅𝒚 ̅′ 𝝏𝑭 𝒅𝒙
= + ′ + (8-13)
̅ 𝒅𝜶 𝝏𝒚
𝒅𝜶 𝝏𝒚 ̅ 𝒅𝜶 𝝏𝒙 𝒅𝜶
𝒅𝜱 𝒙𝟏
̅ 𝝏𝑭 𝒅𝒚
𝝏𝑭 𝒅𝒚 ̅′
=∫ [ + ′ ] 𝒅𝒙 (8-14)
𝒅𝜶 𝒙𝟎 𝝏𝒚̅ 𝒅𝜶 𝝏𝒚̅ 𝒅𝜶
̅
𝒅𝒚 𝒅 ̅′
𝒅𝒚 𝒅 ′
= [𝒚 + 𝜶𝒏] = 𝒏 = [𝒚 + 𝜶𝒏′ ] = 𝒏′ (8-15)
𝒅𝜶 𝒅𝜶 𝒅𝜶 𝒅𝜶
𝒙𝟏 𝒙𝟏
𝝏𝑭 ′ 𝝏𝑭 𝒙 𝒅 𝝏𝑭 (8-17)
∫ ′
𝒏 𝒅𝒙 = ′ 𝒏|𝒙𝟏𝟎 − ∫ 𝒏 ( ′ ) 𝒅𝒙
𝒙𝟎 ̅
𝝏𝒚 ̅
𝝏𝒚 𝒙𝟎 ̅
𝒅𝒙 𝝏𝒚
The first term on the right-hand side is zero, for 𝑛(𝑥0 ) = 𝑛(𝑥1 ) = 0.
Combining the results from equation (8-17) with equation (8-16) gives:
𝒙𝟏
𝒅𝜱 𝝏𝑭 𝒅 𝝏𝑭
= ∫ 𝒏(𝒙) [ − ( ′ )] 𝒅𝒙
𝒅𝜶 𝒙𝟎 ̅ 𝒅𝒙 𝝏𝒚
𝒅𝒚 ̅
𝒙𝟏
∫ 𝒏(𝒙) 𝑮(𝒙)𝒅𝒙 = 𝟎
𝒙𝟎
for every choice of the continuously differentiable function 𝑛(𝑥) for which
𝑛(𝑥0 ) = 𝑛(𝑥1 ) = 0, then 𝐺(𝑥) = 0 identically in the interval 𝑥0 ≤ 𝑥 ≤ 𝑥1 .
𝝏𝑭 𝒅 𝝏𝑭
− ( )=𝟎 (8-4)
𝝏𝒚 𝒅𝒙 𝝏𝒚′
𝝏𝑭 𝝏 𝝏𝑭 𝝏 𝝏𝑭 𝝏 𝝏𝑭
𝒅 ( ′ ) = ′ ( ′ ) 𝒅𝒚′ + ( ′ ) 𝒅𝒚 + ( ) 𝒅𝒙 (8-19)
𝝏𝒚 𝝏𝒚 𝝏𝒚 𝝏𝒚 𝝏𝒚 𝝏𝒙 𝝏𝒚′
or
𝒅 𝝏𝑭 𝝏𝟐 𝑭 𝒅𝟐 𝒚 𝝏𝟐 𝑭 𝒅𝒚 𝝏𝟐 𝑭
( ′) = ′𝟐 𝟐 + + (8-20)
𝒅𝒙 𝝏𝒚 𝝏𝒚 𝒅𝒙 𝒅𝒚𝝏𝒚′ 𝒅𝒙 𝝏𝒙𝝏𝒚′
Substituting equation (8-20) into equation (8-4) and rearranging gives a more
familiar form for a second-order ordinary differential equation:
𝝏𝟐 𝑭 𝒅𝟐 𝒚 𝝏𝟐 𝑭 𝒅𝒚 𝝏𝟐 𝑭 𝝏𝑭
𝟐 𝒅𝒙𝟐
+ ′ 𝒅𝒙
+ ′
− =𝟎 (8-21)
𝝏𝒚 ′ 𝝏𝒚𝝏𝒚 𝝏𝒙𝝏𝒚 𝝏𝒚
𝒅𝟐 𝒚 𝒅𝒚
𝑭𝒚′ 𝒚′ + 𝑭𝒚′ 𝒚 + 𝑭𝒚′𝒙 − 𝑭𝒚 = 𝟎 (8-22)
𝒅𝒙𝟐 𝒅𝒙
𝒅 𝝏𝑭
(𝑭 − 𝒚′ ′ ) = 𝟎 (8-23)
𝒅𝒙 𝝏𝒚
This equation may be integrated once to obtain a form of the Euler equation
given below, which can be a more convenient starting point for problem
solving:
𝝏𝑭
𝑭 − 𝒚′ = 𝒄𝒐𝒏𝒔𝒕𝒂𝒏𝒕 (8-24)
𝝏𝒚′
Determine the function that gives the shortest distance between two given
points. Referring to Figure 8-2, we can state the problem as:
𝒙𝟏
𝑴𝒊𝒏𝒊𝒎𝒊𝒛𝒆: 𝑳 = ∫ 𝒅𝒔
𝒙𝟎
𝒙𝟏
𝑳 = ∫ [𝟏 + (𝒚′ )𝟐 ]𝟏⁄𝟐 𝒅𝒙
𝒙𝟎
Figure 8-2. Diagram to illustrate the shortest distance between two points
for Example 8-1.
Evaluating the partial derivatives for the Euler equation:
𝒅𝟐 𝒚 𝒅𝒚
(𝟏⁄[𝟏 + (𝒚′ )𝟐 ]𝟑⁄𝟐 ) 𝟐
+ (𝟎) + (𝟎) − (𝟎) = 𝟎
𝒅𝒙 𝒅𝒙
Simplifying gives
𝒅𝟐 𝒚
=𝟎
𝒅𝒙𝟐
𝒚 = 𝒄𝟏 𝒙 + 𝒄𝟐
This is the equation of a straight line, and the constants, 𝑐1 and 𝑐2 are
evaluated from the boundary conditions.
𝒙𝟏
𝑻 = ∫ {[𝟏 + (𝒚′ )𝟐 ]𝟏⁄𝟐 ⁄[𝟐𝒈(𝒚 − 𝒚𝟎 )]𝟏⁄𝟐 } 𝒅𝒙 (8-25)
𝒙𝟎
The details of the solution are given by Weinstock (3). The solution is the
equations for a cycloid.
The method of obtaining the Euler equation is used almost directly to
obtain the results for more detailed forms of the integrand of equation (8-2).
The next section extends the results for more complex problems.
