Higher order derivatives - Differentiation - Two-Dimensional Calculus

Two-Dimensional Calculus (2011)

Chapter 2. Differentiation

11. Higher order derivatives

If f(x, y) is continuously differentiable, then fx and fy are by definition continuous functions of x and y. We may ask whether they in turn are differentiable. If so, we may form (fx)x, (fx)y, etc. We adopt the following notation:

Image

These are called the second-order partial derivatives of f(x, y).

Example 11.1

Image

Example 11.2

Image

The equality of the so-called “mixed partial derivatives” fxy and fyx in the first example might pass for a coincidence, but it seems improbable in the second example unless there is a general rule. Indeed there is, and it is the following.

Theorem 11.1 If fxy and fyx exist and are continuous, then they are equal.

PROOF. Choose any point (x0, y0), and assume that the hypotheses of the theorem are satisfied inside some circle about (x0, y0). Let (x1, y1) be any other point in this circle, and form the function (see Fig. 11.1)

Image

By the mean-value theorem there is a y2 between y0 and y1 such that

Image

But by Eq. (11.1),

Image

where the second equality is obtained by applying the mean-value theorem to the function fy(x, y2). Combining Eqs. (11.1), (11.2), and (11.3) yields

Image

where (x2, y2) is some point in the rectangle indicated in Fig. 11.1.

Image

FIGURE 11.1 Notation for proof that fxy = fyx

The same reasoning applied to the function

Image

yields

Image

and

Image

whence

Image

where (x3, y3) is again a point in the same rectangle.

Comparing Eqs. (11.4) and (11.5), we see that

Image

Finally, as (x1, y1) → (x0, y0) we have (x2, y2) → (x0, y0) and (x3, y3) → (x0, y0), so that the continuity of fxy and fyx applied to Eq. (11.6) yields fyx(x0, y0) = fxy(x0, y0), which proves the theorem. Image

Remark Lest it be thought that equality of fxy and fyx is too obvious to need a proof, we give the following example.

Let

Image

where g(x, y) is defined and differentiable everywhere except at the origin. We shall see how to choose g(x, y) so that fxyfyx. First we note that f ≡ 0 on the x and y axes so that fx(0, 0) = fy(0, 0) = 0. Next, if (x, y) ≠ (0, 0), we have

Image

In particular,

Image

and the question reduces to the equality of these two limits. Now if g(x, y) is continuous at the origin, then both limits are the same and equal g(0, 0). However, a simple choice, such as

Image

yields

Image

and

Image

In the example that we have just given, the partial derivatives fxy and fyx exist everywhere, but they are not continuous at the origin, and hence Theorem 11.1 does not apply. The case where some derivative exists but is not continuous may be considered exceptional. In most problems, either derivatives of all orders exist and are continuous, as in the case of polynomials, trigonometric functions, and exponential functions, or else some derivative fails to exist, as in the case of fractional exponents. It is convenient to introduce the notation

Image

meaning f is continuously differentiable, and

Image

meaning all second-order derivatives of f exist and are continuous. Thus, if f Image Image, Th. 11.1 applies, and fxy = fyx, so that f has essentially three distinct second-order derivatives fxx, fxy, fyy.

In a similar manner we can form higher order derivatives, where an nth-order partial derivative is obtained by n successive differentiations, each with respect to either x or y. We introduce the notation

Image

Now if f Image Image, then all the (n − 1)-order partial derivatives of f are continuously differentiable, hence continuous, and therefore f Image Image. Thus, the classes Image are each included in the previous ones:

Image

where the notation Image denotes the class of continuous functions. Furthermore, we have

Image

Applying Th. 11.1, which allows interchanging the order of differentiation at each step, we come to the following conclusion.

If f Image Image, then the nth-order partial derivatives do not depend on the order of differentiation, but only on the total number of times f is differentiated with respect to x and y. If f is differentiated k times with respect to x and l times with respect to y, we obtain an nth-order derivative if k + l = n. The notation for this is

Image

Thus the third derivatives of f(x, y) are

Image

In general, a function f(x, y) Image Image has n + 1 distinct nth-order derivatives

Image

k = 0, 1, 2,…, n.

Example 11.3

Image

Thus, for this particular function, partial differentiation any number of times with respect to the variable y leaves the function unchanged; all derivatives involving one differentiation with respect to x and the rest with respect to y are equal to ey; and all derivatives involving two or more differentiations with respect to x are equal to zero.

Higher Order Directional Derivatives

The meaning of the second-order partial derivatives fxx(x0, y0), and fyy(x0, y0) is quite clear, since they are the ordinary second derivatives of the functions f(x, y0) and f(x0, y), respectively. The necessity for considering the quantity fxy becomes clear if we try to compute the ordinary second-derivative of f with respect to arc length along an arbitrary straight line through (x0, y0). If we let

Image

denote such a line, and use again the notation

Image

then

Image

We now introduce the notation

Image

wherever the derivatives on the right exist. Thus, the nth-order directional derivative of f in the direction α is the ordinary nth-order derivative of f with respect to arc length along the straight line in the direction α.

