LINE INTEGRALS

LINE INTEGRALS

10.1 Introduction

first for real-valued functions defined and bounded on finite intervals, and then for unbounded functions and infinite intervals. The concept was later extended to vector-valued functions and, in Chapter 7 of Volume II, to matrix functions.

In Volume I we discussed the integral

f(x)

This chapter extends the notion of integral in another direction. The interval [a, is replaced by a curve in n-space described by a vector-valued function a, and the integrand is a vector fieldfdefined and bounded on this curve. The resulting integral is called a line

integral, a curvilinear integral, or a contour integral, and is denoted by f da or by some similar symbol. The dot is used purposely to suggest an inner product of two vectors. The curve is called a path of integration.

Line integrals are of fundamental importance in both pure and applied mathematics. They occur in connection with work, potential energy, heat flow, change in entropy, circulation of a fluid, and other physical situations in which the behavior of a vector or

scalar field is studied along a curve.

10.2 Paths and line integrals

Before defining line integrals we recall the definition of curve given in Volume 1. Let a

be a vector-valued function defined on a finite closed interval [a, As runs through the function values a(t) trace out a set of points in n-space called

of the function. If a is continuous on J the graph is called a curve; more specifically, the curve described

by a.

In our study of curves in Volume I we found that different functions can trace out the same curve in different ways, for example, in different directions or with different velocities. In the study of line integrals we are concerned not only with the set of points on a curve but

with the actual manner in which the curve is traced out, that is, with the function a itself. Such a function will be called a continuous path.

which is continuous on J is called a continuous path in n-space. The path is called smooth if the derivative a’ exists and is continuous in the open interval (a,

DEFINITION.

Let J = [a, b] be

closed interval in

A function a:

The path is

Line integrals

smooth if the interval [a, b] can be partitioned into number of subintervals in each of which the path is smooth.

Figure 10.1 shows the graph of a piecewise smooth path. In this example the curve has a tangent line at all but a finite number of its points. These exceptional points subdivide the curve into arcs, along each of which the tangent line turns continuously.

F IGURE 10.1 The graph of a piecewise smooth path in the plane.

Let a be a piecewise smooth path in n-space on an interval [a, b], and let be a vector

DEFINITION OF LINE INTEGRAL.

defined and bounded on the graph of a. The line integral

f along a is denoted by the symbol

da and is defined by the equation

da =

. a’(t)

whenever the integral on the right exists, either as a proper or improper integral. Note: In most examples that occur in practice the dot

bounded on [a,

and continuous except possibly at a finite number of points, in which case the integral exists as a proper integral.

10.3 Other notations for line integrals

If C denotes the graph of a, the line integral

da is also written as

da and is

called the off along C.

If a = a(a) and = a(b) denote the end points of C, the line integral is sometimes written as

da and is called the line integral off from to along a. When the notation

for as

is used it should be kept in mind that the integral depends not only on the end points a and but also on the path a joining them.

When = the path is said to be closed. The symbol is often used to indicate integra- tion along a closed path. When

and a are expressed in terms of their components, say =

the integral on the right of (10.1) becomes a sum of integrals,

In this case the line integral is also written as

da, + +

da,.

325 In the two-dimensional case the path is usually

Other notations for line integrals

by a pair of parametric equations,

y=

and the line integral

dx In the three-dimensional case we use three parametric equations,

da is written as

and write the line integral

f da as

be a two-dimensional vector field given by

y) = Jy +

(0, 0) to (1, 1) along each of the following paths : (a) the line with parametric equations = , = , 0

for all (x, y) withy

0. Calculate the line integral

(b) the path with parametric equations =

0 1. Solution. For the path in part (a) we take a(t) = ti +

Then a’(t) = i + and +

and a’(t) is equal to + t and we find

+ t)j. Therefore the dot product of

da =

(Ji + + t) dt =

For the path in part (b) we take a(t) =

Then a’(t) = 2ti + and +

Therefore

a’(t) =

This example shows that the integral from one point to another may depend on the path joining the two points.

Now let us carry out the calculation for part (b) once more, using the same curve but with a different parametric representation. The same curve can be described by the function

where 0

Line integrals

This leads to the relation .

+ the integral of which from 0 to 1 is

(i

as before. This calculation illustrates that the value of the integral is independent of the parametric representation used to describe the curve. This is a general property of line integrals which is proved in the next section.

10.4 Basic properties of line integrals

Since line integrals are defined in terms of ordinary integrals, it is not surprising to find that they share many of the properties of ordinary integrals. For example, they have a linearity property with respect to the integrand,

and an additiveproperty with respect to the path of integration:

where the two curves

and

make up the curve C. That is, C is described by a function

are those traced out by a(t) as varies over subintervals [a, c] and. [c, b], respectively, for some c satisfying a < c < b. The proofs of these properties follow immediately from the definition of the line integral; they are left as exercises for the reader.

a defined on an interval [a, b], and the curves

and

Next we examine the behavior of line integrals under a change of parameter. Let a be a continuous path defined on an interval [a, b], let be a real-valued function that is differen- tiable, with

never zero on an interval [c, d], and such that the range of is [a, b]. Then the function defined on [c, d] by the equation

is a continuous path having the same graph as a. Two paths a and so related are called equivalent. They are said to provide different parametric representations of the same curve. The function is said to define a change of parameter.

the derivative of is always positive on [c, d] the function is increasing and we say that the two paths a and trace out C in the same direction. If the derivative of is always negative we say that a and trace out C in opposite directions. In the first case the function is said to be orientation-preserving;

Let C denote the common graph of two equivalent paths a and

in the second case is said to be orientation-reversing. An example is shown in Figure 10.2.