MORE COMPLEX PROBLEMS
In the procedure used to obtain the Euler equation, first a function was
constructed to have the integral be a function of a single independent
variable, 𝛼𝑖 , and then the classical theory of maxima and minima was applied
to locate the stationary point. This same method is used for more complex
problems that include more functions, e.g., 𝑦1 , 𝑦2 , …, 𝑦𝑛 ; in higher order
derivatives, e.g., 𝑦, 𝑦 ′ , …, 𝑦 (𝑛) ; and more than one independent variable,
e.g., 𝑦(𝑥1 , 𝑥2 ). It is instructive to take these additional complications in steps.
First, the case will be considered for one function 𝑦 with higher order
derivatives, and then this will be followed by the case of several functions
with first derivatives, all for one independent variable. These results can then
be combined for the case of several functions with higher order derivatives.
The results will be a set of ordinary differential equations to be solved. Then
further elaboration on the same ideas for the case of more than one
independent variable will give a partial differential equation to be solved for
the optimal function. Finally, any number of functions of varying order of
derivatives with several independent variables will require that a set of partial
differential equations be solved for the optimal functions.
Functional with Higher Derivatives in the Integrand: For the case of the
integrand containing higher order derivatives, the integral has the following
form:
𝒙𝟏
𝑰[𝒚(𝒙)] = ∫ 𝑭[𝒙, 𝒚, 𝒚′ , … , 𝒚(𝒎) ] 𝒅𝒙 (8-26)
𝒙𝟎
In this case, boundary conditions will be required for 𝑦(𝑥0 ), 𝑦 ′ (𝑥0 ),…,
𝑦 (𝑚) (𝑥0 ), and 𝑦(𝑥1 ), 𝑦 ′ (𝑥1 ),…, 𝑦 (𝑚) (𝑥1 ).
The function constructed in equation (8-5) is used, and the integral of
equation (8-26) becomes:
𝒙𝟏
̅(𝒙)] = 𝜱(𝜶) = ∫ 𝑭[𝒙, 𝒚
𝑰[𝒚 ̅′ , 𝒚
̅, 𝒚 ̅(𝒎) ] 𝒅𝒙
̅′′ , … , 𝒚 (8-27)
𝒙𝟎
𝒙𝟏
𝒅𝜱 𝒅
=∫ ̅ , 𝒚̅ ′ , 𝒚̅ ′′ , … , 𝒚̅ (𝒎) ] 𝒅𝒙
𝑭[𝒙, 𝒚 (8-28)
𝒅𝜶 𝒙𝟎 𝒅𝜶
and using the chain rule, we can write the integrand as:
̅ 𝝏𝑭 𝒅𝒚
𝒅𝑭 𝝏𝑭 𝒅𝒚 ̅′ 𝝏𝑭 𝒅𝒚
̅′′ ̅(𝒎)
𝝏𝑭 𝒅𝒚
= + ′ + ′′ + ⋯ + (𝒎) (8-29)
̅ 𝒅𝜶 𝝏𝒚
𝒅𝜶 𝝏𝒚 ̅ 𝒅𝜶 𝝏𝒚̅ 𝒅𝜶 ̅
𝝏𝒚 𝒅𝜶
𝒅𝑭 𝝏𝑭 𝝏𝑭 𝝏𝑭 𝝏𝑭
= 𝒏 + ′ 𝒏′ + ′′ 𝒏′′ + ⋯ + (𝒎) 𝒏(𝒎) (8-30)
̅
𝒅𝜶 𝝏𝒚 ̅
𝝏𝒚 ̅
𝝏𝒚 ̅
𝝏𝒚
𝒙𝟏
𝒅𝜱
= ∫ [𝑭𝒚̅ 𝒏 + 𝑭𝒚̅′ 𝒏′ + 𝑭𝒚̅′′ 𝒏′′ + ⋯ + 𝑭𝒚̅(𝒎) 𝒏(𝒎) ] 𝒅𝒙 (8-31)
𝒅𝜶 𝒙𝟎
𝒙𝟏 𝒙𝟏
𝒅
∫ 𝑭𝒚̅′ 𝒏′ 𝒅𝒙 = − ∫ 𝒏 𝑭 ′ 𝒅𝒙 (8-32)
𝒙𝟎 𝒙𝟎 𝒅𝒙 𝒚̅
𝒙𝟏 𝒙𝟏
𝒅 𝒙𝟏
𝒅𝟐
∫ 𝑭𝒚̅′′ 𝒏′′ 𝒅𝒙 = − ∫ 𝒏′ 𝑭𝒚̅′′ 𝒅𝒙 = ∫ 𝒏 𝟐 𝑭𝒚̅′′ 𝒅𝒙
𝒙𝟎 𝒙𝟎 𝒅𝒙 𝒙𝟎 𝒅𝒙
⋮
𝒙𝟏 𝒙𝟏
𝒅(𝒎)
∫ 𝑭𝒚̅(𝒎) 𝒏(𝒎) 𝒅𝒙 = (−𝟏)𝒎 ∫ 𝒏 (𝒎) 𝑭𝒚̅(𝒎) 𝒅𝒙 (8-33)
𝒙𝟎 𝒙𝟎 𝒅𝒙
𝒅𝜱 𝒙𝟏 𝒅𝑭𝒚̅ 𝒅(𝒎)
= ∫ 𝒏 [𝑭𝒚̅ − + ⋯ + (−𝟏)𝒎 (𝒎) 𝑭𝒚̅(𝒎) ] 𝒅𝒙 (8-34)
𝒅𝜶 𝒙𝟎 𝒅𝒙 𝒅𝒙
𝒅𝑭𝒚 𝒅(𝒎)
𝑭𝒚 − + ⋯ + (−𝟏)𝒎 (𝒎) 𝑭𝒚(𝒎) = 𝟎 (8-35)
𝒅𝒙 𝒅𝒙
EXAMPLE 8-2(1)
Determine the optimum function that minimizes the integral in the following
equation:
𝒙𝟏
𝑰[𝒚(𝒙)] = ∫ [𝟏𝟔𝒚𝟐 − (𝒚′′ )𝟐 ] 𝒅𝒙
𝒙𝟎
𝒅𝟐 𝒅
𝑭𝒚′′ − 𝑭 ′ + 𝑭𝒚 = 𝟎
𝒅𝒙 𝟐 𝒅𝒙 𝒚
𝑭 = 𝟏𝟔𝒚𝟐 − (𝒚′′ )𝟐
𝝏𝑭 𝝏𝑭 𝝏𝑭
= 𝑭𝒚′′ = −𝟐𝒚′′ = 𝑭𝒚′ = 𝟎 = 𝑭𝒚 = 𝟑𝟐𝒚
𝝏𝒚′′ 𝝏𝒚′ 𝝏𝒚
𝒅𝟒 𝒚
− 𝟏𝟔𝒚 = 𝟎
𝒅𝒙𝟒
𝒙𝟏
𝑰[𝒚𝟏 (𝒙), 𝒚𝟐 (𝒙), … , 𝒚𝑷 (𝒙)] = ∫ 𝑭[𝒙, 𝒚𝟏 , 𝒚𝟐 , … , 𝒚𝑷 , 𝒚𝟏 ′ , 𝒚𝟐 ′ , … , 𝒚𝑷 ′ ] 𝒅𝒙
𝒙𝟐
(8-37)
and boundary conditions on each of the functions are required, i.e., 𝑦1 (𝑥0 ),
𝑦1 (𝑥1 ), 𝑦2 (𝑥0 ), 𝑦2 (𝑥1 ),…, 𝑦𝑃 (𝑥0 ), 𝑦𝑃 (𝑥1 ).
