Figure 5.1 The peregrine falcon (Falco peregrinus) feeds primarily on pigeons, doves, mice, and shorebirds.

images

Preview

“Nature laughs at the difficulties of integration.”

Pierre-Simon de Laplace (1749–1827)

Calculus has two parts—differential calculus, the topic of the previous chapters, and integral calculus. At the core of differential calculus is the concept of the instantaneous rate of change of a function. We have seen how this concept can be used to locally approximate functions and to identify maxima and minima. Integral calculus, on the other hand, deals with accumulated change and, thereby, recovering a function from a mathematical description of its instantaneous rate of change. This recovery process, interestingly enough, is related to the concept of finding the area under a curve.

Using integral calculus, we can calculate the velocity of a stooping peregrine falcon (see Figure 5.1) as it dives toward Earth at great speed to catch prey; calculate the blossom date of a tree as a function of anticipated temperature patterns, so that an orchard can be stocked with bees to facilitate pollination; or estimate the amount of a drug in the bloodstream of a patient connected to an IV.

A systematic method for estimating the area under the curve was devised by Riemann, one of the great mathematicians of the nineteenth century. This method is commonly known as the Riemann sum and yields in the limit an object called the definite integral. The fathers of calculus, Newton and Leibniz, proved a connection between the problem of finding antiderivatives and finding areas under a curve. This connection, the fundamental theorem of calculus, which is presented in Section 5.4, helps make calculus one of the most powerful mathematical tools for understanding biological and physical processes.

In Sections 5.5 through 5.7, we provide a short apprenticeship in various techniques used to compute and approximate integrals. Armed with these techniques, the chapter concludes with applications to cardiac output, survival and renewal equations, and the scientific notion of work.

5.1 Antiderivatives

Many mathematical operations have an inverse. For example, to undo the addition of b to a we subtract b: a + bb = a. To undo division of a by b we multiply by b: images b = a. To undo exponentiation, we take logarithms: ln ea = a. The process of differentiation can be undone by a process called antidifferentiation.

To motivate antidifferentiation, consider how long it takes an organism to develop when the rate of development depends on environmental factors such as heat, light, and humidity. For example, the developmental rate of plants and insects, which lack internal thermal regulation mechanisms, depend critically on ambient temperature. For ambient temperatures within a range defined by developmental thresholds, a plant's or insect's organismal developmental rate can often be approximated by an increasing linear function of temperature. For example, Eileen Cullen, a doctoral student at the University of California, collected data shown in Table 5.1 on the developmental rate of a particular species of stinkbug (Figure 5.2) reared in the laboratory.

images

Figure 5.2 A green stinkbug

Table 5.1 Developmental rates of stinkbugs

Temperature (°F) Developmental rate (1/days)
64.4 1/89
69.8 1/58
80.6 1/37
89.6 1/25

We see from this table that a stinkbug kept at 64.4°F completes images of its development in one day and all of its development in eighty-nine days. Performing linear regression on these data (i.e., to find the “best-fitting” line as discussed in Section 4.3) yields

images

where T is temperature in degrees Fahrenheit. This relationship is illustrated in Figure 5.3. If T(x) is the temperature at time x and F(x) denotes the fraction of development completed by the stinkbug at time x, then the preceding equation yields the rate at which F(x) changes with time; that is,

images

images

Figure 5.3 Graph of the developmental rate of stinkbugs. The red dots represent the actual data, and the line is the best-fitting line.

Thus, if we know T(x) and want to know how long it takes the stinkbug to complete development, we need to “solve” for F(x). More generally, if we are given that f(x) is the developmental rate at time x, then F(x) must satisfy

images

Understanding solutions of this equation is the main goal of this section.

Antiderivative

Given a function f, an antiderivative F of f is a function F that satisfies

images

For example, x3 is an antiderivative of 3x2 since imagesx3 = 3x2. Is x3 the only antiderivative of 3x2? The answer is no. For example x3, x3 + 1, and x3 + π all have the same derivative 3x2. Consequently, all are antiderivatives of 3x2. Luckily, all antiderivatives of a function are related. Suppose F(x) and G(x) are antiderivatives of f(x) on some interval. Since F′(x) = f(x) = G′(x), the function

images

has derivative

images

on this interval. What functions have a derivative equal to zero on an interval? The mean value theorem implies only the constant function. Hence, there must be a constant C such that F(x) = G(x) + C, and we have the following result.

General Form of an Antiderivative

If F is an antiderivative of f on an interval I, then every antiderivative of f on I has the form

images

where C is a constant. For this reason, we call F(x) + C the general form of the antiderivative.

Because of this general form, finding the general form of an antiderivative amounts to finding an antiderivative of f and adding an arbitrary constant C.

Example 1 Finding general forms of antiderivatives

Find the general forms of the antiderivatives of

images

Solution

  1. Recall that imagesex = ex. Thus, the general form of the antiderivative is ex + C.
  2. Recall that imagessin x = cos x. Hence, the general form of the antiderivative is sin x + C.
  3. Recall that imagesx6 = 6x5. This is not quite what we want, because we are off by a factor of 6. If we divide both sides of the equation by 6, then

    images

    Thus, the general form of the antiderivative of x5 is images + C.

Warning! What we did in part c, namely, divide by 6 because we were off by a factor of 6, worked only because 6 is a constant. It doesn't work in general. For example, suppose we wanted to find an antiderivative of ex2. It would be incorrect to argue that since imagesex2 = 2xex2, we are off by a factor of 2x and the antiderivative is imagesex2. Indeed, imagesex2 does not equal ex2 as you should verify for yourself.

Example 2 Antiderivative of cos(ax)

Find the general form of the antiderivative for cos(ax), where a ≠ 0.

Solution We know images sin(ax) = a cos(ax), but this is not quite what we want, since we are off by a factor of a. If we divide both sides by a, then

images

Thus, the general form of the antiderivative of cos(ax) is images sin(ax) + C.

Corresponding to the many rules of differentiation are rules of antidifferentiation. For instance, if F(x) and G(x) are antiderivatives of f(x) and g(x), respectively, then H(x) = F(x) + G(x) is an antiderivative of h(x) = f(x) + g(x). Indeed, since the derivative of a sum is the sum of the derivatives, we obtain

images

In a similar manner, we can show that antiderivatives have the following properties.

Properties of Antiderivatives

Let F(x) and G(x) be antiderivatives of f(x) and g(x), respectively.

Addition F(x) + G(x) is an antiderivative of f(x) + g(x).

Subtraction F(x) − G(x) is an antiderivative of f(x) − g(x).

Scalar multiplication For any constant c, cF(x) is an antiderivative of cf(x).

Combining the antidifferentiation properties and formulas allows us to compute even more antiderivatives, as the following examples illustrate.

Example 3 Using antiderivative rules

Find the general antiderivative of 3x2 + 3x + 7.

Solution Since an antiderivative of a sum is a sum of antiderivatives, an antiderivative of 3x2 + 3x + 7 is the sum of antiderivatives of 3x2, 3x, and 7. Antiderivatives of 3x2, 3x, and 7 are x3, imagesx2, and 7x. Hence, an antiderivative of 3x2 + 3x + 7 is x3 + imagesx2 + 7x, and the general form of the antiderivative is x3 + imagesx2 + 7x + C where C is an arbitrary constant.

To find a particular antiderivative F(x) of f(x) on an interval, we need to know a value of F(x) at a particular value of x to determine the particular value of the arbitrary constant C. If we have this information, then finding the antiderivative is known as an initial value problem.

Example 4 An initial value problem

Find F(x) such that F(2) = 1 and F′(x) = 3x2 + 3x + 7.

Solution From Example 3, the general form of the antiderivative is F(x) = x3 + imagesx2 + 7x + C. To solve for C, we solve the equation F(2) = 1 as follows:

images

This implies 28 + C = 1 or C = −27. Thus,

images

Example 5 Stinkbug development

Consider the development of the stinkbug from egg to adult, as summarized in Table 5.1. Suppose the stinkbugs are reared under a temperature T(x) that sinusoidally oscillates between 60° and 80°F each day

images

where x is measured in days. Solve the following:

  1. Assuming F(0) = 0, find the function F(x) corresponding to the proportion of development completed after x days.
  2. Use technology to find the number of days for development to be completed; that is, find x such that F(x) = 1.

Solution

  1. Recall that we have F′(x) = −0.06075 + 0.00112T(x). Substituting T(x) into the expression for F′(x) yields

    images

    Since an antiderivative of 0.01765 is 0.01765x and an antiderivative of 0.0112 cos(2πx) is images sin(2πx), it follows that

    images

    To find C, we solve

    images

    which implies C = 0. Hence,

    images

  2. Plotting F(x) as shown below suggests that development is completed in just under fifty-seven days.

    images

Differential equations and slope fields

An equation that involves derivatives is called a differential equation. Consider a function y = F(x). Then any equation of the form y′ = f(x), or in Leibniz's notation

images

is a differential equation, and solving for the antiderivative y = F(x) of f(x) (i.e., F′(x) = f(x)) corresponds to solving this differential equation.