The next theorem shows that a line integral remains unchanged under a change of param- eter that preserves orientation; it reverses its sign if the change of parameter reverses orientation. We assume both intergals

da and

exist.

Basic properties of fine integrals

F IGURE 10.2 A change of parameter defined by = In (a), the function h preserves orientation. In (b), the function h reverses the orientation.

T H EO REM BEH A V IO R O F A LIN E IN T EG RA L U N D ER A C H A N G E O F PA RA M ET ER.

Let

a and be equivalent piecewise smooth paths. Then

if a and trace out C in the same direction; and

a and trace out C in opposite directions. P r o o f , It suffices to prove the theorem for smooth paths; then we invoke

the additive property with respect to the path of integration to deduce the result for wise smooth paths.

The proof is a simple application of the chain rule. The paths a and are related by an equation of the form

where is defined on an interval [c, d] and a is defined on an interval [a, b]. From the chain rule we have

Therefore we find

In the last integral we introduce the substitution = u(t), dv = u’(t) dt to obtain

. a’(v)

. a’(v)

Line integrals

where the + sign is used if a =

and b = u(d), and the

sign is used if a = u(d) and

b = u(c). The first case occurs if a and trace out C in the same direction, the second if they trace out C in opposite directions.

10.5 Exercises

In each of Exercises 1 through 8 calculate the line integral of the vector the path described.

1. = 2xy)i +

from

1) to (1, 1) along the parabola y =

xj, along the path described by

along the path described by a(t) = ti + +

4. y) = +

y )j, from to (2,0) along the curve y = 1 .

= in a counterclock- wise direction.

5. = (x + +

once around the ellipse

6. z) = 2xyi +

+ z)j + yk, from (1,

to

1) along a line segment.

along a line segment.

8. z) = xi +

along the path described by a(t) =

In each of Exercises 9 through 12, compute the value of the given line integral.

9. dx + (y ”

dy , where C is a path from

to (1, 1) along the parabola

y (x

(x

traversed once in a +

10. , where C is the circle +

clockwise direction. dx + dy where C is the square with vertices (1, 0), (0,

1 , 0), and (0, traversed

once in a counterclockwise direction.

12. + zdy + xdz, where (a) C is the curve of intersection of the two surfaces x + y = 2 and +

+ = 2(x + y) . The curve is to be traversed once in a direction that appears clockwise when viewed from the origin.

= traversed once in a direction that appears counterclockwise when viewed from high above the xy-plane.

(b) C is the intersection of the two surfaces z = xy and

10.6 The concept of work as a line intregal

Consider a particle which moves along a curve under the action of a force If the

curve is the graph of a piecewise smooth path a, the w ork done by f is defined to be the line

integral

f da. The following examples illustrate some fundamental properties of work. EXAMPLE 1. W ork done by a constantforce. If f is a constant force, say f= it can be

shown that the work done by f in moving a particle from a point to a point b along any

piecewise smooth path joining a and b is

(b

a), the dot product of the force and the

displacement vector b a. We shall prove this in a special case.

We let a = ...,

be a path joining a and

say a(a) = a and a(b) = b, and we

Line integrals with respect to arc length

write = (c,, . . . , c,). Assume a’ is continuous on [a, b]. Then the work done by f is

equal to

da =

dt =

a,(a)] = c . [a(b) a(a)] c .

For this force field the work depends only on the end points a and b and not on the curve joining them. Not all force fields have this property. Those which do are called conservative. The example on p. 325 is a nonconservative force field. In a later section we shall determine all conservative force fields.

EXAMPLE 2. The principle of work and energy. A particle of mass moves along a curve

under the action of a force field f. If the speed of the particle at time is v(t), its kinetic

energy is defined to be Prove that the change in kinetic energy in any time interval

is equal to the work done by f during this time interval.

Solution. Let r(t) denote the position of the particle at time t. The work done by f

during a time interval [a, b] is

f dr . We wish to prove that

dr =

From Newton’s second law of motion we have

= mu’(t),

where denotes the velocity vector at time The speed is the length of the velocity vector, v(t) =

. Therefore

(u(t) u(t)) = Integrating from a to b we obtain

f [r(t)] r’(t) = f [r(t)] . u(t) = mu’(t) u(t) =

f dr =

f [r(t)] r’(t) dt =

as required.