The function constructed in equation (8-5) is used, except in this case 𝑝
functions are required:
̅ 𝟏 = 𝒚𝟏 + 𝜶𝟏 𝒏𝟏
𝒚
⋮ (8-38)
̅ 𝑷 = 𝒚𝑷 + 𝜶𝑷 𝒏𝑷
𝒚
𝜱( 𝜶𝟏 , 𝜶𝟐 , … , 𝜶𝑷 )
𝒙𝟏
= ∫ 𝑭[𝒙, 𝒚𝟏 + 𝜶𝒏𝟏 , … , 𝒚𝑷 + 𝜶𝒏𝑷 , 𝒚𝟏 ′ + 𝜶𝒏𝟏 ′ , … , 𝒚𝑷 ′
𝒙𝟎
+ 𝜶𝒏𝑷 ′ ] 𝒅𝒙 (8-39)
To locate the stationary point(s) of the integral, the first partial derivatives of
Φ with respect to 𝛼1 , 𝛼2 , … , 𝛼𝑃 are set equal to zero. This gives the
following set of 𝑝 equations:
𝒙𝟏
𝝏𝜱 𝝏𝑭
=∫ 𝒅𝒙 𝒇𝒐𝒓 𝒊 = 𝟏, 𝟐, … , 𝒑 (8-40)
𝝏𝜶𝒊 𝒙𝟎 𝝏𝜶𝒊
𝝏𝑭 𝝏𝑭 𝝏𝒚̅𝒊 𝝏𝑭 𝝏𝒚̅𝒊 ′ 𝝏𝑭 𝝏𝑭 ′
= + = 𝒏𝒊 + 𝒏 (8-41)
𝝏𝜶𝒊 𝝏𝒚̅𝒊 𝝏𝜶𝒊 𝝏𝒚 ′
̅𝒊 𝝏𝜶𝒊 𝝏𝒚 ̅𝒊 ̅𝒊 ′ 𝒊
𝝏𝒚
𝒙𝟏
𝝏𝑭 𝒅 𝝏𝑭
∫ 𝒏𝒊 [ − ( )] 𝒅𝒙 𝒇𝒐𝒓 𝒊 = 𝟏, 𝟐, … , 𝒑 (8-43)
𝒙𝟎 𝝏𝒚𝒊 𝒅𝒙 𝝏𝒚𝒊 ′
𝝏𝑭 𝒅 𝝏𝑭
− ( ) = 𝟎 𝒇𝒐𝒓 𝒊 = 𝟏, 𝟐, … , 𝒑 (8-44)
𝝏𝒚𝒊 𝒅𝒙 𝝏𝒚𝒊 ′
This is a set of Euler equations, and the following example illustrates the use
of these equations to find the optimum set of functions.
EXAMPLE 8-3(1)
Determine the optimum functions that determine the stationary points for the
following integral:
𝒙𝟏
𝑰[𝒚𝟏 , 𝒚𝟐 ] = ∫ [𝟐𝒚𝟏 𝒚𝟐 − 𝟐𝒚𝟏 𝟐 + (𝒚𝟏 ′ )𝟐 − (𝒚𝟐 ′ )𝟐 ] 𝒅𝒙
𝒙𝟎
𝒅 𝝏𝑭 𝝏𝑭 𝒅 𝝏𝑭 𝝏𝑭
[ ]− =𝟎 [ ]− =𝟎
𝒅𝒙 𝝏𝒚𝟏 ′ 𝝏𝒚𝟏 𝒅𝒙 𝝏𝒚𝟐 ′ 𝝏𝒚𝟐
The function 𝐹 and the partial derivatives needed for these Euler equations
are:
𝝏𝑭 𝝏𝑭
= 𝟐𝒚𝟐 − 𝟒𝒚𝟏 = 𝟐𝒚𝟏
𝝏𝒚𝟏 𝝏𝒚𝟐
𝝏𝑭 𝝏𝑭
= 𝟐𝒚𝟏 ′ = −𝟐𝒚𝟐 ′
𝝏𝒚𝟏 ′ 𝝏𝒚𝟐 ′
−𝒚𝟏 ′′ + 𝒚𝟐 − 𝟐𝒚𝟏 = 𝟎 𝒚𝟐 ′′ + 𝒚𝟏 = 𝟎
This set of two linear ordinary differential equations has been solved by
Forray (1), using standard techniques, and the solution is:
𝒅 𝒅
(𝒎)
𝑭𝒚𝟏 (𝒎) − (𝒎−𝟏) 𝑭𝒚𝟏 (𝒎−𝟏) + ⋯ + (−𝟏)𝒎 𝑭𝒚𝟏 = 𝟎
𝒅𝒙 𝒅𝒙
⋮ (8-46)
(𝒌) (𝒌−𝟏)
𝒅 𝒅
𝑭𝒚 (𝒌) − (𝒌−𝟏) 𝑭𝒚𝑷 (𝒌−𝟏) + ⋯ + (−𝟏)𝒌 𝑭𝒚𝑷 = 𝟎
𝒅𝒙(𝒌) 𝑷 𝒅𝒙
This is a set of 𝑝 ordinary differential equations, and the order of each one
is determined by the highest order derivative appearing in the integrand. The
following example illustrates the use of these equations.
EXAMPLE 8-4
Determine the optimal functions that determine the stationary points for the
following integral:
𝒙𝟏
𝑰[𝒚𝟏 , 𝒚𝟐 ] = ∫ {[𝟏 + (𝒚𝟏 ′ )𝟐 ]𝟏⁄𝟐 + 𝟏𝟔𝒚𝟐 𝟐 − (𝒚𝟐 ′′ )𝟐 } 𝒅𝒙
𝒙𝟎
𝒅 𝒅𝟐 𝒅𝑭𝒚𝟐 ′
𝑭 ′ − 𝑭𝒚𝟏 = 𝟎 𝑭 ′′
− + 𝑭𝒚𝟐 = 𝟎
𝒅𝒙 𝒚𝟏 𝒚
𝒅𝒙𝟐 𝟐 𝒅𝒙
Computing the partial derivatives gives:
𝑭 𝒚𝟏 = 𝟎 𝑭𝒚𝟐′ = 𝟎
𝑭𝒚𝟐 = 𝟑𝟐𝒚𝟐
substituting gives
𝒅 𝒚𝟏 ′ 𝒅
[ ]=𝟎 (−𝟐𝒚𝟐 ′′ ) + 𝟑𝟐𝒚𝟐 = 𝟎
𝒅𝒙 [𝟏 + (𝒚𝟏 ′ )𝟐 ]𝟑⁄𝟐 𝒅𝒙𝟐
Functional with More than One Independent Variable: For this case the
integrand contains more than one independent variable. The analogous form
to equation (8-2) for two independent variables is:
where the integral is integrated over the region 𝑅, and 𝑦𝑥1 and 𝑦𝑥2 indicate
partial differentiation of 𝑦 with respect to 𝑥1 and 𝑥2 .