In the next chapter, we discuss differential equations in greater detail. Here, we introduce the topic by considering a physiological phenomenon known as the Weber-Fechner law. This law describes the expected response of an animal or human subject to a stimulus, such as light or sound. More particularly, the Weber-Fechner law in physiological psychology asserts that when a subject is exposed to a stimulus, y, the rate of change of the response x with respect to y is inversely proportional to x. This statement can be written mathematically as

images

where k a positive constant to be determined through experiment. One can interpret this equation as saying if the stimulus x is small, then small changes in stimulus cause large changes in the response. Alternatively, if stimulus x is large, then small changes in the stimulus do not change the response much.

Example 6 Solving the Weber-Fechner differential equation

Find the solution to the Weber-Fechner equation

images

assuming that a threshold stimulus, x0 > 0, is the lowest level for which a response can be detected. In other words, find y(x) subject to the initial condition y(x0) = 0.

Solution This problem requires us to find a function y(x) such that y′ = k/x and y(x0) = 0. Taking the general antiderivative of k/x with respect to x yields

images

where C is a constant. Since y(x0) = 0, solving

images

for C yields

images

Hence, we obtain

images

Equivalently,

images

A particular example of the Weber-Fechner law is the logarithmic decibel scale for measuring the intensity of sound.

For some functions it is impossible to come up with an explicit expression for the antiderivative: for example, f(x) = ex2 and f(x) = sin x2. In such cases, numerical or graphical methods can be used. One graphical approach involves slope fields, where we can use the fact that the slope of a function y = F(x) at any point (x, y) on its graph is given by the derivative F′(x). We exploit this fact to obtain a “picture” of all slopes F′(x) on the (x, y)-plane, as illustrated in the next example.

Example 7 Antiderivatives with slope fields

Use technology to sketch the slope field for the equation

images

for 0 ≤ x ≤ 3 and −2 ≤ y ≤ 2. Sketch by hand antiderivatives F(x) satisfying F(0) = 0 and F(0) = −2, respectively.

Solution Using technology, we draw small line segments of slope F′(x) = sin(x2) at regular intervals in the xy plane, for 0 ≤ x ≤ 3 and −2 ≤ y ≤ 2, to obtain Figure 5.4a. The collection of these line segments is called a slope field or direction field of the function f.

images

Figure 5.4 Slope field with particular solutions

We see that all the line segments anchored at points (0, y) are horizontal lines, because F′(0) = sin(0) = 0. These horizontal line segments correspond to tangent lines of the antiderivatives at x = 0. Line segments anchored at other points (x, y) have slope F′(x) = sin x2, which varies between −1 and 1, and again are independent on the value of y. Sketching an antiderivative F(x) satisfying F(0) = 0 corresponds to sketching a curve that passes through the point (0, 0) and remains tangent to the line segments in Figure 5.4b. This sketch yields the upper curve in the right panel of the Figure 5.4b. The lower curve corresponds to the antiderivative F(x) that satisfies F(0) = −2, since this curve passes through the point (0, −2). Notice that the graphs of each of these antiderivatives are vertical translations of one another, because the slopes for each value of x do not depend on the location of y.

Rectilinear motion

We can use antiderivatives to understand the motion of an object along a straight line. We have previously defined velocity v(t) at time t to be the rate of change of position s(t) of an object—that is, s′(t) = v(t), and acceleration a(t) to be the rate of change of velocity v(t) of an object—that is, v′(t) = a(t). Thus, it follows that position is the antiderivative of velocity which, in turn, is the antiderivative of acceleration.

These definitions allow us to study the stooping behavior of the peregrine falcon shown in Figure 5.1. The peregrine falcon is arguably the fastest animal in the world. This long-winged raptor favors direct pursuit of other birds, and its level speed exceeds the speed of most birds upon which it preys. Peregrines gain speed by launching attacks from high and then stooping (steep diving with feet back against the tail and wings close to the body) to attain speeds of well over 300 km/h.

Example 8 Stooping peregrines

Assume that the peregrine falcon's downward acceleration is due to gravity alone, which is 9.8m/s2, and that there is no air resistance. Determine how far a peregrine falcon would have to free-fall to achieve a speed of 300 km/h.

Solution Let v(t) denote the downward velocity at t seconds after a peregrine falcon has begun its stoop. Assuming acceleration is due purely to gravity, we have

images

To solve for v, we antidifferentiate to obtain v(t) = 9.8t + C where C is a constant. Since the peregrine has no downward velocity at the beginning of its stoop, we have v(0) = 0. Hence, 0 = v(0) = C and

images

To find the position s(t) of the falcon at time t, we have

images

Here, s(t) describes the vertical distance (in meters) from the initial position of the falcon to its position at time t. Antidifferentiating yields s(t) = 4.9t2 + C where C is some constant. Since s(0) = 0, we obtain 0 = s(0) = C and

images

Next, we need to determine how many seconds the peregrine falcon needs to fall to achieve a speed of 300 km/h. By converting 300 km/h to meters/second we obtain

images

Thus, to find the desired time, we solve

images

for t to obtain t ≈ 8.5 seconds. Hence, the distance fallen to achieve a speed of 300 km/h is approximately

images

Thus, the peregrine falcon needs to free-fall about 350 meters to attain a speed of 300 km/h if it relies purely on the force of gravity.

images

Recently, scientists have accurately measured speeds achieved by peregrine falcons during stooping. One falcon was logged by radar at 183 km/h ≈ (114 mph) after a dive of 305 m ≈ (1,000 ft). This is considerably slower than the 300 km/h our current model would predict. One reason for this discrepancy is that we ignored air resistance in our calculations. This shortcoming can be addressed by formulating a differential equation model that includes the effects of air resistance.

PROBLEM SET 5.1

Level 1 DRILL PROBLEMS

Find the general antiderivative of the functions f shown in Problems 1 to 22.

images

17. f(x) = 3ex

18. f(θ) = sec2θ for −π/2 < x < π/2

19. f(x) = x3/2 + x1/2 + x−1 for x > 0

20. f(u) = u3 − 2u + images

21. f(u) = 6u + 3 cos u

22. f(x) = 5x − 4 sin x

Find the antiderivative F(x) of the functions shown in Problems 23 to 28 satisfying the indicated initial condition.

23. f(x) = 2 with F(0) = 1

24. f(x) = 4 with F(1) = −1

25. f(x) = 2x + 3 with F(−3) = 0

26. f(x) = 4 − 5x with F(0) = 4

27. f(x) = 6x4 with F(1) = −2

28. f(x) = 2x−4 for x > 0 with F(2) = 0

29. a. If F′(x) = 1 − 4x, find F so that F(1) = 0.

b. Sketch the graphs of y = F(x), y = F(x) − 2, and y = F(x) + 4.

c. Find a constant C so that the largest value of G(x) = F(x) + C is 0.

30. a. If F′(x) = 2x − 1, find F so that F(2) = 0.

b. Sketch the graphs of y = F(x), y = F(x) − 2, and y = F(x) + 4.

c. Find a constant C so that the smallest value of G(x) = F(x) + C is 0.

The slope F′(x) at each point on a graph is given in Problems 31 to 34 along with one point (x0, y0) on the graph. Use this information to find F graphically.

images

35. Sketch a slope field for

images

for −2 ≤ x ≤ 2 and 0 ≤ y ≤ 5. Over this slope field, sketch the antiderivative of F(x) of x which satisfies F(0) = 1.

36. Sketch a slope field for

images

for −5 ≤ x ≤ 5 and −5 ≤ y ≤ 5. Over this slope field, sketch the antiderivative F(x) of x which satisfies F(0) = 0.

37. Sketch a slope field for

images

for −π ≤ x ≤ π and −2 ≤ y ≤ 2. Over this slope field, sketch the antiderivative F(x) of x which satisfies F(0) = 1.

38. Sketch a slope field for

images

for −1 ≤ x ≤ 1 and −2 ≤ y ≤ 2. Over this slope field, sketch the antiderivative F(x) of x which satisfies F(−1) = 0.

39. Find the general antiderivative of sin(ax) where a ≠ 0.

40. Find the general antiderivative of ekx where k ≠ 0.

Level 2 APPLIED AND THEORY PROBLEMS

41. As discussed in Example 5, the developmental rate of a stinkbug as a function of temperature T is −0.06075 + 0.00112T. Assume that the temperature of a typical spring in Davis, California, x days after the start of the stinkbug development period is adequately modeled by the function

images

  1. Find the function F(x) describing the amount of development completed by day x assuming that F(0) = 0.
  2. Estimate at what time a stinkbug has completed development.

