Sie sind auf Seite 1von 5

Dierential Equations

LECTURE 40

Separation of Variables and Heat Equation IVPs


1. Initial Value Problems Partial dierential equations generally have lots of solutions. To specify a unique one, well need some additional conditions. These conditions are usually motivated by the physics and come in two varieties: initial conditions and boundary conditions. An initial value problem for a PDE consists of the equation, initial conditions, and boundary conditions. An initial condition species the physical state at a given time t0 . For example, an initial condition for the heat equation would be the starting temperature distribution u(x, 0) = f (x). This is the only initial condition required because the heat equation is rst order with respect to time. The wave equation, which we will look at later, is second order with respect to time, and so needs two initial conditions. PDEs are also generally only valid on a certain domain. Boundary conditions specify how the solution is to behave on the boundary of this domain. These need to be specied, because the solution isnt on the one side of the boundary, meaning we might have problems with dierentiabiltiy there. Our heat equation was derived for a one-dimensional bar of length l, so the relevant domain in question can be taken to be the interval 0 < x < l and the boundary consists of the two points x = 0 and x = l. We could have derived a two-dimensional heat equation, for example, in which case the domain would be some region in the xy-plane with boundary some closed curve. It will usually be clear from a physical description of the problem what the appropriate boundary conditions are. We might know that, at the endpoints x = 0 and x = L, the temperatures u(0, t) and u(l, t) are xed. Boundary conditions that give the value of the solution are called Dirichlet conditions. Or we might insulate the ends, meaning there should be no heat ow out of the boundary; this would yield the boundary conditions ux (0, t) = ux (l, t) = 0. If the boundary conditions specify the derivative of the solution, theyre called Neumann conditions. We could also specify that we have one insulated end and at the other, we control the temperature; this is an example of a mixed boundary condition. As weve seen, changing boundary conditions can signicantly change the character of a problem. Initially, to get a feel for our solution method, well work with the homogeneous Dirichlet conditions u(0, t) = u(l, t) = 0, giving us the following initial value problem: (DE) : (BC) : (IC) : ut = kuxx u(0, t) = u(l, t) = 0 u(x, 0) = f (x). (40.1a) (40.1b) (40.1c)

After weve seen the general method, well see what happens with homogeneous Neumann conditions, leaving nonhomogeneous conditions for a later discussion. 1

Dierential Equations

Lecture 40: Separation of Variables and Heat Equation IVPs 2. Separation of Variables

In Section 1, we derived the heat equation (40.1a) for our bar of length L. Suppose weve got an initial value problem such as (40.1). How do we proceed? Well want to try to build up a general solution out of smaller solutions that are easier to nd. A rather straightforward approach is to start by assuming these mini-solutions have some nice form. For example, we might suppose we have a separated solution, where u(x, t) = X(x)T (t). That is, our solution is the product of a function that depends only on x and a function that depends only on t. We can then try to write down an equation depending only on x and another depending only on t before using our knowledge of ODEs to try to solve them. It should be noted that this is very special and should not be expected in general. In fact, this method cannot always be used, and even when it can be used often its hard to move past the rst step. However, it works for all of the equations we will be considering in this class, and is a valuable starting point. How does the method work? We begin by plugging our separated solution into the heat equation (40.1a). 2 [X(x)T (t)] = k 2 [X(x)T (t)] t x X(x)T (t) = kX (x)T (t) Now notice that we can move everything depending on x to one side and everything depending on t to the other. T (t) X (x) = kT (t) X(x) This equation should look a little funny to you. On the left, we have an expression which depends only on t, while on the right, we have an expression that depends only on x. Yet these two sides have to be equal for any choice of x and t we make! The only way this is possible is if both sides of the equation are the same constant. In other words, X (x) T (t) = = . kT (t) X(x) Weve written the minus sign explictly for convenience: it will turn out that > 0, but these expressions should be negative. The equation above really contains a pair of separate ordinary dierential equations, X + X = 0 T + kT = 0. (40.2a) (40.2b)

Notice that, with these separated equations, our boundary conditions (39.4b) become X(0) = 0 and X(l) = 0. Now, (40.2b) is easy enough to solve: we have T = kT , so that T (t) = Aekt . What about (40.2a)? This just gives us the boundary value problem X + X = 0 X(0) = 0 X(l) = 0.

This should look familiar: its very similar to Example 38.6. The only dierence is instead of our second condition occuring at x = 2, its at some x = l. As in that example, it will turn out that 2