The procedure to obtain the differential equation to be solved for the
optimal solution of equation (8-47) follows the mathematical arguments used
for the case of one independent variable. However, Green's theorem is
required for the integration by parts, and the function 𝑛(𝑥1 , 𝑥2 ) is zero on the
surface
of the region 𝑅. The function 𝑦̅(𝑥1 , 𝑥2 ) is constructed from the optimal
function 𝑦(𝑥1 , 𝑥2 ) and the arbitrary function 𝑛(𝑥1 , 𝑥2 ) as:
𝒅𝜱 𝒅
=∫ ∫ 𝑭(𝒙𝟏 , 𝒙𝟐 , 𝒚, 𝒚𝒙𝟏 , 𝒚𝒙𝟐 ) 𝒅𝒙𝟏 𝒅𝒙𝟐 (8-50)
𝒅𝜶 𝑹 𝒅𝜶
Applying the chain rule, as was done previously where 𝐹 is not considered a
function of 𝑥1 and 𝑥2 for changes from surface to surface, the integrand
becomes:
̅
𝒅𝑭 𝝏𝑭 𝝏𝒚 ̅𝒙𝟏
𝝏𝑭 𝝏𝒚 ̅𝒙𝟐
𝝏𝑭 𝝏𝒚
= + + (8-51)
̅ 𝝏𝜶 𝝏𝒚
𝒅𝜶 𝝏𝒚 ̅𝒙𝟏 𝝏𝜶 ̅𝒙𝟐 𝝏𝜶
𝝏𝒚
and
̅
𝝏𝒚 ̅𝒙𝟏
𝝏𝒚 𝝏𝒏 ̅𝒙𝟐
𝝏𝒚 𝝏𝒏
=𝒏 = =
𝝏𝜶 𝝏𝜶 𝝏𝒙𝟏 𝝏𝜶 𝝏𝒙𝟐
The integral in equation (8-50) can be written in the following form, using
equation (8-51):
𝒅𝜱 𝝏𝑭 𝝏𝑭 𝝏𝒏 𝝏𝑭 𝝏𝒏
= ∫ ∫[ 𝒏 + + ] 𝒅𝒙𝟏 𝒅𝒙𝟐 (8-52)
𝒅𝜶 𝑹 ̅
𝝏𝒚 ̅
𝝏𝒚 𝒙𝟏 𝝏𝒙 𝟏 ̅𝒙𝟐 𝝏𝒙𝟐
𝝏𝒚
𝝏𝒇 𝝏𝒇
∫ ∫ [𝑮 +𝑯 ] 𝒅𝒙𝟏 𝒅𝒙𝟐 (8-53)
𝑫 𝝏𝒙𝟏 𝝏𝒙𝟐
𝝏𝑮 𝝏𝑯
= −∫ ∫𝒇[ + ] 𝒅𝒙𝟏 𝒅𝒙𝟐 + ∫ 𝒇(𝑮𝒅𝒙𝟏 − 𝑯𝒅𝒙𝟐 )
𝑫 𝝏𝒙𝟏 𝝏𝒙𝟐 𝑪
This theorem is applied to the second two terms of equation (8-52), where
𝑓 = 𝑛, 𝐺 = 𝜕𝐹 ⁄𝜕𝑦𝑥1 and 𝐻 = 𝜕𝐹 ⁄𝜕𝑦𝑥2 . An equation comparable to
equation (8-18) is obtained by allowing 𝛼 → 0, such that 𝑦̅ → 𝑦, 𝑦̅𝑥1 → 𝑦𝑥1
and 𝑦̅𝑥2 → 𝑦𝑥2 ; and 𝑑𝛷⁄𝑑𝛼 = 0 to have:
𝝏𝑭 𝝏 𝝏𝑭 𝝏 𝝏𝑭
∫ ∫𝒏[ − ( )− ( )] 𝒅𝒙𝟏 𝒅𝒙𝟐 = 𝟎 (8-54)
𝑹 𝝏𝒚 𝝏𝒙𝟏 𝝏𝒚𝒙𝟏 𝝏𝒙𝟐 𝝏𝒚𝒙𝟐
𝝏𝑭 𝝏 𝝏𝑭 𝝏 𝝏𝑭
− ( )− ( )=𝟎 (8-55)
𝝏𝒚 𝝏𝒙𝟏 𝝏𝒚𝒙𝟏 𝝏𝒙𝟐 𝝏𝒚𝒙𝟐
Also, this equation can be expanded using the chain rule to give an equation
that corresponds to equation (8-21), which is:
𝝏𝟐 𝒚 𝝏𝟐 𝒚 𝝏𝟐 𝒚 𝝏𝒚 𝝏𝒚
𝑭𝒚𝒙 𝒚 + 𝟐𝑭 𝒚 𝒙𝟏 𝒚 𝒙𝟐 + 𝑭𝒚 𝒙𝟐 𝒚 𝒙𝟐 + 𝑭𝒚𝒙 𝒚 + 𝑭𝒚𝒙 𝒚
𝟏 𝒙𝟏 𝝏𝒙 𝟐 𝝏𝒙𝟏 𝝏𝒙𝟐 𝝏𝒙𝟐 𝟐 𝟏 𝝏𝒙𝟏 𝟐 𝝏𝒙𝟐
𝟏
+ 𝑭𝒚𝒙 𝒙𝟏 + 𝑭𝒚𝒙 𝒙𝟐 − 𝑭𝒚 = 𝟎 (8-56)
𝟏 𝟐
The derivation of this equation follows the one for two independent
variables.
The following example illustrates an application of equation (8-55). Other
applications are given by Forray (1) and Schechter (4).
𝝏𝒚 𝟐 𝝏𝒚 𝟐
𝑰 = 𝟏⁄𝟐 ∫ ∫ {𝑨 ( ) + 𝑨( ) − 𝟐𝑩𝒚} 𝒅𝒙𝟏 𝒅𝒙𝟐
𝑫 𝝏𝒙𝟏 𝝏𝒙𝟐
Obtain the differential equation that is to be solved for the optimum shape.