42. Recall that the developmental rate of a stinkbug as a function of temperature T is −0.06075 + 0.00112T. Assume that the temperature of an atypical day in Davis, California, x days after the start of the stinkbug development period is adequately modeled by the function

images

  1. Find the function F(x) describing the amount of development completed by day x assuming that F(0) = 0.
  2. Estimate at what time a stinkbug has completed development.

43. Entomologists Godfrey and Anderson* studied the developmental rates of the hydrilla tuber weevil, which is a species that consumes a weed found in ponds and waterways. As illustrated in Figure 5.5, the developmental rate as a function of temperature (in degrees Celsius) is given by

images

images

Figure 5.5 Developmental rate as a function of temperature

  1. Suppose that the temperature in C° is given by the function T(t) = 30 + 10 sin(2πt) where t is measured in days. Find the fraction of development F(t) that has been completed at time t for eggs laid at time 0.
  2. Estimate how many days it takes the weevil to develop to adulthood.

44. Assume that the temperature in Problem 43 is given by the function T(t) = 50 + images. Using the developmental rate for the tuber weevil presented in that problem, estimate how many days it takes the weevil to complete half of its development.

45. A peregrine falcon stoops for 305 meters. Assuming a constant acceleration of 9.8 m/s2, find its speed at the end of the stoop.

46. Suppose a food package is dropped out of a balloon that is 100 ft above the ground and ascending at a rate of 10 ft/s. Determine how long it takes the package to hit the ground, assuming a constant gravitational acceleration of 9.8 m/s2.

47. Apollo 15 astronaut David Scott dropped a hammer and a feather on the moon to demonstrate that in a vacuum all objects fall at the same rate. He dropped both items from a height of approximately 4 ft. How long did it take each object to hit the ground? (Acceleration on the moon due to gravity is −5.2ft/s2.) How long would it take for a hammer to hit the ground on Earth if dropped from a height of 4 ft? (Gravitational acceleration on Earth is −32 ft/s2.)

48. Assume the brakes of a certain automobile produce a constant deceleration of 22 ft/s2. If the car is traveling at 60 mi/h (88 ft/s) when the brakes are applied, how far will it travel before coming to a complete stop?

49. It is estimated that t months from now, the population of Ferndale, California, will be changing at the rate of 4 + 5t2/3 people per month. If the current population is 2,000, what will the population be eight months from now?

50. A hypothetical study of a community suggests that t years from now the level of carbon monoxide in the air will be changing at the rate of 0.1t + 0.1 ppm/year. If the current level of carbon monoxide in the air is 3.4 ppm, what will be the level three years from now?

51. One of Poiseuille's laws for the flow of blood in an artery says that if v(r) is the velocity of flow r centimeters from the central axis of the artery, then the velocity decreases at a rate proportional to r. That is, v′(r) = ar where a is a negative constant. Find an expression for v(r) assuming that v(R) = 0, where R is the radius of the artery.

52. Suppose that a silviculturist finds that a certain type of tree grows in such a way that its height h(t)t years after planting is changing at the rate of

images

If the tree was two feet tall when it was planted, how tall will it be in twenty-seven years?

53. Suppose that a woman driving a sports car down a straight road at 60 mi/h (88 ft/s) sees a cow start to cross the road 200 feet ahead. She takes 0.7 seconds to react to the situation before hitting the brakes, which decelerates the car at the rate of 28 ft/s2. Does she stop in time to avoid hitting the cow?

54. A hypothetical population, N, grows in such a way that at time t (years), the growth rate is given by

images

where N(t) is measured in thousands of individuals and N(0) = 5.

  1. Find N(t).
  2. What is the minimum population size? When does it occur?

5.2 Accumulated Change and Area under a Curve

In this section, we deal with the problem of finding accumulated change, which can be interpreted as the area under a curve. The early Greeks, particularly Archimedes (287–212 BC), estimated the areas of geometrical objects using the “method of exhaustion,” a precursor to integral calculus. They found increasingly better approximations by filling in areas with increasingly smaller elements of known area (much as we do later in this section in taking Riemann sums).

In elementary school you learned formulas to find areas of squares, triangles, and other polygons. You also are familiar with the formula for the area of a circle with radius r: A = πr2. The Egyptians were the first to use this formula more than 5,000 years ago, but the Greeks derived the area of a circle by drawing inscribed polygons or circumscribed polygons, as shown in Figure 5.6, and then using triangles to find the area of those polygons as an approximation. This method, called the method of exhaustion, involves finding the area of a circle by inscribing polygons with increasing numbers of sides (Archimedes stopped at a ninety-six-sided polygon). The area of the circle is the limit of the areas of the inscribed polygons as the number of polygonal sides increases.

images

Figure 5.6 Using the limit of a sequence of inscribed polygons to find the area of a circle

images

Figure 5.7 Area under a curve y = f(x)

In this section, we focus on estimating the area under a curve y = f(x) over an interval a to b. As illustrated in Figure 5.7, this means estimating the area defined by the region bounded by the curves y = f(x) (with f(x) ≥ 0 on [a, b]), x = a, x = b, and y = 0. Similar to the method of exhaustion, we find these areas by approximating them with collections of finer and finer rectangles.

To motivate finding area under a curve, we show that area under a curve corresponds to accumulated change. We do this in the context of two biological processes: physiological time for insects and plants and disease incidence as studied by epidemiologists.

Physiological time: degree-days

Plants and insects often require a certain amount of heat to develop from one stage in their life cycle to another stage in their life cycle. This measure of accumulated heat is known as physiological time, and the units used are called degree-days—the accumulated product of time and temperature between the organism's lower and upper developmental thresholds. The lower developmental threshold is the temperature below which the insect or plant cannot develop, while the upper development threshold is the temperature above which it cannot develop.

To simplify the presentation, we initially assume that the temperature remains between the lower and upper developmental thresholds. The more general case is considered later. One degree-day is one day (24 hours) with the temperature one degree above the lower developmental threshold. For example, if the lower developmental threshold of the organism is 47°F and the temperature remains at 48°F for one day or 47.2°F for five days, then in each case, one degree-day is accumulated (i.e., 1 × (48 − 47) = 5 × (47.2 − 47) = 1).

The concept of degree-days is used widely in agriculture and developmental biology. In agricultural publications we may come across the following types of statements: “Around 1539 degree-days are required for sweet corn to ripen, assuming a lower development threshold of 50°F”; or, “corn earworms (pests of corn) have a lower developmental threshold of 54.7°F and require 760 degree-days to develop from egg to adult”. These statements allow us to estimate the time it takes sweet corn to mature or corn earworms to develop in geographical regions with different temperature profiles, as well as to anticipate how these times may change in response to global warming.

Example 1 Degree-days under constant temperature

According the University of California at Davis Integrated Pest Management Program website, the lower developmental threshold of Thompson seedless grapevines is 50°F, and this variety of grape requires approximately 3,000 degree-days for its fruit to mature. If the temperature were to remain constant at 70°F, how long would it take for the fruit to mature?

Solution The amount of degree-days accumulated in x days is

images

Solving 20 x = 3,000 yields x = 150 days. Therefore, it would take 150 days for the grapes to mature. Notice that this answer can be interpreted as the following shaded rectangular area:

images

Unlike the preceding example, temperatures in fields vary over time. Consequently, computing the accumulation of degree-days, as we shall see, requires finding the area of an appropriate region defined by the temperature curve and the lower developmental threshold.

For example, the temperature in Lincoln, Nebraska, on June 23, 2006 is given by the function f(t)°F, illustrated in Figure 5.8, where t going from 0 to 1 represents one day from midnight to midnight. If an organism of interest (say, sweet corn) has a lower developmental threshold of 50°F, then, as we explore in the next example, the total accumulated degree-days is given by the area between the curves y = f(t) and y = 50 from t = 0 to t = 1. This area corresponds to the shaded area in Figure 5.8.

images

Figure 5.8 Temperature for Lincoln, Nebraska, on June 23, 2006. The shaded area corresponds to the accumulated degree-days for an organism with a lower developmental threshold of 50°F.

Example 2 Sweet corn in Nebraska

Estimate the accumulation of degree-days for sweet corn in Lincoln, Nebraska, on June 23, 2006 using the data in Table 5.2.

