The Euler-Lagrange Equation, or Euler's Equation

Definition 2   Let Ck[a, b] denote the set of continuous functions defined on the interval axb which have their first k-derivatives also continuous on axb.

The proof to follow requires the integrand F(x, y, y') to be twice differentiable with respect to each argument. What's more, the methods that we use in this module to solve problems in the calculus of variations will only find those solutions which are in C2[a, b]. More advanced techniques (i.e. beyond MATH0043) are designed to overcome this last restriction. This isn't just a technicality: discontinuous extremal functions are very important in optimal control problems, which arise in engineering applications.

Theorem 1   If I(Y) is an extremum of the functional

I(y) = $\displaystyle \int_{a}^{b}$ F(x, y, y') dx

defined on all functions yC2[a, b] such that y(a) = A, y(b) = B, then Y(x) satisfies the second order ordinary differential equation

% latex2html id marker 1635
$\displaystyle {\frac{{{\rm d}}}{{{\rm d} x}}}$$\displaystyle \left(\vphantom{ \frac{\partial F}{\partial y'} }\right.$$\displaystyle {\frac{{\partial F}}{{\partial y'}}}$$\displaystyle \left.\vphantom{ \frac{\partial F}{\partial y'} }\right)$ - $\displaystyle {\frac{{\partial
F}}{{\partial y}}}$ = 0. (1)

Definition 3   Equation ([*]) is the Euler-Lagrange equation, or sometimes just Euler's equation.

Warning 1   You might be wondering what ${\frac{{\partial F}}{{\partial y'}}}$ is suppose to mean: how can we differentiate with respect to a derivative? Think of it like this: F is given to you as a function of three variables, say F(u, v, w), and when we evaluate the functional I we plug in x, y(x), y'(x) for u, v, w and then integrate. The derivative ${\frac{{\partial F}}{{\partial y'}}}$ is just the partial derivative of F with respect to its second variable v. In other words, to find ${\frac{{\partial F}}{{\partial y'}}}$, just pretend y' is a variable.

Equally, there's an important difference between % latex2html id marker 1665
$ {\frac{{{\rm d} F}}{{{\rm d} x}}}$ and ${\frac{{\partial F}}{{\partial x}}}$. The former is the derivative of F with respect to x, taking into account the fact that y = y(x) and y' = y'(x) are functions of x too. The latter is the partial derivative of F with respect to its first variable, so it's found by differentiating F with respect to x and pretending that y and y' are just variables and do not depend on x. Hopefully the next example makes this clear:

Example 1   Let F(x, y, y') = 2x + xyy' + y'2 + y. Then

$\displaystyle {\frac{{\partial F}}{{\partial y'}}}$ = xy + 2y'    
% latex2html id marker 1686
$\displaystyle {\frac{{{\rm d}}}{{{\rm d} x}}}$$\displaystyle \left(\vphantom{ \frac{\partial F}{\partial y'} }\right.$$\displaystyle {\frac{{\partial F}}{{\partial y'}}}$$\displaystyle \left.\vphantom{ \frac{\partial F}{\partial y'} }\right)$ = y + xy' + 2y'    
$\displaystyle {\frac{{\partial
F}}{{\partial y}}}$ = xy' + 1    
$\displaystyle {\frac{{\partial F}}{{\partial x}}}$ = 2 + yy'    
% latex2html id marker 1693
$\displaystyle {\frac{{{\rm d} F}}{{{\rm d} x}}}$ = 2 + yy' + xy2 + xyy′′ +2y′′y′ + y    

and the Euler-Lagrange equation is

y + xy' + 2y' = xy' + 1$\displaystyle \qedhere$    

Warning 2   Y satisfying the Euler-Lagrange equation is a necessary, but not sufficient, condition for I(Y) to be an extremum. In other words, a function Y(x) may satisfy the Euler-Lagrange equation even when I(Y) is not an extremum.