The extension of the Euler equation for this case of two independent
variables was given by equation (8-55):
𝝏 𝝏𝑭 𝝏 𝝏𝑭 𝝏𝑭
( )+ ( )− =𝟎
𝝏𝒙𝟏 𝝏𝒚𝒙𝟏 𝝏𝒙𝟐 𝝏𝒚𝒙𝟐 𝝏𝒚
𝝏𝒚 𝟐 𝝏𝒚 𝟐
𝑭 = 𝑨( ) + 𝑨( ) − 𝟐𝑩𝒚
𝝏𝒙𝟏 𝝏𝒙𝟐
The following results are obtained from evaluating the partial derivatives:
𝝏𝑭 𝝏𝒚 𝝏𝑭 𝝏𝒚 𝝏𝑭
= 𝟐𝑨 = 𝟐𝑨 = −𝟐𝑩
𝝏𝒚𝒙𝟏 𝝏𝒙𝟏 𝝏𝒚𝒙𝟐 𝝏𝒙𝟐 𝝏𝒚
𝝏𝟐 𝒚 𝝏𝟐 𝒚 𝑩
𝟐
+ + =𝟎
𝝏𝒙𝟏 𝝏𝒙𝟐 𝟐 𝑨
Generally, there are two procedures used for solving variational problems
that have constraints. These are the methods of direct substitution and
Lagrange Multipliers. In the method of direct substitution, the constraint
equation is substituted into the integrand; and the problem is converted into
an unconstrained problem, as was done in Chapter 2. In the method of
Lagrange Multipliers, the Lagrangian function is formed, and the
unconstrained problem is solved using the appropriate forms of the Euler or
Euler-Poisson equation. However, in some cases the Lagrange Multiplier is
a function of the independent variables and is not a constant. This is an added
complication that was not encountered in Chapter 2.
𝒅 𝝏𝑳 𝝏𝑳
( )− =𝟎 (8-60)
𝒅𝒙 𝝏𝒚′ 𝝏𝒚
The classic example to illustrate this procedure is the problem of finding the
path of a unit mass particle on a sphere from point (0,0,1) to point (0,0, −1)
in time 𝑇, which minimizes the integral of the kinetic energy of the particle.
The integral to be minimized and the constraint to be satisfied are:
𝑻
𝑴𝒊𝒏𝒊𝒎𝒊𝒛𝒆: 𝑰[𝒙, 𝒚, 𝒛] = ∫ [(𝒙′ )𝟐 + (𝒚′ )𝟐 + (𝒛′ )𝟐 ]𝟏⁄𝟐 𝒅𝒕
𝟎
𝑺𝒖𝒃𝒋𝒆𝒄𝒕 𝒕𝒐: 𝒙𝟐 + 𝒚𝟐 + 𝒛𝟐 = 𝟏
𝒅 𝝏𝑳 𝝏𝑳 𝒅 𝝏𝑳 𝝏𝑳 𝒅 𝝏𝑳 𝝏𝑳
( ′) − =𝟎 ( ′) − =𝟎 ( ′) − =𝟎
𝒅𝒕 𝝏𝒙 𝝏𝒙 𝒅𝒕 𝝏𝒚 𝝏𝒚 𝒅𝒕 𝝏𝒛 𝝏𝒛
𝒅𝟐 𝒙 𝒅𝟐 𝒚 𝒅𝟐 𝒛
+ 𝝀𝒙 = 𝟎 + 𝝀𝒚 = 𝟎 + 𝝀𝒛 = 𝟎
𝒅𝒕𝟐 𝒅𝒕𝟐 𝒅𝒕𝟐
𝒙 + 𝒄𝟏 𝒚 + 𝒄𝟐 𝒛 = 𝟎
which is the equation of a plane through the center of the sphere. The
intersection of this plane and the sphere is a great circle, which is the optimal
path. It can be shown that the minimum kinetic energy is 𝜋 2 ⁄𝑇.
Integral Constraints: Isoperimetric problems (1) are ones where an integral
is to be optimized subject to a constraint, which is another integral having a
specified value. This name came from the famous problem of Dido of finding
the closed curve of given perimeter for which the area is a maximum. For the
Euler equation the problem can be stated as:
𝒙𝟏
𝑶𝒑𝒕𝒊𝒎𝒊𝒛𝒆: 𝑰[𝒚(𝒙)] = ∫ 𝑭(𝒙, 𝒚, 𝒚′ ) 𝒅𝒙
𝒙𝟎
𝒙𝟏
(8-61)
′)
𝑺𝒖𝒃𝒋𝒆𝒄𝒕 𝒕𝒐: 𝑱 = ∫ 𝑮(𝒙, 𝒚, 𝒚 𝒅𝒙
𝒙𝟎
and the following unconstrained Euler equation is solved along with the
constraint equation:
𝒅 𝝏𝑳 𝝏𝑳
( ′) − =𝟎 (8-60)
𝒅𝒙 𝝏𝒚 𝝏𝒚
Determine the shape of the curve of length 𝐽 that encloses the maximum area.
The integral to be maximized and the integral constraint are as follows:
𝒙𝟏
𝑴𝒂𝒙𝒊𝒎𝒊𝒛𝒆: 𝑰[𝒚(𝒙)] = 𝟏⁄𝟐 ∫ [𝒚𝟏 𝒚𝟐 ′ − 𝒚𝟐 𝒚𝟏 ′ ]𝒅𝒙
𝒙𝟎
𝒙𝟏
𝟐 𝟐 𝟏⁄𝟐
𝑺𝒖𝒃𝒋𝒆𝒄𝒕 𝒕𝒐: 𝑱 = ∫ [𝒚𝟏 ′ + 𝒚𝟐 ′ ] 𝒅𝒙
𝒙𝟎
The Lagrangian function is:
𝟐 𝟐 𝟏⁄𝟐
𝑳 = 𝒚𝟏 𝒚𝟐 ′ − 𝒚𝟐 𝒚𝟏 ′ + 𝝀[𝒚𝟏 ′ + 𝒚𝟐 ′ ]
𝒅 𝝏𝑳 𝝏𝑳 𝒅 𝝏𝑳 𝝏𝑳
( ′
)− =𝟎 ( ′
)− =𝟎
𝒅𝒙 𝝏𝒚𝟏 𝝏𝒚𝟏 𝒅𝒙 𝝏𝒚𝟐 𝝏𝒚𝟐
Performing the differentiation and substituting into the Euler equations give:
𝒅 𝟐 𝟐 −𝟏⁄𝟐
{−𝒚𝟐 + 𝒚𝟏 ′ 𝝀[𝒚𝟏 ′ + 𝒚𝟐 ′ ] } − 𝒚𝟐 ′ = 𝟎
𝒅𝒙
𝒅 𝟐 𝟐 −𝟏⁄𝟐
{𝒚 + 𝒚𝟐 ′ 𝝀[𝒚𝟏 ′ + 𝒚𝟐 ′ ] } − 𝒚𝟏 ′ = 𝟎
𝒅𝒙 𝟏
The two equations above can be integrated once to obtain the following
results:
−𝒚𝟏 ′ 𝝀 −𝒚𝟐 ′ 𝝀
𝒚 𝟐 − 𝒄𝟐 = 𝟏⁄𝟐
𝒚 𝟏 − 𝒄𝟏 = 𝟏⁄𝟐
𝟐[𝒚𝟏 ′ 𝟐 + 𝒚𝟐 ′ 𝟐 ] 𝟐[𝒚𝟏 ′ 𝟐 + 𝒚𝟐 ′ 𝟐 ]
Squaring both sides and adding the two equations gives the following:
which is the equation of a circle. Thus, a circle encloses the maximum area
for a given length curve.