Table 5.2 Temperature in Lincoln, Nebraska, starting at midnight of June 23, 2006 and reported at two-hour intervals

images

Solution To approximate the number of degree-days that have accumulated, break up the day into two-hour intervals (i.e., intervals of width images day). Within each interval, assume that the temperature is relatively constant. Then, accumulated degree-days within the first interval [0, 1/12] is given by

images

This quantity simply corresponds to the area of a rectangle with height 15.8 and width images days as illustrated in Figure 5.9. The total 286.6 in Table 5.2, when divided by 12 yields images ≈ 23.9 degree-days, which corresponds to the sum of the areas of rectangles depicted in Figure 5.9. This sum of areas is an approximation for the shaded area in Figure 5.8.

images

Figure 5.9 Accumulated degree-days approximated by the area of rectangles whose width is images days

The Bombay plague epidemic

In epidemiology, scientists keep track of various rates associated with disease, including the incidence rate, which measures the number of new disease cases per unit of time (e.g., day or week), and the mortality rate, which reports the number of deaths due to the disease per unit of time. For instance, during the outbreak of the plague in Bombay in 1905–1906, the weekly mortality rates due to the plague were recorded, and the values obtained are plotted in Figure 5.10.

images

Figure 5.10 Mortality rate from plague in Bombay (now called Mumbai) from December 17, 1905 to July 21, 1906. Data shown in red and the fitted function in blue.

In a landmark paper,* two mathematicians, W. O. Kermack and A. G. McKendrick, showed that these data could be reasonably well fitted by the function

images

where t is measured in weeks and the hyperbolic secant function sech x equals images. If we want to estimate the total number of deaths using this function, what do we need to compute? Consider a small interval of time, from t to t + Δt. Since the mortality rate over this interval is given approximately by f(t), the number of deaths over this time interval is approximately f(tt, that is, the area of a rectangle of width Δt and height f(t). Notice how the units work out in this product: f(t) has units of deaths/week and Δt has units of weeks. The product f(tt has units of deaths. These arguments suggest that the area under the curve f(t) from t = 0 to t = 30 equals the total number of deaths, as discussed in the next example. Since we know the actual number of deaths—we simply add up all the weekly data—the next example examines how the area under f(t) approximates the actual number of deaths.

Example 3 Mortality due to the plague

Use the function f(t) = 890 sec h2(0.2t − 3.4) to approximate the total number of deaths in Bombay from t = 0 to t = 30 using intervals of five weeks.

Solution Begin by breaking the interval from t = 0 to t = 30 into six subintervals of length 5, as shown in Figure 5.11. For the mortality rate in each interval, we can evaluate f(t) at the right endpoint of each interval. This yields the following table (entries rounded to one decimal place):

images

Summing up the deaths yields 8,847 deaths, which is approximately 2% short of the actual 9,043 recorded number of deaths. This approximation is illustrated in Figure 5.11.

images

Figure 5.11 Using the data in Figure 5.10 to approximate the number of deaths

In Figure 5.11 notice that when the curve is on the rise, as in the first three rectangles, the area is overestimated (the green area above the curve), and when the curve is on the decline, the area is underestimated (the white area below the curve). This is a result of the height of the rectangles being defined by the value of the function on the right side of each interval. The reverse would be true if the height of the rectangles were defined by the value of the function on the left side of each interval.

The area problem

The previous examples illustrate the importance of finding areas under curves. These examples also show that we can approximate areas by approximating the region with rectangles, computing the area of each rectangle, and summing up the areas. This observation is the key that unlocks the area problem. We pursue this approach further in the following example.

Example 4 Estimating the area under a curve

Consider the function f(x) = x2 over the interval [0, 1]. Use rectangles to find upper and lower bounds for the area under x2, above the y axis, between x = 0 and x = 1.

Solution Let A denote the area under y = f(x), above y = 0, and between the lines x = 0 and x = 1, as shown in Figure 5.12.

Now, we find the area A by taking successive approximations. Notice that the largest value of x2 on the interval [0, 1] is 1 at x = 1. Hence, the region under x2 is contained in a rectangle of height 1 and width 1. Thus, A < 1 × 1 = 1. On the other hand, A is clearly greater than 0. To obtain a better estimate, subdivide the interval [0, 1] into two subintervals [0, 1/2] and [1/2, 1], each with width Δx = 1/2, as shown 0.2 in Figure 5.13.

images

Figure 5.12 Area A under y = x2 on [0, 1]

images

Figure 5.13 Left and right sum approximations of the area under y = x2

The greatest values that x2 takes on these subintervals are f(1/2) = 1/4 and f(1) = 1. Hence, the two rectangles over the intervals [0, 1/2] and [1/2, 1] with heights, 1/4 and 1, respectively, enclose our region. Therefore, A < images. Alternatively, the minimum values of x2 on [0, 1/2] and [1/2, 1] are 0 and 1/4, respectively. Therefore, A > 0 × images.

Subdividing the interval once improved our estimates, so more subdivisions should also improve our estimates. Suppose we divide the interval into n subintervals images. Since x2 is an increasing function on the interval [0, 1], the maximum values of f(x) = x2 on these subintervals are images. The area of the n rectangles determined by these heights is given by

images

which is greater than A. Since the minimum values of x2 on these subintervals of width images, A is greater than

images

Thus,

images

This relationship is illustrated with n = 4 in Figure 5.14 where

images

images

Figure 5.14 Estimating the area of A using four subintervals

Computing these estimates by hand for large n is tedious. Using technology, we calculated the entries in the following table to a precision of ten decimal places, but zeros are included only when necessary.

images

The sums suggest that as n becomes large, Ln and Rn both converge to images.

Example 4 suggests that area under x2 over the interval [0, 1] is images. But how can we really be sure that these numbers converge to images? We address this question in the next example.

Example 5 Finding the exact area under x2 on the interval [0, 1]

Use the formula (which can be proved inductively)

images

to prove that

images

Solution We have that

images

Thus,

images

Similarly (see Problem 19 in Problem Set 5.2), it can be shown that images Ln = images. Since LnARn for all n ≥ 1, it follows that

images

Examples 4 and 5 provide the core idea of how to define area under a nonnegative function y = f(x) from x = a to x = b. First, we divide the interval [a, b] into n equally spaced subintervals of width Δx = images. Let

images

To approximate the height of f over a subinterval [ai, ai+1], choose a point xi on the interval [ai, ai+1]. The points xi are called sample points. In our examples, we chose left or right endpoints as our sample points, but we could have picked any point in each interval. The height of f over [ai, ai+1] is approximately f(xi). The area of f over [ai, ai+1] is approximately f(xi) Δx. Adding all these rectangular areas up yields

images

This sum is known as a Riemann sum after the brilliant mathematician Georg Friedrich Bernhard Riemann (1826–1866; see the images in Problem Set 5.2). Now we write this sum succinctly, using the summation notation presented in the regression subsection of Section 4.3.

Riemann Sum

Suppose a continuous function f is defined on the interval [a, b]. If the interval is divided into n subintervals so that Δx = images and

images

then a Riemann sum associated with f is the sum

images

where xi is any point we chose to select in the interval [ai−1, ai].

We have seen how area can be approximated by a Riemann sum and how this approximation improves as n become large, approaching the true area as n → ∞. Therefore, we write

images

We cannot know that the method really works unless we have a theorem that tells us that a limit exists and that this limit is independent of the way we choose the sample points in the subintervals. For continuous functions such a theorem does exist, but its proof is a topic for a course in real analysis. (The “real” refers to real-valued functions in contrast to the “complex” of complex numbers and complex-valued functions.)

Theorem 5.1 Limit of a Riemann sum theorem

If f(x) is continuous on [a, b], then

images

exists and is independent of the choice of sample points xi.

In Problem Set 5.2, some of the problems require one or more of the following summation formulas. These formulas can be verified using mathematical induction.

Summation Formulas

The following formulas can be verified using mathematical induction. You may use these formulas to find certain Riemann sums.

images

PROBLEM SET 5.2

Level 1 DRILL PROBLEMS

First sketch the region under the graph of y = f(x) on the interval [a, b] in Problems 1 to 12. Then approximate the area of each region by using right endpoints and the formula

images

for Δx = images and the indicated values of n.

1. f(x) = 2x + 1 on [0, 1] for n = 4

2. f(x) = 4x + 1 on [0, 1] for n = 8

3. f(x) = x2 on [0, 2] for n = 4

4. f(x) = x2 on [0, 2] for n = 6

5. f(x) = x3 on [1, 3] for n = 4

6. f(x) = 4x2 + 2 on [0, 1] for n = 4

7. f(x) = x2 + x3 on [0, 1] for n = 4

8. f(x) = ex on [0, 1] for n = 4

9. f(x) = x−1 on [1, 2] for n = 4

10. f(x) = images on [1, 4] for n = 4

11. f(x) = cos x on images for n = 4

12. f(x) = x + sin x on images for n = 3

Use a calculator to estimate the area under the curve y = f(x) on each interval given in Problems 13 to 18 as a sum of ten terms evaluated at right endpoints.

13. f(x) = 4x on [0, 1]

14. f(x) = x2 on [0, 4]

15. f(x) = cos x on images

16. f(x) = x + sin x on images

17. f(x) = ln(x2 + 1) on [0, 3]

18. f(x) = e−3x2 on [0, 1]

Use a summation formula in Problems 19 to 24.

19. Prove that

images

for Ln as defined in Example 4.