10.7 Line integrals with respect to arc length

Let a be a path with a’ continuous on an interval [a, b]. The graph of a is a rectifiable curve. In Volume I we proved that the corresponding arc-length function is given by the integral

The derivative of arc length is given by

Line integrals

Let be a scalar field defined and bounded on C, the graph of a. The line integral of with respect to arc length along C is denoted by the symbol

and is defined by the equation

ds =

whenever the integral on the right exists. Now consider a scalar field given by

the dot product of a vector fieldfdefined on C and the unit tangent vector T(t) =

. In this case the line integral

is the same as the line integral

da because

. a’(t) =

When f denotes a velocity field, the dot product T is the tangential component of

integral off along C. When C is a closed curve the flow integral is called the circulation off along C. These terms are commonly used in the theory of fluid flow.

velocity, and the line integral

f T ds is called

10.8 Further applications of line integrals

Line integrals with respect to arc length also occur in problems with mass distribution along a curve. For example, think of a curve C in 3-space as a wire made of a thin material of varying density. Assume the density is described by a scalar field

where is the

z) of C. Then the total mass of the wire is defined to be the line integral of with respect to arc length:

unit length at the point (x,

The center of mass of the wire is defined to be the point whose coordinates are determined by the equations

A wire of constant density is called uniform. In this case the center of mass is also called the centroid.

EXAMPLE 1. Compute the mass of one coil of a spring having the shape of the helix whose vector equation is

if the density at (x, z) is

Solution. The integral for A4 is

ds =

(a”

t+

dt .

331 Since s’(t) =

Exercises

we have s’(t) and hence

and a’(f) = -a sin

i + a cos tj

(a’ +

In this example the z-coordinate of the center of mass is given by

+ z”)

The coordinates and are requested in Exercise 15 of Section 10.9.

Line integrals can be used to define the moment of inertia of a wire with respect to an axis. If

z) represents the perpendicular distance from a point (x, z) of C to an axis L, the moment of inertia

is defined to be the line integral

where z) is the density at (x, z). The moments of inertia about the coordinate

axes are denoted by I,, I,, and I,.

EXAMPLE 2. Compute the moment of inertia of the spring coil in Example 1. Solution. Here

+ + so we have

ds =

+ z”) ds =

where is the mass, as computed in Example 1.

10.9 Exercises

. Compute the work done by this force in moving a particle from (0,

1. A force field f in

is given by

z) = xi + + (xz

along the line segment joining these two points.

to (1,

in moving a particle (in a counterclockwise direction) once around the square bounded by the coordinate axes and thelinesx =aandy

2. Find the amount of work done by the

>O.

3. A two-dimensional force fieldfis given by the y) = cxyi + where c is a positive constant. This force acts on a particle which must move from (0, 0) to the line x = 1 along a curve of the form

where a> 0

and

b >O .

Find a value of a (in terms of c) such that the work done by this force is independent of b.

4. A force

is given by the

z) = yzi +

+ + Calculate

the work done by fin moving a particle once around the triangle with vertices (0, 0, 0), (1, 1,

(-1, 1, -1) in that order.

Line integrals

+ (x along the curve of intersection of the sphere

5. Calculate the work done by the force

+ (z

+ z = 4 and the plane z = tan where 0< <

The path is transversed in a direction that appears counterclockwise when viewed from high above the xy-plane. 6. Calculate the work done by the force field

+ along the curve of intersection of the sphere

z) =

= ax, where z 0 and a > 0. The path is traversed in a direction that appears clockwise when viewed from high above the xy-plane.

and the cylinder

Calculate the line integral with respect to arc length in each of Exercises 7 through IO. 7. (x +

traversed in a counterclockwise direction. 8. where C has the vector equation

where C is the triangle with vertices (0, 0), (1, 0), and (0,

a(t) =

sin

cos

9. + where C has the vector equation a(t) =

t + t sin

t cos

10. z ds, where C has the vector equation

a(t) = tcosti + tsintj + tk,

11. Consider a uniform semicircular wire of radius (a) Show that the centroid lies on the axis of symmetry at a distance

from the center. (b) Show that the moment of inertia about the diameter through the end points of the wire

is where is the mass of the wire. 12. A wire has the shape of the circle

Determine its mass and moment of inertia about a diameter if the density at (x,

is

13. Find the mass of a wire whose shape is that of the curve of intersection of the sphere + + = 1 and the plane x + + z = 0 if the density of the wire at (x,

is 14. A uniform wire has the shape of that portion of the curve of intersection of the two surfaces

Find the z-coordinate of its centroid. 15. Determine the coordinates

= x connecting the points (0,

and (1, 1,

of the center of mass of the spring coil described in Example 1 of Section 10.8. 16. For the spring coil described in Example 1 of Section 10.8, compute the moments of inertia and .

10.10 Open connected sets. Independence of the path

Let S be an open set in R”. The set S is called connected if every pair of points in can be joined by a piecewise smooth path whose graph lies in S. That is, for every pair of points a and in S there is a piecewise smooth path a defined on an interval [a,

such that a(t) S for each in [a, b], with a(a)

a and a(b) =

Three examples of open connected sets in the plane are shown in Figure 10.3. Examples in 3-space analogous to these would be (a) a solid ellipsoid, (b) a solid polyhedron, and (c)

a solid torus; in each case only the interior points are considered.