𝒅 𝝏𝑳 𝝏𝑳
( ′) − =𝟎 (8-60)
𝒅𝒙 𝝏𝒚 𝝏𝒚
subject to:
𝒅𝒚𝟏
= 𝒚 𝟐 − 𝒚𝟏
𝒅𝒙
𝑳 = 𝒚𝟏 𝟐 + 𝒚𝟐 𝟐 + 𝝀(𝒚𝟏 ′ − 𝒚𝟐 + 𝒚𝟏 )
Using equation (8-60) obtains the two Euler equations for 𝑦1 and 𝑦2 . They
are to be solved with the constraint equation, and this gives the following set
of equations:
−𝟐𝒚𝟏 − 𝝀 + 𝝀′ = 𝟎
𝟐𝒚𝟐 − 𝝀 = 𝟎
𝒚𝟏 ′ + 𝒚𝟏 − 𝒚𝟐 = 𝟎
The solutions for 𝑦1 and 𝑦2 are obtained by manipulating and integrating the
equation set to give:
𝒚𝟏 = 𝒄𝟏 𝒆√𝟐𝒙 + 𝒄𝟐 𝒆√𝟐𝒙
𝒚𝟐 = 𝒄𝟏 (𝟏 + √𝟐)𝒆√𝟐𝒙 + 𝒄𝟐 (𝟏 − √𝟐)𝒆−√𝟐𝒙
where the constants of integration 𝑐1 and 𝑐2 are evaluated using the boundary
conditions. A particular solution for 𝑦1 (0) = 1 and 𝑦2 (𝑥1 ) = 0 is given by
Beveridge and Schechter (10).
√𝟐
𝑰[𝑻𝟏 (𝒕), 𝑻𝟐 (𝒕)] = ∫ {[𝑻𝟏 (𝒕) − 𝟑𝟐]𝟐 + [𝑻𝟐 (𝒕) − 𝟑𝟐]𝟐 } 𝒅𝒕
𝟎
An unsteady-state energy balance on the water in the tank at time 𝑡 gives the
following equation relating the temperatures 𝑇1 (𝑡), 𝑇2 (𝑡) and the system
parameters:
𝒅
𝑪𝑷 𝑾 𝑻 (𝒕) = 𝑾𝑪𝑷 [𝑻𝟏 (𝒕) − 𝟑𝟐] − 𝑾𝑪𝑷 [𝑻𝟐 (𝒕) − 𝟑𝟐]
𝒅𝒕 𝟐
For water the heat capacity, 𝐶𝑃 , is equal to 1.0 𝐵𝑇𝑈⁄𝑙𝑏℉, and this equation
simplifies to the following form:
𝒅
𝑻 (𝒕) = 𝑻𝟏 (𝒕) − 𝑻𝟐 (𝒕)
𝒅𝒕 𝟐
√𝟐
𝑴𝒊𝒏𝒊𝒎𝒊𝒛𝒆: 𝑰[𝑻𝟏 (𝒕), 𝑻𝟐 (𝒕)] = ∫ {[𝑻𝟏 (𝒕) − 𝟑𝟐]𝟐 + [𝑻𝟐 (𝒕) − 𝟑𝟐]𝟐 } 𝒅𝒕
𝟎
𝒅
𝑺𝒖𝒃𝒋𝒆𝒄𝒕 𝒕𝒐: 𝑻 (𝒕) − 𝑻𝟏 (𝒕) + 𝑻𝟐 (𝒕) = 𝟎
𝒅𝒕 𝟐
𝒅 𝝏𝑳 𝝏𝑳 𝒅 𝝏𝑳 𝝏𝑳
( )− =𝟎 ( )− =𝟎
𝒅𝒕 𝝏𝑻𝟏 ′ 𝝏𝑻𝟏 𝒅𝒕 𝝏𝑻𝟐 ′ 𝝏𝑻𝟐
𝝏𝑳 𝝏𝑳
= 𝟐(𝑻𝟏 − 𝟑𝟐) − 𝝀 = 𝟐(𝑻𝟐 − 𝟑𝟐) + 𝝀
𝝏𝑻𝟏 𝝏𝑻𝟐
𝝏𝑳 𝝏𝑳 𝒅 𝝏𝑻 𝒅 𝝏𝑻 𝒅𝝀
=𝟎 = 𝝀(𝒕) ( )=𝟎 ( ′) =
𝝏𝑻𝟏 ′ 𝝏𝑻𝟐 ′ 𝒅𝒕 𝝏𝑻𝟏 ′ 𝒅𝒕 𝝏𝑻𝟐 𝒅𝒕
𝒅 𝝏𝑳 𝒅 𝝏𝑳 𝒅𝝀
( )=𝟎 ( )=
𝒅𝒕 𝝏𝑻𝟏 ′ 𝒅𝒕 𝝏𝑻𝟐 ′ 𝒅𝒕
Substituting into the Euler equations gives the following set of equations:
𝟐(𝑻𝟏 − 𝟑𝟐) − 𝝀 = 𝟎
𝒅𝝀
𝟐(𝑻𝟐 − 𝟑𝟐) + 𝝀 − =𝟎
𝒅𝒕
𝒅
𝑻 − 𝑻𝟏 + 𝑻𝟐 = 𝟎
𝒅𝒕 𝟐
The third equation is the constraint, and these equations are solved for 𝑇1 (𝑡)
and 𝑇2 (𝑡). The set has two ordinary differential equations and one algebraic
equation. Manipulating and solving this set for one equation in terms of 𝑇2 (𝑡)
gives:
𝑻𝟐 ′′ − 𝟐𝑻𝟐 = −𝟔𝟒
The solutions for the optimal functions, 𝑇1 (𝑡) and 𝑇2 (𝑡) are tabulated and
plotted in Figure 8-3. As shown in the figure, the warm water temperature
increases to 209 °𝐹 for the water temperature in the tank to reach 104 °𝐹 in
√2 ℎ𝑜𝑢𝑟𝑠.