20. Use Riemann sums and left endpoints to prove that the area under y = x from x = 0 to x = 2 equals 2.

21. Use Riemann sums and right endpoints to prove that the area under y = x from x = 0 to x = 4 equals 8.

22. Use Riemann sums and right endpoints to prove that the area under y = x3 from x = 0 to x = 4 is 64.

23. Use Riemann sums and left endpoints to prove that the area under y = x3 from x = 0 to x = 2 is 4.

24. Use Riemann sums and right endpoints to prove that the area under y = x + 3x2 from x = 0 to x = 2 is 10.

Level 2 APPLIED AND THEORY PROBLEMS

25. The lower developmental threshold of sweet corn is 50°F and requires 1,587 degree-days for maturing. If the temperature were to remain a constant 75°F, how long would it take for the corn to mature?

26. The pistachio has a lower developmental threshold of 50°F and requires 1,197 degree-days for shell hardening. If the temperature were to remain a constant 72°F, how long would it take for the pistachio's shell to harden?

27. The black turtle bean has a lower developmental threshold of 41°F and requires 1,365.5 degree-days for 50% anthesis (i.e., until 50% of all the flowers have blossomed). If the temperature were to remain a constant 68.5°F, how long would it take to reach the required 50% anthesis?

28. Estimate mortality due to the plague in Bombay (Mumbai) by approximating the region under

images

deaths per week from t = 0 to t = 30 with rectangles of width 15 weeks. Would you expect your answer to be more or less accurate than the result of Example 3?

29. Estimate mortality due to the plague by approximating the region under

images

deaths per week from t = 0 to t = 30 with rectangles of width three weeks and height given by their right endpoints. Would you expect your answer to be more of less accurate than the result of Example 3?

30. The weekly rate of cases of influenza A (strain unknown) studied by WHO/NREVSS during the 2003–2004 season is plotted in Figure 5.15. Estimate the total number of cases (i.e., the area under the curve) over the interval [40, 56] using the right endpoints of two-week intervals. Sketch the corresponding rectangles in the figure.

images

Figure 5.15 Weekly rate of cases of influenza A

31. Repeat Problem 30 using left endpoints.

32. The daily maximum (max) and minimum (min) temperatures at Westmoreland, California, for the period from April 1 to 9 are given in Table 5.3.

Table 5.3 Early morning minimum and mid-afternoon maximum temperatures at Westmoreland, California, for nine days (minimum on April 10 included for later use)

images

The lower developmental threshold for cotton is k = 60°F. If Mi and mi are used to denote the max and min temperatures on day i, respectively, then estimate the degree-day accumulation over the nine-day period using this formula:

degree-days accumulated on day i

images

Note that the logic behind this formula is that on days for which kmi we use the average temperature above the threshold, and on days for which mi < k < Mi we estimate that the temperature is above threshold for a proportion images of the day, and then use the average images as applying over this period of time to obtain the middle expression in the equation.

33. The lower developmental threshold for the elm leaf beetle is 52°F. Use the formula in Problem 32 to estimate the degree-day accumulation for the elm leaf beetle in Stockton, California, over a two-week period of time using the data listed in Table 5.4.

Table 5.4 Early morning minimum and mid-afternoon maximum temperatures at Stockton, California (fire station # 4) for 14 days (minimum on March 15 included for later use)

images

34. In the figure that follows, mi and mi+1 are the minimum daily temperatures on consecutive days i and (i + 1), respectively, which are assumed to be exactly one day apart. Mi is the maximum daily temperature on day i and is assumed to occur exactly half way between the two minimum temperatures. The parameter k is the lower developmental threshold for a particular species.

images

If these values satisfy Mi > mi > k and k > mi+1, then show that the degree-day accumulation over the one-day period, as depicted by the green area in the figure, is given by the expression

images

Alternatively, if Mi > mi > k and Mi > mi+1 > k then show that the degree-day accumulation over day i is given by

images

35. The obliquebanded leafroller, is an agricultural pest with a lower developmental threshold of 43°F. Estimate the accumulated degree-days for this insect growing in Westmoreland, California, over the period from the morning of April 2nd to the morning of April 3rd using the data in Table 5.3 (see Problem 32), and the formula presented in Problem 34.

36. The codling moth, is an agricultural pest with a lower developmental threshold of 50°F. Estimate the accumulated degree-days for this insect growing in Stockton, California, over the period from the morning of April 1st to the morning of April 2nd using the data in Table 5.3 (see Problem 32), and the formula presented in Problem 34.

37. Assume the temperature in degrees Fahrenheit is given by

images

where t is time in days. Assume the lower development threshold is 40°F and estimate the number of degree-days that accumulate from t = 0 to t = 10 days using time intervals of width two.

38. Use the temperature variation model in Problem 37 to estimate the number of degree-days accumulated from t = 0 to t = 20 for citrus flower which has a lower developmental threshold of 49°F. Use time intervals of width four days.

39. Suppose the velocity v (in meters per second) of a runner during the first few seconds of a race is given by

images

Plot these points in the t v plane. Sketch the velocity curve. Estimate the distance traveled by the runner by estimating the area under the velocity curve; use rectangles with heights given by a right endpoint approximation.

40. A pneumotachograph is a medical device used to measure the rate at which air is exhaled by a patient's lungs. Suppose Figure 5.16 shows the rate of exhalation for a particular patient. The area under the graph provides a measure of the total volume of air in the lungs during exhalation. Use a Riemann sum with n = 8 and right-endpoint subinterval representatives to estimate the volume.

images

Figure 5.16 Rate of exhalation

41. An industrial plant spills pollutant into a lake. Suppose that the pollutant spread out to form the pattern shown in Figure 5.17. All distances are in feet.

images

Figure 5.17 Pollutant spill

Use a Riemann sum with n = 6 and right-endpoint subinterval representatives to estimate the area of the spill.

42. images

images

In this section, we saw that history honored Georg Riemann by naming an important process after him. In his personal life Riemann was frail, bashful, and timid; but in his professional life, he was one of the all-time giants in mathematics. In his book, Space Through the Ages, Cornelius Lanczos wrote, “Although Riemann's collected papers fill only one single volume of 538 pages, this volume weighs tons if measured intellectually. Every one of his many discoveries was destined to change the course of mathematical science.” One of these discoveries is the Riemann zeta function

images

which had already been considered by Euler. In this function, the sum is over all natural numbers n, and the product is over all prime numbers. While Euler considered the zeta function in the context of a real variable z, Riemann considered the zeta function in the context of a complex variable z. Except for a few trivial exceptions, the roots of ζ(s) all lie between 0 and 1. Riemann conjectured that the zeta function had infinitely many nontrivial roots, all with their real parts equal to 1/2. This is the famous Riemann hypothesis, which remains one of the most important unsolved problems in mathematics. The Clay Mathematics Institute, for example, has offered a million dollar prize for solving this conjecture. (For more information about this institute, visit http://www.claymath.org.) So you can become a millionaire doing mathematics!

Write a paper on Georg Riemann; in particular, discuss this million dollar prize.

5.3 The Definite Integral

Previously we defined area under a nonnegative function as the limit of a Riemann sum. In this section we define this limit for any continuous function (positive or negative) and develop its geometrical meaning as well as its properties.

For a nonnegative continuous function f(x) from x = a to x = b, we defined the area under the curve as

images

where Δx = images and xi is a point from the interval [a + (i − 1) Δx, a + iΔx]. Theorem 5.1 from the previous section implies that

images

exists and is independent of the sample points xi whenever f is continuous. When f takes on negative values, the integral no longer corresponds to the area under the curve, but the signed area as we soon shall see. The existence of the limit is so important that Leibniz (see images, page 538) developed the following special notation for it.

Definite Integral

Let f be continuous on [a, b]. Then the definite integral of f from a to b is defined to be

images

In the definition of the definite integral, the function f that is being integrated is called the integrand; the interval [a, b] is the interval of integration; and the endpoints a and b are called, respectively, the lower and the upper limits of integration. The variable x is called the variable of integration. Notice that in taking the limit the Greek letters are supplanted by the Roman letters: Δ becomes a d and Σ becomes an elongated S.

Example 1 From sums to integrals

Write the sum

images

as a definite integral.

Solution There are several ways we can answer this problem depending on how we view the Riemann sum. For instance, we can view this Riemann sum corresponding to an integrand sin x with sample points xi = images. Since the first sample point x1 = images approaches 0 as n increases, the lower limit of integration must be 0. Since the last sample point xn = images = 2π for all n, the upper limit of integration must be 2π. Hence, the definite integral is

images

Alternatively, we can always represent the limit of the Riemann sum as an integral from x = 0 to x = 1.(In fact, we can choose the limits of integration to be any a < b and still get things to work out!) With this view, our sample points need to be xi = images. Hence, the argument of the sum is equal to sin(2πxi)2πΔx and the Riemann sum converges to

images

Note that the two expressions obtained for the integrals in Example 1 must be the same for the theory of integration to be consistent. This will be demonstrated in Section 5.5, after we consider how to change the variable of integration.