333 An open set is said to be disconnected if is the union of two or more disjoint

The second fundamental theorem of calculus for line integrals

empty open sets. An example is shown in Figure 10.4. It can be shown that the class of open connected sets is identical with the class of open sets that are not disconnected.?

be a vector field that is continuous on an open connected set S. Choose two points and in S and consider the line integral off from to along some piecewise smooth path in S. The value of the integral depends, in general, on the path joining a to

Now let

For some vector fields, the integral depends only on the end points a and and not on the path which joins them. In this case we say the integral is independent of

from to We say the line integral off is independent of

in if it is independent of the path from

a to b for every pair of points a and b in S.

SS

F IGURE 10.3 Examples of open connected sets. F IGURE 10.4 A disconnected set S, the union of two disjoint circular disks.

Which have line integrals independent of the path? To answer this question we extend the first and second fundamental theorems of calculus to line integrals.

10.11 The second fundamental theorem of calculus for line integrals

The second fundamental theorem for real functions, as proved in Volume I (Theorem states that

provided that is continuous on some open interval containing both a and b. To extend this result to line integrals we need a slightly stronger version of the theorem in which continuity of

is assumed only in the open interval (a, b).

THEOREM 10.2. Let be a realfunction that is continuous on a closed interval [a, b] and

assume that the integral y’(t) dt exists. If is continuous on the open interval (a, b), we

have

Proof. For each in [a, b] define f(x) = dt . We wish to prove that (10.2)

For a further discussion of connectedness, see Chapter 8 of the author’s Analysis,

Line integrals

By Theorem 3.4 of Volume continuous on the closed interval [a, b]. By Theorem 5.1 of Volume

differentiable on the open interval (a, b), = for each x in (a, b). Therefore, by the zero-derivative theorem (Theorem 5.2 of Volume I), the difference

is constant on the open interval (a, b). By continuity, f is also

constant on the closed interval [a, b]. In particular, f(b)

But = 0, this proves (10.2).

= f(a)

THEOREM

10.3. SECOND FUNDAMENTAL THEOREM OF CALCULUS FOR LINE INTEGRALS.

Let be a

with a continuous gradient

on an open connected set

S in R". Then for any two points a and b joined by a piecewise smooth path a in S we have

da =

Choose any two points a and b in S and join them by a piecewise smooth path a in S defined on an interval [a, b]. Assume first that a is smooth on [a, b]. Then the line integral of

Proof.

from a to b along a is given by

da =

a’(t) dt

By the chain rule we have

where g is the composite function defined on [a, b] by the formula

The derivative is continuous on the open interval (a, b) because is continuous on S and a is smooth. Therefore we can apply Theorem 10.2 to to obtain

This proves the theorem if a is smooth.

When a is piecewise smooth we partition the interval [a, b] into a finite number (say r)

of subintervals , in each of which a is smooth, and we apply the result just proved to each subinterval. This gives us

as required.

As a consequence of Theorem 10.3 we see that the line integral of a gradient is inde- pendent of the path in any open connected set S in which the gradient is continuous. For a closed path we have b = a, so

= 0. In other words, the line integral of a

335 continuous gradient is zero around every piecewise smooth closed path in S. In Section 10.14

Applications to mechanics

we shall prove (in Theorem 10.4) that gradients are the continuous vector fields with this property.

10.12 Applications to mechanics

If a vector field is the gradient of a scalar field then is called a potential function In 3-space, the level sets of are called equipotential surfaces; in 2-space they are called equipotential lines. (If

denotes temperature, the word “equipotential” is replaced by “isothermal”; if denotes pressure the word “isobaric” is used.)

+ For every integer we have

EXAMPLE 1. In

where r = xi + + (See Exercise of Section 8.14.) Therefore is a potential of the vector field

z) =

The equipotential surfaces of are concentric spheres centered at the origin. EXAMPLE 2. The Newtonian potential. Newton’s gravitation law states that the force

which a particle of mass exerts on another particle of mass m is a vector of length where G is a constant and is the distance between the two particles. Place the origin at the particle of mass M, and let r = xi +

be the position vector from the origin to the particle of mass m. Then =

and --Y/r is a unit vector with the same direction as

f, so Newton’s law becomes

Taking =

1 in Example 1 we see that the gravitational force is the gradient of the scalar field given by

z) =

This is called the Newtonian potential.

The work done by the gravitational force in moving the particle of mass m from

where = + +

If the two points lie on the same equipotential surface then

and =

and no work is done.

EXAMPLE 3. Theprinciple of conservation of mechanical energy. Let f be a continuous force

field having a potential in an open connected set S. Theorem 10.3 tells us that the work done by fin moving a particle from to along any piecewise smooth path in S is

Line integrals

the change in the potential function. In Example 2 of Section 10.6 we proved that this work is also equal to the change in kinetic energy of the particle, k(x)

k(a) where k(x) denotes the kinetic energy of the particle when it is located at x. Thus, we have

or (10.3)

The scalar is called the potential

of the particle.

If a is kept fixed and is allowed to vary over the set Equation (10.3) tells us that the sum of k(x) and

is a gradient, the sum of the kinetic andpotential energies of aparticle moving in

is constant. In other words, if a

is constant. In mechanics this is called the principle of conservation of (mechanical) energy. A force field with a potential function is said to be conservative because the total energy, kinetic plus potential, is conserved. In a conservative field, no net work is done in moving a particle around a closed curve back to its starting point. A force field will not be conservative if friction or viscosity exists in the system, since these tend to convert mechanical energy into heat energy.