CLOSURE
2. Optimize:
𝒙𝟏
𝑰[𝒚(𝒙)] = ∫ 𝑭[𝒙, 𝒚, 𝒚′ , … , 𝒚(𝒎) ] 𝒅𝒙
𝒙𝟎
Solve an ordinary differential equation of order 2𝑚:
𝒅(𝒎) 𝝏𝑭 𝒅(𝒎−𝟏) 𝝏𝑭 𝝏𝑭
( ) − (𝒎−𝟏) ( (𝒎−𝟏) ) + ⋯ + (−𝟏)𝒎 =𝟎
𝒅𝒙(𝒎) 𝝏𝒚(𝒎) 𝒅𝒙 𝝏𝒚 𝝏𝒚
3. Optimize:
𝒙𝟏
𝑰[𝒚𝟏 , 𝒚𝟐 , … , 𝒚𝑷 ] = ∫ 𝑭[𝒙, 𝒚𝟏 , 𝒚𝟏 ′ , … , 𝒚𝟏 (𝒎) , … , 𝒚𝑷 , 𝒚𝒑 ′ , … , 𝒚𝑷 (𝒌) ] 𝒅𝒙
𝒙𝟎
Solve a set of 𝑝 Euler-Poisson equations:
𝒅(𝒎) 𝝏𝑭 𝒅(𝒎−𝟏) 𝝏𝑭 𝝏𝑭
(𝒎) ( (𝒎) ) − ( ) + ⋯ + (−𝟏)𝒎 =𝟎
𝒅𝒙 𝝏𝒚𝟏 𝒅𝒙(𝒎−𝟏) 𝝏𝒚𝟏 (𝒎−𝟏) 𝝏𝒚𝟏
⋮
𝒅(𝒌) 𝝏𝑭 𝒅(𝒌−𝟏) 𝝏𝑭 𝝏𝑭
( ) − (𝒌−𝟏) ( (𝒌−𝟏) )
+ ⋯ + (−𝟏)𝒌 =𝟎
𝒅𝒙(𝒌) 𝝏𝒚𝑷 (𝒌) 𝒅𝒙 𝝏𝒚𝑷 𝝏𝒚𝑷
4. Optimize:
𝑰[𝒚(𝒙𝟏 , 𝒙𝟐 )] = ∫ ∫ 𝑭(𝒙𝟏 , 𝒙𝟐 , 𝒚, 𝒚𝒙𝟏 , 𝒚𝒙𝟐 ) 𝒅𝒙𝟏 𝒅𝒙𝟐
𝑹
Solve a second-order partial differential equation:
𝝏 𝝏𝑭 𝝏 𝝏𝑭 𝝏𝑭
( )+ ( )− =𝟎
𝝏𝒙𝟏 𝝏𝒚𝒙𝟏 𝝏𝒙𝟐 𝝏𝒚𝒙𝟐 𝝏𝒚
5. Optimize:
𝒙𝟏
𝑰[𝒚(𝒙)] = ∫ 𝑭[𝒙, 𝒚, 𝒚′ ] 𝒅𝒙
𝒙𝟎
Subject to:
𝒙𝟏
𝑮(𝒙, 𝒚) = 𝟎 𝑱 = ∫ 𝑮(𝒙, 𝒚, 𝒚′ ) 𝒅𝒙 𝑮(𝒙, 𝒚, 𝒚′ ) = 𝟎
𝒙𝟎
𝑐𝑜𝑛𝑠𝑡𝑟𝑎𝑖𝑛𝑡𝑠: 𝑎𝑙𝑔𝑒𝑏𝑟𝑎𝑖𝑐 𝑖𝑛𝑡𝑒𝑔𝑟𝑎𝑙 𝑑𝑖𝑓𝑓𝑒𝑟𝑒𝑛𝑡𝑖𝑎𝑙 𝑒𝑞𝑢𝑎𝑡𝑖𝑜𝑛
Form the Lagrangian function 𝐿 = 𝐹 + λ𝐺 and solve the Euler equation with the
constraint equation. The Lagrange Multiplier is a constant for integral constraints
and is a function of the independent variable for algebraic and differential
equation constraints.
__________________________________________________________________________
to the Euler-Poisson equation with higher order derivatives are available in
Weinstock (3).
When constraints are involved, the Lagrange function is formed, as
shown in the table. This gives an unconstrained problem that can be solved
by the Euler and/or Euler Poisson equation, along with the constraint
equations.
The purpose of the chapter was to give some of the key results of the
calculus of variations and to emphasize the similarities and differences
between finding an optimal function and an optimal point. Consequently, it
was necessary to select the methods given here from some equally important
methods that were omitted. Two of these are the concept of a variation and
the use of the second variation for the sufficient conditions to determine if
the function was actually a maximum or minimum. These are discussed by
Courant and Hilbert (5) along with the problem of the existence of a solution.
Also, most texts discuss the moving (or natural) boundary problem where
one or both of the limits on the integral to be optimized can be a function of
the independent variable. This leads to extensions of the Brachistochrone
problem, and Forray's discussion (1) is recommended. With the background
of this chapter, extension to Hamilton's principle follows, and is typically the
next topic presented on the subject. Also, this material leads to extensions
that include Pontryagin's maximum principle; Sturm-Liouville problems;
and application in optics, dynamics of particles, vibrations, elasticity, and
quantum mechanics.
The calculus of variations can be used to solve transport phenomena
problems, i.e., obtain solutions to the partial differential equations
representing the conservation of mass, momentum, and energy of a system.
In this approach the partial differential equations are converted to the
corresponding integral to be optimized from the calculus of variations. Then
approximate methods of integration are used to find the minimum of the
integral, and this yields the concentration, temperature, and/or velocity
profiles required for the solution of the original differential equations. This
approach is described by Schechter (4) in some detail.
Again, the purpose of the chapter was to introduce the topic of finding the
optimal function. The references at the end of the chapter are recommended
for further information; they include the texts by Fan (12) and Fan and Wang
(13) on the maximum principle and Kirk (14), among others, (7, 15, 16) on
optimal control.
REFERENCES
1. Forray, M. J., Variational Calculus in Science and Engineering, McGraw-Hill Book Co., New
York (1968).
2. Ewing, G. M., Calculus of Variations with Applications, W. W. Norton Co., Inc., New York
(1969).
3. Weinstock, R., Calculus of Variations, McGraw-Hill Book Co., New York (1952).
4. Schechter, R. S., The Variational Method in Engineering, McGraw-Hill Book Co., New York
(1967).
5. Courant, R., and D. Hilbert, Methods of Mathematical Physics, Vol. 1, John Wiley & Sons,
Inc., New York (1953).
6. Sagan, H., Introduction to the Calculus of Variations, McGraw-Hill Book Co., New York
(1969).
7. M. M. Denn, Optimization by Variational Methods, McGraw-Hill Book Co., New York
(1970).
8. Wylie, C. R., and L. C. Barrett, Advanced Engineering Mathematics, 5th Ed., McGraw-Hill
Book Co., New York (1982).
9. Burley, D. M., Studies in Optimization, John Wiley & Sons, Inc., New York (1974).
10. Beveridge, G. S. G., and R. S. Schechter, Optimization: Theory and Practice, McGraw-Hill
Book Co., New York (1970).