Suppose now that f Image Image. Then

Image

and since the right-hand side is continuously differentiable, we have

Image

or

Image

The importance of this formula is that it allows us to describe very accurately how the surface z = f(x, y) behaves near any given point (x0, y0). We need only observe that if the left-hand side, Image(x0, y0), is positive, then Image > 0, so that the curve on the surface lying over the straight line C is concave upwards (Fig. 11.2). Similarly, Image < 0 means that this curve is concave downwards. Suppose now that we fix (x0, y0) and set

Image

Then a, b, c are constants, and as α varies from 0 to 2π, (x, y) goes once around the circle X2 + Y2 = 1. Equation (11.7) becomes

Image

Image

FIGURE 11.2 Geometric interpretation of the sign of the second directional derivative

and we see that the totality of .values of Image(x0, y0) is precisely the set of values of a quadratic form on the unit circle. We may apply all the results of the previous section to describe the behavior of the surface z = f(x, y). For example, if the quadratic form is positive definite, then over each straight line the surface will be convex upwards, while if it is negative definite, the surface will be convex downwards (Fig. 11.3). If the quadratic form takes on both positive and negative values, then the surface will be convex upwards over certain lines through (x0, y0) and convex downwards over others. Such a point is called a saddle point (Fig. 11.4).

Image

FIGURE 11.3 Picture of surfaces whose second directional derivatives are positive definite or negative definite

Image

FIGURE 11.4 Saddle surface

Example 11.4

Image

This surface is concave up in the x direction and concave down in the y direction. Every point is a saddle point.

We shall discuss these cases in more detail in the next section, and also their use in maximum-minimum problems. We give now just one illustration of this line of reasoning.

Theorem 11.2 Suppose f(x, y) Image Image and

Image

Then f(x, y) can have neither a local maximum nor a local minimum at (x0, y0).

Remark This is a result that has no analog for functions of a single variable, since there is no condition analogous to (11.10).

PROOF. Condition (11.10) is precisely the condition acb2 < 0 that guarantees that the quadratic form (11.9) takes on both positive and negative values. In other words, there are a pair of lines C1, C2 through (x0, y0) such that for the corresponding directions α1, α2 we have Image Thus, Image and Image But from the theory of functions of a single variable, this means that Image cannot have a local maximum and Image cannot have a local minimum. But if f(x, y) had a local maximum or minimum at (x0, y0), then the same would be true of both Image and Image, which are merely restrictions of f to straight lines. Hence f can have neither, and the theorem is proved.

Image

Remark Condition (11.10) means geometrically that (x0, y0) is a saddle point, so that the theorem is intuitively obvious.

Example 11.5

Image

Here

Image

This function cannot have a local maximum or minimum anywhere.

Example 11.6

f(x, y) is a harmonic function. This means that f(x, y) satisfies the equation

Image

This equation is met in many areas of mathematics and its applications, and the study of its solutions has led to a theory, known as potential theory. We note that from Eq. (11.11) we have fxx = −fyy, and hence

Image

everywhere, with equality possible only if

Image

We have the following basic property: a harmonic function cannot have a local maximum or minimum unless it is constant. In fact, by what we have just observed, either

Image

in which case f cannot have a local maximum or minimum by Th. 11.2 or else

Image

in which case all the second-partial derivatives are zero. We can also show in this case that f cannot have a local maximum or minimum unless all higher derivatives vanish too, in which case f is constant.7 Examples of harmonic functions include all constants, all linear functions, and the quadratic function given in Example 11.5.

We conclude this section with the observation that Eq. (11.7) may be extended by induction to the general formula

Image

where Image is the binomial coefficient,

Image

Exercises

11.1 Find all second-order partial derivatives of the following functions.

a. f(x, y) = x6 − 5x4y2 + 3xy5

b. f(x, y) = ex + log y

c. f(x, y) = sin (y/x)

d. f(x, y) = tan (x + y)

e. f(x, y) = (xy)5

f. Image

11.2 Show that each of the following functions f(x, y) is harmonic, that is, fxx + fyy ≡ 0.

a. f(x, y) = xy

b. f(x, y) = x3 − 3xy2

c. f(x, y) = x/(x2 + y2)

d. f(x, y) = sin x cosh y

11.3 Suppose that the functions f(x, y), g(x, y) satisfy the Cauchy-Riemann equations

Image

Show that if f(x, y) Image Image and g(x, y) Image Image, then f(x, y) and g(x, y) are both harmonic.

11.4 Verify the statement of Ex. 11.3 for the following pairs of functions.

a. f(x, y) = x2y2, g(x, y) = 2xy

b. f(x, y) = ex cos y, g(x, y) = ex sin y

c. Image

11.5 a. Show that a quadratic form is a harmonic function if and only if it can be expressed as a(x2y2) + 2bxy, where a and b are arbitrary constants.

b. Show that if f(x, y) Image Image and f(x, y) is harmonic, then

Image

are harmonic also.

Note: in Exs. 11.6–11.9 assume that all functions are at least twice continuously differentiable.