10.13 Exercises

Determine which of the following open sets

in

are connected. For each connected set,

choose two arbitrary distinct points in and explain

curve in S connecting the two points.

2. Given a two-dimensional vector field

where the partial derivatives

are continuous on an open set S. If is the gradient of some potential

and

prove that

at each point of S.

a gradient. Then find a closed path C such that =

3. For each of the following vector fields, use the result of Exercise 2 to prove that is

xi. =

4. Given a three-dimensional vector field

Some authors refer to

as the potential function

that the potential energy at x will be equal to the that the potential energy at x will be equal to the

line integrals

where the partial derivatives

are continuous on an open set S. Iffis the gradient of some potential function prove that

at each point of

5. For each of the following vector fields, use the result of Exercise 4 to prove that is not a gradient. Then find a closed path C such that

+xj +xk.

(b) z) = +

6. A force field is defined in

by the equation

(a) Determine whether or not fis conservative. (b) Calculate the work done in moving a particle along the curve described by

a(t) = cos ti + sin

as t runs from 0 to

7. A two-dimensional force field F is described by the equation

(a) Show that the work done by this force in moving a particle along a curve

(b) Find the amount of work done = 1, f(b) = 2, g(a) = 3, g(b) = 4.

8. A force field is given in polar coordinates by the equation

= -4 sin

4 sin

Compute the work done in moving a particle from the point to the origin along the spiral whose polar equation is r =

9. A radial or “central” force field F in the plane can be written in the form

y)

where r = xi + and r = . Show that such a force field is conservative.

+ 2)i + 16xj. in moving a particle from 1, to (1, 0) along the upper half of the ellipse

10. Find the work done by force

Which ellipse (that is, which value of b) makes the work a minimum?

10.14 The first fundamental theorem of calculus for line integrals

Section 10.11 extended the second fundamental theorem of calculus to line integrals. This section extends the first fundamental theorem. We recall that the first fundamental theorem states that every indefinite integral of a continuous function

has a derivative

Line integrals

equal to f. That is, if

then at the points of continuity off we have

=f

To extend this theorem to line integrals we begin with a vector continuous on an open connected set S, and integrate it along a piecewise smooth curve C from a fixed point a in S to an arbitrary point x. Then we let denote the scalar field defined by the line integral

where a describes C. Since S is connected, each point in S can be reached by such a curve. For this definition of

to be unambiguous, we need to know that the integral depends only on and not on the particular path used to join a to x. Therefore, it is natural

to require the line integral off to be independent of the path in S. Under these conditions, the extension of the first fundamental theorem takes the following form:

THEOREM

10.4. FIRST FUNDAMENTAL THEOREM FOR LINE INTEGRALS. Let f be a vector that is continuous on an open connected set in

and assume that the line integral off is independent of the path in S. Let a be

a on S by the equation

of S and

where a is any piecewise smooth path in S joining a to x. Then the gradient of exists and is equal to f; that is,

= f(x)

for every in

Proof, We shall prove that the partial derivative exists and is equal to the kth component of f(x), for each k = 2, . . . , and each in S. Let

r) be an n-ball with center at and radius r lying in S. If is a unit vector, the point

also lies in S for every real h satisfying 0 < < r, and we can form the difference quotient

Because of the additive property of line integrals, the numerator of this quotient can be written as

da,

and the path joining to + can be any piecewise smooth path lying in S. In particular,

Necessary and

conditions for a

to be a gradient

we can use the line segment described by

a(t) = + thy,

where 0

Since a’(t) = , the difference quotient becomes

Now we take y = the kth unit coordinate vector, and note that the integrand becomes + thy) . y =

+ the,). Then we make the change variable = ht , = h dt , and we write (10.4) in the form

du

where g is the function defined on the open interval (-r, r) by the equation

the first fundamental theorem for ordinary integrals tells us that g’(t) exists for each in (-r, r) and that

Since each component

is continuous on

In particular, g’(0)

Therefore, if we let h

0 in (10.5) we find that

g’(0) =

This proves that the partial derivative

exists and

as asserted.

10.15 Necessary and sufficient conditions for a vector field to be a gradient

The first and second fundamental theorems for line integrals together tell us that a necessary and sufficient condition for a continuous vector field to be a gradient on an open connected set is for its line integral between any two points to be independent of the path.

We shall prove now that this condition is equivalent to the that the line integral is zero around every piecewise smooth closed path. All these conditions are summarized in the following theorem.

T H EO REM N EC ES S A RY A N D S U FFIC IEN T C O N D IT IO N S FO R A V EC T O R FIELD T O BE A GRAD I ENT .

Let f be a continuous on an open connected set S in Then the three statements are equivalent. (a) is the gradient of somepotentialfinction in S. (b) The line integral off is independent of the path in S. (c) The line integral off is zero around every piecewise smooth closedpath in S.

Line integrals

We shall prove that (b) implies (a), (a) implies (c), and (c) implies (b). State- ment (b) implies (a) because of the first fundamental theorem. The second fundamental theorem shows that (a) implies (c).