Dierential Equations

Lecture 40: Separation of Variables and Heat Equation IVPs

our eigenvalues have to be positive. Letting = 2 for > 0, our general solution is X(x) = B cos(x) + C sin(x). The rst boundary condition says B = 0. The second condition says that X(l) = C sin(l) = 0. To avoid only having the trivial solution, we must have l = n. In other words, n = n l
2

and

Xn (x) = sin

nx l

for n = 1, 2, 3, . . . So we end up having found an innite number of solutions to our boundary value problem (40.1a) and (40.1b), one for each positive integer n (and no other values of n, since our eigenvalues are all positive). They are n 2 nx un (x, t) = An e( l ) kt sin . l The heat equation is linear and homogeneous. As such, the Principle of Superposition still holds: a linear combination of solutions is again a solution. So any function of the form
N

u(x, t) =
n=0

An e(

n 2 kt l

sin

nx l

(40.3)

is also a solution to (40.1a) and (40.1b). Notice that we havent used our initial condition (40.1c) yet, which is why we referred to (39.6) as a solutions to just the boundary value problem. How does our initial data come into play? We have
N

f (x) = u(x, 0) =
n=0

An sin

nx . l

So if our initial condition has this form, (40.3) works perfectly for us, with the coecients An just being the associated coecients from f (x). Example 40.1. Find the solution to the following heat equation problem on a rod of length 2. ut = uxx u(0, t) = u(2, t) = 0 u(x, 0) = sin 3x 2 5 sin(3x)

In this problem, we have k = 1. Now, we know that our solution will have the form of something like (40.3), since our initial condition is just the dierence of two sine functions. We just need to gure out which terms are represented and what the coecients An are. This isnt too hard to do: our initial condition is f (x) = sin 3x 2 5 sin(3x).

Looking at (40.3) with l = 2, we can see that the rst term corresponds to n = 3 and the second n = 6, and there are no other terms. Thus we have A3 = 1, A6 = 5, and all other An = 0. Our solution is then 2 2 3x 9 t 4 u(x, t) = e sin 5e(9 )t sin (3x) . 2

Dierential Equations

Lecture 40: Separation of Variables and Heat Equation IVPs

This isnt very satisfying: theres no reason to suppose that our initial distribution is a nite sum of sine functions. Physically, such situations are clearly very special. What do we do if we have a more general initial temperature distribution? Lets consider what happens if we take an innite sum of our separated solutions. Then our solution to (40.1a) and (40.1b) is

u(x, t) =
n=0

An e(

n 2 kt l

sin

nx . l

Now the initial condition species that the coecients must satisfy

f (x) =
n=0

An sin

nx . l

(40.4)

This idea is due to the French mathematician Joseph Fourier1, and (40.4) is called the Fourier sine series for f (x). There are several very important questions raised by this, however. Why should we believe that our initial condition f (x) ought to be able to be written as an innite sum of sines? Why should we believe that such a sum would converge to anything? Well come to these in good time, but keep them in mind. 2.1. Neumann boundary conditions. Now lets consider a heat equation problem with homogeneous Neumann conditions, (DE) : (BC) : (IC) : ut = kuxx ux (0, t) = ux (l, t) u(x, 0) = f (x). (40.5a) (40.5b) (40.5c)

Well start by again supposing that our solution to (40.5a) is separable, so we have u(x, t) = X(x)T (t) and, as were using the same dierential equation as before, we obtain the pair of ODEs (40.2a) and (40.2b). The solution to (40.2b) is still T (t) = Aekt . Now we need to determine the boundary conditions for (40.2a). Our boundary conditions (40.5b) are on ux (0, t) and ux (l, t); thus they are conditions on X (0) and X (l), as the t-derivative isnt controlled at all. So we have the boundary value problem X + X = 0 X (0) = 0 X (l) = 0. Along the lines of the analogous computation last lecture, this has eigenvalues and eigenfunctions n 2 n = l nx yn (x) = cos l for n = 0, 1, 2, . . . So individual separable solutions to (40.5a) and (40.5b) have the form n 2 nx un (x, t) = An e( l ) kt cos . l
1Fourier (1768-1830) was a big promoter of the French Revolution, and traveled with Napoleon to Egypt, where he was made governor of Lower Egypt. Fourier has been credited with being one of the rst to understand the greenhouse eect and, more generally, planetary energy balance. Theres a great (though almost certain apocryphal) story about what led Fourier to study the heat equation. Reportedly, what deeply concerned Fourier was being able to nd the ideal depth to build his wine cellar so that the wine would be stored at the perfect temperature year-round, and so he proceeded to try to understand the way heat propagated through the ground.

Dierential Equations

Lecture 40: Separation of Variables and Heat Equation IVPs

Taking nite linear combinations of these work similarly to the Dirichlet case (and is the solution to (40.5) when f (x) is a nite linear combination of constants and cosines, in direct analogy to (40.3)), but in general, were interested in knowing when we can take innite sums, i.e. 1 u(x, t) = A0 + 2

An e (

n 2 kt l

cos

Notice how we wrote the n = 0 case, as reason for this will be made clear in the future. The initial condition (40.5c) means that we need 1 f (x) = A0 + 2

n=1 1 2 A0 ; the

nx . l

An cos
n=1

nx . l

(40.6)

An expression of the form (40.6) is called the Fourier cosine series of f (x). 2.2. Other boundary conditions. Its also possible for certain boundary conditions to require the full Fourier series of the initial data; this is an expression of the form 1 f (x) = A0 + 2

An cos
n=1

nx nx + Bn sin l l

(40.7)

but for most of our purposes, well want to solve problems with Dirichlet or Neumann conditions. However, in the process of learning about Fourier sine and cosine series, well also learn how to compute the full Fourier series of a function.

Das könnte Ihnen auch gefallen