11.6 Let f(x, y) = g(x) + h(y).

a. Show that fxy ≡ 0.

b. Show that if g″(x) > 0 and h″(y) < 0, then the surface z = f(x, y) cannot have a local maximum or minimum at any point.

c. Verify parts a and b for the function of Ex. 11.1b.

11.7 a. Show that the functions in Exs. 11.1d, e, f, satisfy fxxfyy ≡ 0.

b. Show that any function of the form

Image

satisfies the equation fxxfyy = 0.

11.8 Suppose that f(x, y) = g(x2 + y2).

a. Find fxx, fxy, and fyy in terms of the function g and its derivatives.

*b. Show that if f(x, y) is a harmonic function that depends only on the distance r of the point (x, y) to the origin, then

Image

where c and d are constants. (Hint: write f(x, y) in the form g(x2 + y2). Using part a, show that, with t = x2 + y2, g(t) satisfies the equation

Image

11.9 Let f(x, y) be homogeneous of degree k.

a. Show that f(x, y) satisfies the equations

Image

b. Show that f(x, y) satisfies

Image

11.10 Verify the equation in Ex. 11.9b for the functions of Exs. 11.1a, c, e.

11.11 Let f(x, y) = ex sin y.

a. Find all third-order derivatives of f(x, y).

b. Find

Image

c. Find

Image

d. Given any two integers k and l, with k + l = n, describe how nf/∂xk ∂yl depends on k and l.

11.12 Let f(x, y) = exy.

a. Find all fourth-order derivatives of f(x, y).

b. What is kf/∂xk?

c. What is k + 1f/∂xk ∂y?

d. What is k + 1f/∂x ∂yk?

*e. What is the value at the origin of k + 1f/∂xk ∂yl, for arbitrary k and l?

11.13 Let f(x, y) = 3x2 − 8xy + 2y2 + 12x − 4y + 17.

a. Find Image for arbitrary α.

b. Find an α such that Image > 0.

c. Find an α such that Image < 0.

d. Can f(x, y) have a local maximum or minimum at any point?

11.14 Let f(x, y) = ax2 + 2bxy + cy2 + dx + ey + f, where a and c have opposite signs.

a. By examining the function on any pair of lines x = x0 and y = y0, show that it is concave upwards in one case and downwards in the other, so that f(x, y) cannot have a local maximum or minimum at (x0, y0).

b. Use Th. 11.2 to derive the same conclusion.

11.15 Let f(x, y) = log [1/(3x + 7xy + 2y + 1)].

a. Find Image

b. For which values of α is Image(0, 0) > 0?

11.16 Let f(x, y) = exp {Image [7x2 + 10xy + 7y2]}. Find the maximum and minimum of Image(0, 0) for 0 ≤ α ≤ 2π.

11.17 Given fc(t) = f(x(t), y(t)), find Image.

11.18 Explain why the terms involving fx and fy in the answer to Ex. 11.17 do not appear in Eq. (11.7).

11.19 Let r = (x2 + y2)1/2.

a. Find rxx, rxy, ryy.

b. If x and y are functions of t, find d2r/dt2, first by using the answer to Ex. 11.17, and then by direct differentiation.

c. If Image is considered as a quadratic form in the variables X = cos α, Y = sin α, classify it according to the categories of the previous section, that is, positive definite, positive semidefinite, etc.

d. Sketch the surface z = (x2 + y2)1/2, and interpret geometrically your answer to part c.

11.20 Let z = (1 − x2y2)1/2, for x2 + y2 < 1.

a. Find zxx, zxy, zyy.

b. Answer Ex. 11.19c for this function.

c. Answer Ex. 11.19d for this function.

*11.21 a. Suppose the equation f(x, y) = c satisfies the assumptions of the implicit function theorem and defines a curve y = g(x). Show that if f(x, y) Image Image, then g(x) is twice continuously differentiable and

Image

b. Using the formula

Image

for the radius of curvature of a curve y = g(x), where y″ ≠ 0, find a formula for the radius of curvature of a curve in the implicit form f(x, y) = c.

c. Apply the formula of part b to find the radius of curvature at an arbitrary point of the circle x2 + y2 = r2.

11.22 a. Find the expression for Image by taking the directional derivative α of the right-side of Eq. (11.7).

b. Find Image by substituting n = 3 in Eq. (11.12), and compare the result with your answer to part a.

11.23 Carry out Ex. 11.22a, b for Image.

*11.24 Prove Eq. (11.12) by induction.

*11.25 Show that the function f(x, y) = x3 + x2y − 2xy2 + Imagey3 cannot have a local maximum or a local minimum at any point.

11.26 a. If f(x, y) = cxiyj, find kf/∂xk.

b. If f(x, y) = cxiyj, find k + lf/∂xk ∂yl.

c. If f(x, y) = cxiyj, find the value of k + lf/∂xk ∂y1 at (0, 0).

d. If

Image

is a homogeneous polynomial of degree n, find (mf/∂xk ∂yl)(0, 0), where k + l = m.

e. Show that the homogeneous polynomial f(x, y) of part d may be written in the form

Image

f. Show that an arbitrary polynomial P(x, y) of degree m may be written in the form

Image