To complete the proof we show that (c) implies (b). Assume (c) holds and let and

be any two piecewise smooth curves in with the same end points. Let

be the graph of a function a defined on an interval [a,

and let

be the graph of a function defined on

Define a new function y as follows:

if =

if

Then y describes a closed curve C such that

Since

dy = because of (c), we have

f da =

f so the integral off is

independent of the path. This proves (b). Thus, (a), (b), and (c) are equivalent.

closed curve C, then f is not a gradient. However, if a line integral

Note: If f 0 for a

f is zero for a particular closed curve C or even for infinitely many closed curves, it does not necessarily follow that f is a gradient. For example, the reader

can easily verify that the line integral of the vector field f(x,

= xi +

is zero for every

circle C with center at the origin. Nevertheless, this particular vector field is not a gradient.

10.16 Necessary conditions for a vector field to be a gradient

The first fundamental theorem can be used to determine whether or not a given vector field is a gradient on an open connected set S. If the line integral off is independent of the path in

we simply define a scalar field by integrating f from some fixed point to an arbitrary point in along a convenient path in S. Then we compute the partial derivatives of and compare

= for every in and every k, then f is a gradient on and is a potential. If

the kth component off. If

for some k

and some x, then f is not a gradient on S. The next theorem gives another test for determining when a vector field f is not a gradient.

This test is especially useful in practice because it does not require any integration.

THEOREM 10.6. NECESSARY CONDITIONS FOR A VECTOR FIELD TO BE A GRADIENT. Let .,

vector field on an open set in R”. is a gradient on then the partial derivatives of the components off are related by the equations

be a continuously

for j = . . . , n and every in

Necessary conditions for a vector field to be a gradient

Proof.

a gradient, then for some potential function This means that

foreachj= Differentiating both members of this equation with respect to we find

Similarly, we have

Since the partial derivatives

are continuous on S, the two mixed partial derivatives

and

and must be equal on S. This proves (10.6).

EXAMPLE

Determine whether or not the vector field

is a gradient on any open subset of Solution. Here we have

The partial derivatives

and

are given by

y) =

this vector field is not a gradient on any open subset of

except when x = 0 or y =

The next example shows that the conditions of Theorem 10.6 are not always sufficient for a vector field to be a gradient.

be the vector field defined on S by the equation

EXAMPLE 2. Let S be the set of all (x, y)

in

and let

Show that

everywhere on S but that, nevertheless, f is not a gradient on S.

y) for all y) in S. (See Exercise 17 in Section 10.18.)

Solution.

The reader can easily verify that

y) =

To prove thatfis not a gradient on S we compute the line integral offaround the unit circle given by

Line integrals

We obtain

da =

a'(t)

Since the line integral around this closed path is not zero, f is not a gradient on S. Further

properties of this vector field are discussed in Exercise 18 of Section 10.18.

At the end of this chapter we shall prove that the necessary conditions of Theorem 10.6 are also sufficient if they are satisfied on an open

set. (See Theorem 10.9.)

10.17 Special methods for constructing potential functions

The first fundamental theorem for line integrals also gives us a method for constructing

potential functions. If f is a continuous gradient on an open connected set S, the line

integral off is independent of the path in S. Therefore we can find a potential simply by

using any piecewise smooth path lying in S. The scalar field so obtained depends on the choice of the initial point a. If we start from another initial point, say

integrating f from some fixed point a to an arbitrary point

in

we obtain a new potential But, because of the additive property of line integrals,

can differ only by a constant, this constant being the integral off from a to

and

The following examples illustrate the use of different choices of the path of integration.

F IGURE 10.5 Two polygonal paths from (a, to (x,

EXAMPLE 1. Construction of a potential on an open rectangle. If f is a continuous gradient

can be constructed by integrating from a fixed point to an arbitrary point along a set of line segments parallel to the coordinate axes.

on an open rectangle in

a potential

A two-dimensional example is shown in Figure 10.5. We can integrate first from (a, b) to (x,

along a vertical segment. Along the horizontal segment we use the parametric representation

along a horizontal segment, then from (x,

to (x,

343 and along the vertical segment we use the representation

Special methods for constructing potential functions

a(t) = xi + tj,

y) + the resulting formula for a potential y) is

We could also integrate first from (a, b) to (a, y) along a vertical segment and then from (a, y) to (x, y) along a horizontal segment as indicated by the dotted lines in Figure 10.5. This gives us another formula for

y),

Both formulas (10.7) and (10.8) give the same value y) because the line integral of a gradient is independent of the path.

EXAMPLE 2. Construction of a potentialfunction by the use of integrals. The use of indefinite integrals often helps to simplify the calculation of potential functions. For example, suppose a three-dimensional vector field

is the gradient of a potential function on an open set in

Then we

everywhere on S. Using indefinite integrals and integrating the first of these equations with respect to x (holding y and z constant) we find

where z) is a “constant of integration” to be determined. Similarly, if we integrate the equation

with respect to y and the equation

with respect to z we

obtain the further relations

and

=j

where and y) are functions to be determined. Finding means finding three functions

z), B(x, z), and y) such that all three equations for y, z) agree in their right-hand members. In many cases this can be done by inspection, as illustrated by

Line integrals

EXAMPLE 3. Find a potential function for the vector field defined on by the equation =

4xy)j +

+ 2)k.