11. Fan, L. T., E. S. Lee, and L. E. Erickson, Proc. of the Mid-Am. States Univ. Assoc. Conf on
Modern Optimization Techniques and their Application in Engineering Design, Part I, Kansas
State University, Manhattan, Kansas (Dec. 19-22, 1966).
12. L. T. Fan, The Continuous Maximum Principle, John Wiley & Sons Inc., New York (1966).
13. Fan, L. T. and C. S. Wang, The Discrete Maximum Principal, John Wiley & Sons, Inc., New
York (1964).
14. Kirk, D. E. Optimal Control Theory, An Introduction. Prentice-Hall, Inc., Englewood Cliffs,
N. J. (1970).
15. Miele, A., Optimization Techniques with Applications to Aerospace Systems, Ed., G. Leitman,
Ch.4, Academic Press, New York (1962).
16. Connors, M. M., and D. Teichroew, Optimal Control of Dynamic Operations Research
Models, International Textbook Co., Scranton, Pa. (1967).
PROBLEMS
where 𝑐1 and 𝑐2 are cost coefficients and 𝑃′ = 𝑑𝑃⁄𝑑𝑡. The total cost for the
change in the production schedule is given by:
𝒕𝟏
𝑪𝑻 = ∫ [𝒄𝟏 𝑷′𝟐 + 𝒄𝟐 𝒕𝑷′ ] 𝒅𝒕
𝒕𝟎
𝑳
𝑫 = 𝟒𝝅𝝆𝝂𝟐 ∫ 𝒚(𝒚′ )𝟑 𝒅𝒙
𝟎
where 𝜈 and 𝜌 are the free stream velocity and density, respectively.
a. Obtain the differential equation and boundary conditions that are to
be solved to obtain the optimum body shape.
b. Show that the following is the solution to the differential equation
obtained in (a):
𝒚 = (𝒅⁄𝟐)(𝒙⁄𝑳)𝟑⁄𝟒
𝒙𝟏
𝑰 = 𝟐𝝅 ∫ 𝒚(𝟏 + 𝒚′𝟐 )𝟏⁄𝟐 𝒅𝒙
𝒙𝟎
𝝏𝑭
𝑭 − 𝒚′ = 𝒄𝒐𝒏𝒔𝒕𝒂𝒏𝒕
𝝏𝒚′
𝒙 = 𝒄𝟏 𝒕 + 𝒄𝟐
𝒚𝟏 = 𝒄𝟏 𝐜𝐨𝐬𝐡(𝒕)
which is the parametric form of a family of catenaries, and 𝑐1 and
𝑐2 are boundary conditions for the end points of the curve.
9
8-4. Find the shape at equilibrium of a chain of length that hangs from
two points at the same level. The potential energy, 𝐸, of the chain is given
by the following equation:
𝒙𝟏
𝑬[𝒚(𝒙)] = −𝝆𝒈 ∫ 𝒚(𝟏 + 𝒚′𝟐 )𝟏⁄𝟐 𝒅𝒙
𝒙𝟎
𝒙𝟏
𝑳 = ∫ (𝟏 + 𝒚′𝟐 )𝟏⁄𝟐 𝒅𝒙
𝒙𝟎
Make the substitution 𝑘(𝑦 + 𝜆) = cosh 𝜃, and obtain the solution given
below:
This curve is the catenary, and the constants 𝑘, 𝛼, 𝜆 can be obtained from the
boundary conditions and the constraint on 𝐿.
8-5.14 A simple optimal control problem related to an electromechanical
system can be formulated as:
𝑻
𝑴𝒊𝒏𝒊𝒎𝒊𝒛𝒆: 𝑰 = ∫ [𝒚𝟐 𝟐 (𝒕) − 𝒚𝟑 𝟐 (𝒕)] 𝒅𝒕
𝟎
𝑺𝒖𝒃𝒋𝒆𝒄𝒕 𝒕𝒐: 𝒚𝟏 ′ + 𝒚𝟏 = 𝒚𝟑
𝒚 𝟐 ′ − 𝒚𝟏 = 𝟎
𝝏𝟐 𝝂 𝝏𝟐 𝝂 𝟏 𝒅𝑷
+ =
𝝏𝒙𝟐 𝝏𝒚𝟐 𝝁 𝒅𝒛
where 𝜈 is the axial velocity; 𝜇 is the viscosity of the fluid; 𝑑𝑃⁄𝑑𝑧 is the
pressure gradient, a constant; and 𝑎 is the length of one-half of the side of
the duct. The boundary conditions are that there is no slip at the wall, i.e.,
𝜈 = 0 at 𝑥 = 𝑎 for 0 < 𝑦 < 𝑎 and at 𝑦 = 𝑎 for 0 < 𝑥 < 𝑎.
a. Show that the integral to be minimized corresponding to the
differential equation is:
𝒂 𝒂
𝝏𝝂 𝟐 𝝏𝝂 𝟐 𝟐 𝒅𝑷
𝑰[𝝂(𝒙, 𝒚)] = ∫ ∫ [( ) + ( ) − 𝝂] 𝒅𝒙𝒅𝒚
𝟎 𝟎 𝝏𝒙 𝝏𝒚 𝝁 𝒅𝒛
𝝂 = 𝑨(𝒂𝟐 − 𝒙𝟐 )(𝒂𝟐 − 𝒚𝟐 )
𝒘 = 𝟎. 𝟓𝟓𝟔𝝆(𝒅𝑷⁄𝒅𝒛)𝒂𝟒 ⁄𝝁
The analytical solution has the same form, but the coefficient is
0.560. Thus, the approximate solution is within 1% of the exact
solution.
8-7. In a production scheduling problem, the production rate is to be changed
from 100 units per unit time to 300 units per unit time in 10 time units, i.e.,
𝑝(0) = 100 and 𝑝(10) = 300. The costs as a function of time are associated
with changes in machines, personnel, and raw materials. For this simple
problem this cost is given as:
where 𝑝′ = 𝑑𝑝⁄𝑑𝑡
Determine the production rate as a function of time that minimizes the
cost over the time.
8-8.1 Determine the deflection in an uniformly loaded, cantilever beam,
𝑦(𝑥), where 𝑦 is the deflection as a function of distance down the beam from
the wall (𝑥 = 0) to the end of the beam (𝑥 = 𝐿). The total potential energy
of the system to be minimized is given by:
𝑳
𝑰[𝒚(𝒙)] = ∫ [(𝑬⁄𝟐)(𝒚′′ )𝟐 − 𝒒𝒚] 𝒅𝒙
𝟎
where 𝐸 is the bending rigidity and 𝑞 is the load. The boundary conditions
at the wall end are 𝑦(0) = 𝑦 ′ (0) = 0, and at the supported end are 𝑦 ′′′ (𝐿) =
𝑦 ′′ (𝐿) = 0.