Example 2 From integrals to sums

Write the integral

images

as a limit of a Riemann sum.

Solution Our integral images f(x)dx has integrand f(x) = images and the limits of integration a = 1 and b = 4. If we break up the interval [1, 4] into n subintervals of equal width, then

images

Choosing the right endpoints of the intervals as sample points gives

images

Hence, the definite integral equals

images

Example 3 Approximating integrals with sums

Approximate the integral

images

by the sum

images

where the xi correspond to right endpoints.

Solution Since the integrand, tan x, is continuous on the interval [−1, 0.5], the integral is well defined. The summation expression in the problem statement implies that n = 6. Thus, we choose Δx = images = 0.25, in which case x0 = −1, x1 = −0.75, x2 = −0.5, x3 = −0.25, x4 = 0, x5 = 0.25, and x6 = 0.5. The Riemann sum with right end points is

images

A graphical representation of this sum is shown in green in Figure 5.18. Notice that we got a negative number, as the areas of the rectangles below the x axis were greater than the areas of the rectangles above the x axis.

images

Figure 5.18 Graph of y = tan x with approximating rectangles

Example 4 Computing an integral using a summation formula

Use a summation formula to compute

images

Solution Break the interval [0, 2] into n subintervals whose endpoints are images. The corresponding Riemann sum is

images

with Δx = images. Expanding and rearranging terms yields

images

Taking the limit of this expression as n → ∞ yields

images

Again, we have an integral that is negative.

Geometrical meaning of the definite integral

We saw previously that images f(x)dx corresponds to the area under the curve y = f(x) provided that f(x) ≥ 0 from x = a to x = b. The following example uses this fact to evaluate an integral.

Example 5 Integral of dx rule

Evaluate

images

Solution Let f(x) = 1 with limits of integration x = a and x = b.

images

If we plot f over [a, b], we can see this is the area of a rectangle of height 1 and width (ba). Thus,

images

What happens if f(x) changes sign on the interval? In this case, images f(x)dx is the signed area of the region R determined by the curve y = f(x) and the lines y = 0, x = a, and x = b. More specifically, if f changes sign on the interval [a, b], then the region R breaks up into two pieces: one piece, call it R, that lies below the x axis as illustrated by the red region in Figure 5.19 and another piece, call it R+, that lies above the x axis as illustrated by the green region in Figure 5.19.

If A+ and A denote the areas of R+ and R, respectively, then

images

images

Figure 5.19 Geometry of the definite integral

Example 6 Evaluating integrals using signed areas

Use the signed area interpretation of integrals to find

images

images

Figure 5.20 Graph of f

images

Figure 5.21 Graph of g

Solution

  1. Let f(x) = x on [−1, 2], as shown Figure 5.20. The graph forms two triangles, R+ and R, that lie above and below the x axis, respectively. The area of R+ is 2 and the area of R is images. Hence,

    images

  2. Let g(x) = images on [−3, 3], as shown in Figure 5.21. The graph forms a semicircle of radius 3. The graph is always above the axis; consequently, we need its area. Using the formula for the area of a circle,

    images

  3. Let h(x) = x5 on [−3, 3], as shown in Figure 5.22. Notice that this graph is symmetric with respect to the origin; consequently, it has the same area above and below the x axis. Therefore,

    images

    images

    Figure 5.22 Graph of h

Properties of definite integrals

Integrals satisfy several useful properties, some of which are summarized in the following box.

Properties of the Definite Integral: Part I

Let f and g be continuous functions on the interval [a, b].

Sum rule images

Difference rule images

Scalar rule images

Opposite rule images

These properties can be proved using Riemann sums and limit laws (see Problem Set 5.3).

Example 7 Using the properties of definite integrals

Evaluate images.

Solution

images

Combining the properties of integrals with the geometrical interpretation of the integral allows one to quickly compute certain integrals.

Example 8 Growing grapes

Thompson seedless grapes have a lower developmental threshold of 50°F and require approximately 3,000 degree-days to ripen after pollination. Suppose the temperature in the fields over a particular ten-day period is given by

images

where x is time in days. Write an expression involving definite integrals that represents the number of degree-days accumulated from x = 0 to x = 10, and evaluate this expression. Use the expression to determine the percent of development that takes place during this ten-day period.

Solution We are interested in finding the area between the curves y = 50 and y = 70 + 10 sin(2πx) from x = 0 to x = 10, as illustrated in Figure 5.23.

Since this area is computed by finding the area below the curve y = 70 + 10 sin(2πx) and then subtracting the area below the curve y = 50, the result is

images

images

Figure 5.23 Degree-days accumulated for ten days

Since the integral of sin(2πx) has equal area above and below the x axis on the interval [0, 10], its value is zero. Hence, the number of degree-days accumulated is 200. This area could be found by noticing that the “hills” of the temperature functions fit in the valleys, yielding a 20 by 10 rectangle.

Since 200 of 3000 degree-day units accumulated over the ten-day period, this period represents p = 200/3000 ≈ 0.0667 of the total number of degree-days needed to ripen. Hence, 6.67% of needed ripening took place during this ten-day period.

We conclude this section with some additional properties of the definite integral.

Properties of the Definite Integral: Part II

Assuming all integrals exist, we have the following properties.

Nonnegativity If f(x) ≥ 0 from x = a to x = b, then

images

Dominance If f(x) ≥ g(x) from x = a to x = b, then

images

Bounding If mf(x) ≤ M from x = a to x = b, then

images

Splitting

images

Definite integral at a point

images

Nonnegativity can be proved using the definition of a definite integral, and nonnegativity, in turn, can be used to prove dominance and bounding. For example, to prove dominance, suppose that f(x) ≥ g(x) from x = a to x = b. Then, f(x) − g(x) ≥ 0 from x = a to x = b. Applying the property of differences and nonnegativity yields

images

Hence,

images

images

Figure 5.24 Geometrical depiction of the splitting property where the signed area images f(x)dx from x = a to x = b equals the signed area images f(x)dx from x = a to x = c plus the signed area images f(x)dx from x = c to x = b

If we set M and m to be the maximum value and minimum value, respectively, of f on the interval [a, b], then the bounding property provides crude estimates for the value of a definite integral. When working through detailed computations by hand, these crude estimates allow us to see whether our work has resulted in a reasonable answer. Finally, a proof of the splitting property is somewhat subtle, but geometrically intuitive, as illustrated in Figure 5.24.

Example 9 Using bounds

Show that

images

Solution Since the sine function is bounded between −1 and 1, it follows that 8 ≤ 10 + 2 sin x2 ≤ 12 for all x as illustrated in the left margin. The bounding property implies that

images

images

which yields the desired result.

Example 10 Using the splitting property

Suppose that images.

Solution By the splitting property,

images

Thus,

images

PROBLEM SET 5.3

Level 1 DRILL PROBLEMS

Express the limits in Problems 1 to 6 as definite integrals of the form images f(x)dx.

images

Express the definite integrals in Problems 7 to 12 as limits of Riemann sums.

images

First sketch the region under the graph of y = f(x) on the interval [a, b]. Then use the interpretation of the definite integral images f(x)dx as a signed area to evaluate the integrals in Problems 13 to 16.

images

Evaluate each of the integrals in Problems 17 to 22 by using the following information together with the sum rule and the splitting property:

images

Use integral properties to establish the statements in Problems 23 to 26.

images

27. Use the fact that images x2dx = images and the geometrical interpretation of the integral to find

images

28. Use the graph of y = cos x to evaluate

images

on the indicated interval.

images

Using right endpoints with n = 5, approximate the definite integrals in Problems 33 to 36. Indicate whether each approximation is greater than or less than the actual definite integral.

images

images

Use Riemann sums with right endpoints, along with a summation formula (see Section 5.2) to evaluate the integrals in Problems 37 and 38.

images

Show that each statement about area in Problems 39 to 42 is true in general, or if not, provide a counterexample. It will probably help to sketch the indicated region for each problem.

39. If C > 0 is a constant, the region under the line y = C on the interval [a, b] has area A = C(ba).

40. If C > 0 is a constant and b > a ≥ 0, the region under the line y = Cx on the interval [a, b] has area A = imagesC(ba).

41. Let f be a function that satisfies f(x) ≥ 0 for x in the interval [a, b]. Then the area under the curve y =[f(x)]2 on the interval [a, b] must always be greater than the area under y = f(x) on the same interval.

42. A function f is said to be even if f(−x) = f(x). If f is even and f(x) ≥ 0 throughout the interval [−a, a], then the area under the curve y = f(x) on this interval is twice the area under y = f(x) on [0, a].