Solution. Without knowing in advance whether or not has a potential function we try to construct a potential as outlined in Example 2, assuming that a potential

exists. Integrating the

with respect to x we find

Integrating with respect to

and

with respect to

we find

By inspection we see that the choices

z) = +x and y)

z) =

x will make all three equations agree; hence the function given by the equation

is a potential for f on

EXAMPLE 4. Construction of a potential on a convex set.

A set S in

is called convex

if every pair of points in S can be joined by a line segment, all of whose points lie in S. An example is shown in Figure 10.6. Every open convex set is connected.

Iff is a continuous gradient on an open convex set, then a potential can be constructed by integratingf from a fixed point in

along the line segment joining

to an arbitrary point

to The line segment can be parametrized by the function

1. This gives us a’(t) =

a(t)

where 0

a, so the corresponding potential is given by the integral

F IGURE 10.6 In a convex set S, the segment joining a and is in for all points a and

345 If S contains the origin we can take a = 0 and write (10.9) more simply as

Exercises

10.18 Exercises In each of Exercises 1 through 12, a vector fieldfis defined by the formulas given. In each case

determine whether or not is the gradient of a scalar field. When is a gradient, find a correspond- ing potential function

1. = xi

2. y) = +

3. f(x, y) =

5. f(x, y) = [sin (xy) + xy cos

13. A fluid flows in the xy-plane, each particle moving directly away from the origin. If a particle is at a distance r from the origin its speed is

where a and are constants. (a) Determine those values of a and for which the velocity vector field is the gradient of some scalar field.

(b) Find a potential function of the velocity whenever the velocity is a gradient. The case = -1 should be treated separately.

14. If both and are potential functions for a continuous vector field f on an open connected set S in

prove that is constant on S. 15. Let S be the set of all x 0 in

Let r =

, and let f be the vector field defined on S by

the equation

f(x) =

where

is a real constant. Find a potential function for f on S. The case = -2 should be treated separately. 16. Solve Exercise 15 for the vector field defined by the equation

where is a real function with a continuous derivative everywhere on

The following exercises deal with the vector field f defined on the set S of all points (x, y)

in by the equation

Line integrals

In Example 2 of Section 10.16 we showed that f is not a gradient on S, even though

everywhere on S.

17. Verify that for every point (x,

in S we have

18. This exercise shows that f a gradient on the set

consisting of all points-in the xy-plane except those on the nonpositive x-axis. (a) If (x, y)

T, express and in polar coordinates,

sin

wherer > Oand < Prove that is given by the formulas

Y if > 0, if

if < 0.

t is the unique real number which satisfies the two conditions tan = t and

[Recall the definition of the arc tangent function: For any real t,

< < (b) Deduce that

ae Y

for all (x, y) in T. This proves that is a potential function for f on the set T.

This exercise illustrates that the existence of a potential function depends not only on the vector field f but also on the set in which the vector field is defined.

10.19 Applications to exact differential equations of first order

Some differential equations of the first order can be solved with the aid of potential functions. Suppose we have a first-order differential equation of the form

If we multiply both sides by a nonvanishing factor y) we transform this equation to the form

y) for y) and use the Leibniz notation for derivatives, writing

y) = 0. If we then write

for y’, the differential equation takes the form

Applications to exact

equations

order 347

We assume that P and Q are continuous on some open connected set S in the plane. With each such differential equation we can associate a vector field V, where

The components and Q are the coefficients of dx and dy in Equation (10.11). The differ- ential equation in (10.11) is said to be exact in S if the vector field V is the gradient of a potential; that is, if

where is some scalar field. When such a exists we have

y) =

y) for each point (x, y) in

= Q, and the differential equation in (10.11) becomes

= P and

We shall prove now that each solution of this equation satisfies the relation y)

C, where C is a constant. More precisely, assume there is a solution Y of the

differential equation (10.11) defined on an open interval (a,

such that the point (x, Y(x)) is in for each x in (a, b). We shall prove that

for some constant C. For this purpose we introduce the composite function defined on

(a,

by the equation

By the chain rule, the derivative of g is given by

y)+ = 0, so g’(x) = 0 for each x in (a,

where y = Y(x) and y’ = Y’(x). If y satisfies

then

and, therefore, is constant on (a, b). This proves that every solution y satisfies the equation

y) = C.

Now we may turn this argument around to find a solution of the differential equation. Suppose the equation

defines y as a differentiable function of x, say y Y(x) for x in an interval (a, b), and let g(x) =

Y(x)]. Equation (10.13) implies that is constant on (a, b). Hence, by y) +

= 0, so y is a solution. Therefore, we have proved the following theorem :

THEOREM 10.7. Assume that the

equation

Line integrals

is exact in an open connected set and let be a satisfying

and

-Q

everywhere in S. Then every solution y = Y(x) of (10.14) whose graph lies in

the equation

Y(x)] = C for some constant C. Conversely, the equation

y implicitly as a differentiable function of x, then this function is a solution of the

equation (10.14). The foregoing theorem provides a straightforward method for solving exact differential

equations of the first order. We simply construct a potential function and then write the equation.

y) = C, where C is a constant. Whenever this equation defines y implicitly as a function of the corresponding y satisfies (10.14). Therefore we can use Equation (10.13) as a representation of a one-parameter family of integral curves. Of course, the only admissible values of C are those for which

= C for some , in S.