Proof. Consider functions Yε(x) of the form

Yε(x) = Y(x) + εη(x)    

where η(x)∈C2[a, b] satisfies η(a) = η(b) = 0, so that Yε(a) = A and Yε(b) = B, i.e. Yε still satisfies the boundary conditions. Informally, Yε is a function which satisfies our boundary conditions and which is `near to' Y when ε is small.1 I(Yε) depends on the value of ε, and we write I[ε] for the value of I(Yε):

I[ε] = $\displaystyle \int^{b}_{a}$ F(x, Yε, Yε') dx.

When ε = 0, the function I[ε] has an extremum and so

% latex2html id marker 1732
$\displaystyle {\frac{{{\rm d}I}}{{{\rm d}\epsilon}}}$ = 0    when    ε = 0.

We can compute the derivative % latex2html id marker 1734
$ {\frac{{{\rm d}I}}{{{\rm d}\epsilon}}}$ by differentiating under the integral sign:

% latex2html id marker 1736
$\displaystyle {\frac{{{\rm d}I}}{{{\rm d}\epsilon}}}$ = % latex2html id marker 1737
$\displaystyle {\frac{{{\rm d}}}{{{\rm d}\epsilon}}}$$\displaystyle \int^{b}_{a}$F(x, Yε, Yε') dx = $\displaystyle \int^{b}_{a}$% latex2html id marker 1740
$\displaystyle {\frac{{{\rm d}F}}{{{\rm d} \epsilon}}}$(x, Yε, Yε') dx

We now use the multivariable chain rule to differentiate F with respect to ε. For a general three-variable function F(u(ε), v(ε), w(ε)) whose three arguments depend on ε, the chain rule tells us that

% latex2html id marker 1746
$\displaystyle {\frac{{{\rm d} F}}{{{\rm d} \epsilon}}}$ = $\displaystyle {\frac{{\partial F}}{{\partial
u}}}$% latex2html id marker 1748
$\displaystyle {\frac{{{\rm d} u}}{{{\rm d} \epsilon}}}$ + $\displaystyle {\frac{{\partial F}}{{\partial
v}}}$% latex2html id marker 1750
$\displaystyle {\frac{{{\rm d} v}}{{{\rm d} \epsilon}}}$ + $\displaystyle {\frac{{\partial F}}{{\partial
w}}}$% latex2html id marker 1752
$\displaystyle {\frac{{{\rm d} w}}{{{\rm d} \epsilon}}}$.    

In our case, the first argument x is independent of ε, so % latex2html id marker 1756
$ {\frac{{{\rm d} x}}{{{\rm d} \epsilon}}}$ = 0, and since Yε = Y + εη we have % latex2html id marker 1759
$ {\frac{{{\rm d} Y_\epsilon}}{{{\rm d} \epsilon}}}$ = η and % latex2html id marker 1761
$ {\frac{{{\rm d} Y_\epsilon'}}{{{\rm d} \epsilon}}}$ = η'. Therefore

% latex2html id marker 1763
$\displaystyle {\frac{{{\rm d} F}}{{{\rm d} \epsilon}}}$(x, Yε, Yε') = $\displaystyle {\frac{{\partial
F}}{{\partial y}}}$ η(x) + $\displaystyle {\frac{{\partial F}}{{\partial y'}}}$ η'(x).

Recall that % latex2html id marker 1767
$ {\frac{{{\rm d} I}}{{{\rm d} \epsilon}}}$ = 0 when ε = 0. Since Y0 = Y and Y0' = Y',

0 = $\displaystyle \int_{a}^{b}$$\displaystyle {\frac{{\partial
F}}{{\partial y}}}$(x, Y, Y')η(x) + $\displaystyle {\frac{{\partial F}}{{\partial y'}}}$(x, Y, Y')η'(x) dx. (2)

Integrating the second term in ([*]) by parts

$\displaystyle \int_{a}^{b}$$\displaystyle {\frac{{\partial F}}{{\partial y'}}}$η'(x) dx = $\displaystyle \left[\vphantom{ \frac{\partial
F}{\partial y'}\eta(x) }\right.$$\displaystyle {\frac{{\partial F}}{{\partial y'}}}$η(x)$\displaystyle \left.\vphantom{ \frac{\partial
F}{\partial y'}\eta(x) }\right]^{b}_{a}$ - $\displaystyle \int_{a}^{b}$% latex2html id marker 1782
$\displaystyle {\frac{{{\rm d}}}{{{\rm d} x}}}$$\displaystyle \left(\vphantom{ \frac{\partial F}{\partial y'} }\right.$$\displaystyle {\frac{{\partial F}}{{\partial y'}}}$$\displaystyle \left.\vphantom{ \frac{\partial F}{\partial y'} }\right)$η(x) dx.