Level 2 APPLIED AND THEORY PROBLEMS

43. We saw in Example 8 that Thompson seedless grapes have a lower developmental threshold of 50°F and require approximately 3,000 degree-days to ripen. Suppose the temperature in the field is given by

images

over a 15 day period where x is time in days. Write an expression involving definite integrals that represents the number of degree-days accumulated from x = 0 to x = 15, and evaluate this expression. What percentage of ripening took place during this 15 day period?

44. Assume for a particular variety of grape that the lower developmental threshold is 30°F and that it takes 2,500 degree-days to ripen. If the temperature in degrees Fahrenheit over a particular 15 day period is given by

images

where x is the time in days, then write an expression involving definite integrals that represents the number of degree-days accumulated from x = 0 to x = 15. Evaluate this expression and calculate the percentage ripening the occurred over this 15 day period.

45. Assume that over a 50 day period the rate at which individuals in a population die from disease is given by the function f(x), where x is days and the units of f(x) are individuals per day. If f(x) has the form

images

then write an expression involving definite integrals that represents the number of deaths over this 50 day period and evaluate this expressing using your knowledge of how to calculate areas of geometrical objects.

46. Assume that over a 60 day period the rate at which individuals in a population die from disease is given by the function f(x), where x is days and the units of f(x) are individuals per day. If f(x) has the form

images

then write an expression involving definite integrals that represents the number of deaths over this 60 day period and evaluate this expressing using your knowledge of how to calculate areas.

47. A function f is said to be odd if f(−x) = −f(x). Show that if f is odd on the interval [−a, a], then images f(x)dx = 0.

48. Prove the sum rule for integrals using the definition of a definite integral and properties of limits and summations.

49. Generalize the splitting property by showing that for acdb

images

whenever all these integrals exist.

50. Prove the bounding rule for definite integrals: If f is continuous on the closed interval [a, b] and mf(x) ≤ M for constants, m, M, and all x in the closed interval, then

images

51. images Gilles de Roberval (1602–1675) started his study of mathematics at the age of 14 years. He had a distinguished career and was a founding member of the Académie Royale des Sciences. In 1669 he invented the Roberval balance (Figure 5.25), which is still widely used today.

images

Figure 5.25 Roberval's balance variations

Roberval had a chair position as professor of mathematics at the Collége Royale. Every three years there was a contest by competitive examination (written by the incumbent!) to determine who would occupy this position. It is said for this reason, Roberval kept many of his techniques of integration secret until his death. We do know, however, that he developed powerful methods in the early study of integration. These methods are described in his treatise Traité des indivisibles. For instance, in this treatise, he computed the definite integral of sin(x) using obscure trigonometric identities. It is these identities that the computer algebra system uses to simplify the sum and take the limit. For this quest, you may stand on the shoulders of Roberval and use technology to write a Riemann sum for

images

using right endpoints. Then use technology to simplify this sum to evaluate this definite integral.

5.4 The Fundamental Theorem of Calculus

In this section, we discuss the evaluation theorem and the fundamental theorem of calculus. These theorems link antiderivatives, which we can compute relatively easily, with definite integrals and Riemann sums. We show that antiderivatives and Riemann sums, when they both exist, are the same thing.

Evaluation theorem and net change

Theorem 5.2 Evaluation theorem

Let f be a continuous function on [a, b] and F be any antiderivative of f. Then,

images

The proof of this theorem is a corollary of the fundamental theorem of calculus (FTC), which is discussed later in this section. Why is the evaluation theorem useful? Well, as we saw in the beginning of this chapter, finding antiderivatives is much easier than taking limits of Riemann sums. This theorem allows us to evaluate definite integrals by finding and evaluating an antiderivative. This fact is so important that this theorem is sometimes itself called the fundamental theorem of calculus, even though it is only a corollary to the more powerful theorem of the same name.

Example 1 Using the evaluation theorem

Evaluate the following definite integrals.

images

Solution

  1. Since an antiderivative of f(x) = x7 is F(x) = imagesx8, the evaluation theorem tells us

    images

    Notice that if we took another antiderivative, say F(x) = imagesx8 + 14, we still get

    images

    as the constant term 14 cancels out.

  2. Since an antiderivative of f(x) = x−1 is F(x) = ln|x|, the evaluation theorem tells us

    images

    Notice that we used the fact that images is continuous on the interval [1, 2]. The fundamental theorem would not apply to images is not continuous on the interval [−1, 1].

  3. Since an antiderivative of f(x) = x−1 is F(x) = ln|x|, the evaluation theorem tells us

    images

    Compare this with the previous answer, which is + ln 2.

  4. Since the antiderivative of f(x) = sin x is F(x) = −cos x, the evaluation theorem tells us

    images

To appreciate the power of the evaluation theorem, compare the work of Example 1 with work required to find the limit of the Riemann sum for each of these functions.

To simplify our work, we introduce the following notation:

images

Or, when there is no ambiguity,

images

Example 2 Using evaluation notation

Evaluate images.

Solution Since the antiderivative of sec2x is tan x, the evaluation theorem tells us

images

Notice that we used the fact that sec2x is continuous on the interval from x = π to 5π/4. The fundamental theorem does not apply on the interval from x = π to 2π, as sec2x is not defined at x = 3π/2.

As illustrated in Example 3 of Section 5.2, an important interpretation of the evaluation theorem is that it relates the accumulated change of a function over an interval to the area under its derivative. Thus, we have the following procedure for calculating accumulated change.

Accumulated Change

If F is an antiderivative of f, then by the evaluation theorem we have

images

For instance, if F(x) represents the total number of births in the world by year x, then F(b) − F(a) is the number of births that occurred between years a and b. Now suppose that f(x) = F′(x), which by definition implies that F′(x) is the instantaneous birthrate x. The evaluation theorem asserts that the area under the instantaneous birthrate equals the accumulated change in births. Does this make sense? At the level of units it certainly does. The instantaneous birthrate has units births per year. Hence, the area over an interval of time has units births per year multiplied by years. This equals births, which has the same units as the accumulated change. Moreover, if we broke up the interval [a, b] into small subintervals, then the number of births in a given subinterval would be approximately the instantaneous birthrate at some point in the subinterval in question multiplied by the length of the subinterval, that is, the area of a rectangle lying above this subinterval. Adding up all these little rectangular areas would give us simultaneously an approximation for the number of births over [a, b] and the area under F′. More generally, if one integrates the rate of change over an interval [a, b], then one gets the accumulated change over [a, b].

Example 3 Horn increase for the bighorn ram

Bighorn sheep (Ovis canadensis) (see Figure 5.26) inhabit remote mountain and desert regions. They are restricted to semi-open, precipitous terrain with rocky slopes, ridges, and cliffs or rugged canyons. Forage, water, and escape terrain are the most important components of their habitat. Jon Jorgenson and colleagues* found that the rate of increase of a bighorn ram's horn is approximated by the function

images

for x between three and nine years.

Find the accumulated change in the length of a bighorn ram's horn from age x = 3 to x = 9.

images

Figure 5.26 Bighorn ram

Solution Let F(x) denote the length of a ram's horn at age x years. Then,

images

The evaluation theorem implies

images

Fundamental theorem of integral calculus

The answer to Example 3 gives us the net increase in the length of the ram's horn over the whole six-year period. Suppose, though, that we want to find the net increase at any age x during the six-year period from age 3 to age 9. The net increase is given by the function

images

for 3 ≤ x ≤ 9. For example, we found F(9) ≈ 33.8148 cm. In writing the integral defining F(x), we took advantage of the fact that the variable of integration is a dummy variable (it disappears once the integration has been performed). Consequently, to avoid confusion, we choose the variable u of integration to be different from our time variable x. By the evaluation theorem,

images

images

Figure 5.27 Estimated growth of a ram's horn in centimeters

Plotting F(x) from x = 3 to x = 9 as shown in Figure 5.27 illustrates how the net increase of the length of the horn changes over this time interval. Notice, as we might expect, the length of the horn is increasing at a decreasing rate.

We can now generalize this idea for any continuous function f defined on the interval [a, b] by considering the function

images

If we interpret f(x) as a rate, then F(x) describes how the accumulated change varies as a function of x. Alternatively, F(x) describes how the signed area under f confined to the interval [a, x] varies as a function of x.

Theorem 5.3 Fundamental theorem of calculus (FTC)

Consider a continuous function f on the interval [a, b]. Then, F defined by

images

is an antiderivative of f(x) on (a, b). In other words,

images

Why should this be true? The idea of the proof is as follows. The splitting property of integrals implies that

images

On the other hand, continuity of f implies that f(u) ≈ f(x) for u between x and x + Δx. Furthermore, this approximation gets better and better as Δx approaches zero. Thus,

images

Dividing both sides by Δx and letting Δx go to zero suggests that

images

To really show that this final statement is true requires a bit more care using images-δ arguments of the type introduced in Section 2.2.