EXAMPLE 1. Consider the differential equation

Clearing the fractions we may write the equation as

dx +

This is now a special case of (10.14) with

and y) = + Integrating P with respect to and Q with respect toy and comparing results, we find that

y) = +

a potential function is given by the formula

y)=

By Theorem 10.7, each solution of the differential equation satisfies

for some C. This provides an implicit representation of a family of integral curves. In this particular case the equation is quadratic in and can be solved to give an explicit formula for y in terms of and C.

EXAMPLE 2. Consider the first-order differential equation (10.15)

Exercises

Here y) = y and

= 2, this differential equation is not exact. However, if we multiply both sides by

y) = 2x. Since

= a.nd

we obtain an equation that is exact:

+ 2xy dy = 0.

and every solution of (10.16) satisfies the relation

A potential of the vector field

is

y) =

= C. This relation also represents a family of integral curves for Equation (10.15).

The multiplier y which converted (10.15) into an exact equation is called an integrating

fact o r. In general, if multiplication of a first-order equation by a factor y) results in an exact equation, the multiplier

y) is called an integrating factor of the original equation. A differential equation may have more than one integrating factor. For example,

y) = is another integrating factor of (10.15). Some special differ- ential equations for which integrating factors can easily be found are discussed in the following set of exercises.

10.20 Exercises

Show that the differential equations in Exercises through are exact, and in each case find a one-parameter family of integral curves.

6. Show that a linear first-order equation, y’ +

has the integrating factor = Use this to solve the equation.

Let y) be an integrating factor of the differential equation

y) dx +

y ) dy = 0.

Show that

Use this equation to deduce the following rules for finding integrating factors: (a) If

is an integrating factor. (b) If

is a function of x alone, say j”(x), then

is a function ofy alone,

then

is an integrating factor.

Use Exercise 7 to find integrating factors for the following equations, and determine a parameter family of integral curves.

dx + dy } is an integrating factor of the differential equation

If show

exp

y) dy = 0. Find such an integrating factor for each of the following equations and obtain a one-parameter family of integral curves. (a)

tany)dx + dy = 0.

Line integrals

10. The following differential equations have an integrating factor in common. Find such an integrating factor and obtain a one-parameter family of integral curves for each equation.

10.21 Potential functions on convex sets In Theorem 10.6 we proved that the conditions

are necessary for a continuously differentiable vector

,... to be a gradient on an open set S in R”. We then showed, by an example, that these conditions are not always sufficient. In this section we prove that the conditions are sufficient whenever the set S is convex. The proof will make use of the following theorem concerning differentiation under the integral sign.

THEOREM 10.8. Let S be a closed interval in

with

interior and let J = [a, b]

be a closed interval in

Write each point in as (x, t), where S and t J,

Let

be the closed interval S x J in

Assume that is a scalarjeld

on

such that the partial derivative is continuous

wherekisoneof

n.

a scalarjeld on S by the equation

Then the partial derivative exists at each interior point of S and is given by the formula

In other words, we have

Note: This theorem is usually described by saying that we can differentiate under the integral sign.

Proof. Choose any in the interior of S. Since int S is open, there is an r > 0 such that + he, int S for all h satisfying 0 <

is the kth unit coordinate vector in R”. For such h we have

< r. Here

+ he,, t)

t)} dt .

351 Applying the mean value theorem to the integrand we have

Potential functions on convex sets

where z lies on the line segment joining and + he,. Hence (10.17) becomes

The last integral has absolute value not exceeding

where the maximum is taken for all z on the segment joining to + he,, and all in [a, b]. Now we invoke the uniform continuity of

on S (Theorem 9.10) to conclude that for every > 0 there is a > 0 such that this maximum is

a), whenever 0<

Therefore

wlhenever 0 < < 6.

h -a This proves that

t) dt, as required. Now we shall use this theorem to give the following necessary and sufficient condition for

exists and equals

a vector field to be a gradient on a convex set.

vector on an open convex set S in

be a continuously

Then is a gradient on S and only we have (10.18)

n.

Proof. We know, from Theorem 10.6, that the conditions are necessary. To prove sufficiency, we assume (10.18) and construct a potential on S. For simplicity, we assume that S contains the origin. Let

be the integral the line segment from 0 to an arbitrary point in S. As shown earlier in Equation (10.10) we have

dt =

t)

Line integrals

where

t) = f(tx)

. There is a closed n-dimensional subinterval T of S with

empty interior such that satisfies the hypotheses of Theorem 10.8 in T x

where = [0,

Therefore the partial derivative exists for each = 2, . . . , n and can be computed by differentiating under the integral sign,

To compute

we differentiate the dot

. and obtain

where in the last step we used the relation (10.18). Therefore we have

Now let =

. By the chain rule we have

so the last formula for

t) becomes

Integrating this from 0 to 1 we find

dt . We integrate by parts in the last integral to obtain

g(t) dt. Therefore (10.19) becomes

This shows that

= f on S, which completes the proof.