The first term on the right hand side vanishes because η(a) = η(b) = 0. Substituting the second term into ([*]),

$\displaystyle \int^{b}_{a}$ % latex2html id marker 1789
$\displaystyle \left(\vphantom{ \frac{\partial F}{\partial y} - \frac{{\rm d}}{{\rm d} x}
\frac{\partial
F}{\partial y'}}\right.$$\displaystyle {\frac{{\partial
F}}{{\partial y}}}$ - % latex2html id marker 1791
$\displaystyle {\frac{{{\rm d}}}{{{\rm d} x}}}$$\displaystyle {\frac{{\partial F}}{{\partial y'}}}$% latex2html id marker 1793
$\displaystyle \left.\vphantom{ \frac{\partial F}{\partial y} - \frac{{\rm d}}{{\rm d} x}
\frac{\partial
F}{\partial y'}}\right)$η(x) dx = 0.

The equation above holds for any η(x)∈C2[a, b] satisfying η(a) = η(b) = 0, so the fundamental lemma of calculus of variations (explained on the next page) tells us that Y(x) satisfies

% latex2html id marker 1798
$\displaystyle {\frac{{{\rm d}}}{{{\rm d} x}}}$$\displaystyle \left(\vphantom{ \frac{\partial F}{\partial y'} }\right.$$\displaystyle {\frac{{\partial F}}{{\partial y'}}}$$\displaystyle \left.\vphantom{ \frac{\partial F}{\partial y'} }\right)$ - $\displaystyle {\frac{{\partial
F}}{{\partial y}}}$ = 0.$\displaystyle \qedhere$

$\qedsymbol$

Definition 4   A solution of the Euler-Lagrange equation is called an extremal of the functional.2

Exercise 1   Find an extremal y(x) of the functional

I(y) = $\displaystyle \int_{0}^{1}$(y' - y)2 dx,      y(0) = 0, y(1) = 2,    $\displaystyle \left[\vphantom{ \mbox{Answer:}
~~ y(x)=2 \frac{\sinh{x}}{\sinh{1}} }\right.$Answer:  y(x) = 2$\displaystyle {\frac{{\sinh{x}}}{{\sinh{1}}}}$$\displaystyle \left.\vphantom{ \mbox{Answer:}
~~ y(x)=2 \frac{\sinh{x}}{\sinh{1}} }\right]$.$\displaystyle \qedhere$

Exercise 2   By considering y + g, where y is the solution from exercise 1 and g(x) is a variation in y(x) satisfying g(0) = g(1) = 0, and then considering I(y + g), show explicitly that y(x) minimizes I(y) in Exercise 1 above. (Hint: use integration by parts, and the Euler-Lagrange equation satisfied by y(x) to simplify the expression for I(y + g)).

Exercise 3   Prove that the straight line y = x is the curve giving the shortest distance between the points (0, 0) and (1, 1).

Exercise 4   Find an extremal function of

I[y] = $\displaystyle \int_{1}^{2}$x2(y')2 + y dx,      y(1) = 1, y(2) = 1,    $\displaystyle \left[\vphantom{ \mbox{Answer:}~~
y(x)=\frac{1}{2}\ln{x}+\frac{\ln{2}}{x}+1-\ln{2} }\right.$Answer:  y(x) = $\displaystyle {\frac{{1}}{{2}}}$lnx + $\displaystyle {\frac{{\ln{2}}}{{x}}}$ +1 - ln2$\displaystyle \left.\vphantom{ \mbox{Answer:}~~
y(x)=\frac{1}{2}\ln{x}+\frac{\ln{2}}{x}+1-\ln{2} }\right]$.