An important consequence of the fundamental theorem of calculus is that it proves that every continuous function f has an antiderivative F, even though F cannot always be expressed using a combination of elementary functions (e.g., polynomial, exponential, and trigonometric functions). A corollary of the fundamental theorem of calculus is the evaluation theorem, as discussed in Problem 33 in Problem Set 5.4.

Example 4 Derivatives via the fundamental theorem

Compute the following derivatives.

images

Solution

  1. Since images is a number, we are taking the derivative of a constant and images.
  2. By the fundamental theorem of calculus, images.

Example 5 From integrals to integrands

Suppose

images

Find f and a.

Solution Let F(x) = images f(u)du. By the fundamental theorem of calculus,

images

To find a, notice that

images

Therefore, a = images.

Using the fundamental theorem, we can easily compute the accumulation of degree-days.

Example 6 Seedless grapes

Thompson seedless grapes (see Figure 5.28) have a lower developmental threshold of 50°F and require approximately 3,000 degree-days to ripen.

Suppose the temperature (degrees Fahrenheit) in the fields is given by

images

where t is time in days. Write an expression involving definite integrals that represents the number of degree-days accumulated from day 0 to day x, evaluate this expression, and find the time x at which 3,000 degree-days have accumulated.

images

Figure 5.28 Seedless grapes

Solution Since T(t) ≥ 50 for all t (i.e., the lower developmental threshold does not require consideration), the number of degree-days accumulated by day x is given by F(x) = images(T(t) − 50)dt. Integrating yields

images

Since F′(x) = 20 + 10 sin(2πx) > 0, F is an increasing function. Therefore, if we find a positive solution, then it is the only solution. Notice that if x is an integer, then

images

Solving

images

The grapes will be ready for picking in 150 days. Although the first part of this solution required the fundamental theorem of calculus, the latter part of the problem could be answered using the geometrical interpretation of the definite integral.

Indefinite integrals

Since the fundamental theorem of calculus ensures the existence of antiderivatives via integrals, it is appropriate to introduce a notation for the general antiderivative.

Indefinite Integral

If f(x) is a continuous function, then

images

is called the indefinite integral of f and is equal to the general antiderivative of f.

The fact that the indefinite integral has no upper limit of integration or lower limit of integration distinguishes it from a definite integral. It is important to remember that the indefinite integral represents a family of functions (a particular function plus an arbitrary constant), whereas the definite integral of a specified function represents a value.

Example 7 Finding indefinite integrals

Compute the following indefinite integrals.

images

Solution

  1. ex dx = ex + C where C is an arbitrary constant.
  2. ∫ (x3 + 2) dx = images + 2x + C where C is an arbitrary constant.
  3. ∫ sec2x dx = tan(x) + C where C is an arbitrary constant.

PROBLEM SET 5.4

Level 1 DRILL PROBLEMS

In Problems 1 to 8, evaluate the definite integral.

images

images

images

Find the indefinite integrals in Problems 9 to 16.

images

Compute F′(x) for the functions given in Problems 17 to 22.

images

images

images

Figure 5.29 Graph of f

  1. For what values of x does F(x) have a local maximum or minimum?
  2. For what values of x is F concave up? concave down?
  3. At what values of x does F(x) achieve a global maximum? global minimum?
  4. Sketch the graph of F(x).

27. Let G(x) = imagesg(u)du where the graph of g is shown in Figure 5.30.

images

Figure 5.30 Graph of g

  1. For what values of x does G(x) have a local maximum or minimum?
  2. For what values of x is G concave up? concave down?
  3. At what values of x does G(x) achieve a global maximum? global minimum?
  4. Sketch the graph of G(x).

Level 2 APPLIED AND THEORY PROBLEMS

28. Use the model for bighorn rams formulated by Jorgenson and colleagues (see Example 3) to find the net increase in length of a ram's horn from x = 3 to x = 7.

29. Use the model for bighorn rams formulated by Jorgenson and colleagues (see Example 3) to find the net increase in length of a ram's horn from x = 5 to x = 9.

30. Citrus flowers in Tulare County, California, have a lower developmental threshold of 49°F and require approximately 767 degree-days to reach petal-fall (i.e., 50% of citrus flowers have lost their petals). Suppose the temperature in the fields is given by

images

where x is the time in days.

  1. Write an expression involving definite integrals that represents the time, x, at which 767 degree-days have accumulated.
  2. Use technology to estimate the time x.

31. Sweet corn in western Oregon has a lower developmental threshold of 50°F and requires approximately 1,597 degree-days to reach maturity. Suppose the temperature in the fields is given by

images

where x is the time in days.

  1. Write an expression involving definite integrals that represents the time, x, at which 1,597 degree-days have accumulated.
  2. Use technology to estimate the time x.

32. The rate of change of ant diversity along an elevational gradient is given by

images

where x is elevation above sea level measured in kilometers. If F(1) = 6.9, find an expression for F(x), the number of ant species at an elevation of x km. Compare your answer to Example 4 of Section 2.7.

33. Prove the evaluation theorem (Theorem 5.2) using the fundamental theorem of calculus (Theorem 5.3).

5.5 Substitution

In the next two sections, we discuss three techniques of integration: substitution, integration by parts, and partial fractions. The first two of these techniques are counterparts to rules of differentiation. Unlike differentiation, techniques of integration are incomplete; not every function has an elementary representation of its indefinite integral. Consequently, part of the skill you need to acquire is knowing which integration techniques apply to which functions. Since it is possible to compute integrals using technology, you may be asking: Why bother with these techniques? There are several reasons. First, by learning integration techniques and determining when and in which order to use them, you gain insight into how these techniques work. Second, sometimes technology needs a helping hand; implementing an integration technique (especially substitution) by hand may allow technology to complete a calculation it could not otherwise do. Third, computing integrals builds important mathematical skills: it's like lifting weights to improve your strength for sports.

Substitution for indefinite integrals

Our integration efforts begin with an antidifferentiation form of the chain rule. Consider the integral

images

Basic rules of antidifferentiation provide no direct way of computing this integral. However, look carefully at this integral, and note that the derivative of the denominator equals the numerator. This observation suggests introducing a new variable, u = x2 + 5, that, on differentiation, yields images = 2x. In this next step we replace 2xdx = imagesdx in the integrand with du. A mathematical justification for this step is dealt with in more advanced texts. Finally, integrating with respect to the variable u, we obtain

images

Although we have not justified our steps rigorously, we can verify our answer by differentiating. Since

images

it follows that ln(x2 + 5) + C is the general antiderivative of images.

To see why this approach worked, consider the integral

images

for functions f and g. In our example imagesdx, we have f(x) = 1/x and g(x)= x2 + 5. If F is an antiderivative of f, then

images

Equivalently, if we make the change of variables u = g(x), then

images

We summarize this procedure in the following box.

Integration by Substitution

Over any interval of x for which u = g(x) is continuously differentiable and f is continuous on the range of this function, the relationship

images

holds.

Example 1 Integration by substitution

Find

images

Solution For the procedure of substitution, we need to identify the appropriate change of variables.

images

This procedure may seem difficult now, because you may not be sure of what to let the variable u represent. Just remember, at least initially, you are looking for one part of the integrand that is the derivative of another part of the integrand. If you practice enough, things will get easier!

Example 2 Substitution with a radical function

Find images.

Solution Let u = 3x + 7, so du = 3dx or dx = images. Substituting and integrating yields

images

As the previous two examples illustrate, after making a substitution and simplifying, there should be no x values in the integrand. Sometimes eliminating all the x's requires additional work.

Example 3 Substitution with leftover x values

Find images.

Solution Let u = 4x + 5. Then, du = 4dx and dx = images.

images

Example 4 Substitution with a trigonometric function

Find

images

Solution Recall that tan x = images cos x = −sin x. Hence, using u = cos x gives

images

Some of you will, no doubt, have access to a calculator or a computer program that can assist in the process of integration. Although technology is very useful, there are times when evaluating an integral by hand will result in a simpler form of the answer. In other cases, technology will give an incomplete form that needs to be adjusted.

Example 5 Using technology to integrate

Use technology to find

images

and compare the result with the results shown in Example 4.

Solution Some calculators or software programs will return the answer −ln(|cos(x)|) and others may return −ln cos x. In either case, compared with the previous example, the constant, C, is missing.

Lest you think that if you purchase a calculator you will not need to study and master techniques of integration, consider the following example.

Example 6 Use substitution with the help of technology

Find a closed-form algebraic expression for

images

Solution If we apply technology to solve this problem using the form of the integrand as it stands, we will likely not obtain a satisfactory solution. However, if we first substitute u = x ln x and du = (1 + ln x) dx then we have

images

Using technology on this transformed integrand will yield one of the following two particular antiderivatives:

Expression 1

images

where sinh θ = (eθe−θ)/2 is called the hyperbolic sine function and sinh−1 is its inverse.

Expression 2

images

Despite looking quite different, these two expressions must be algebraically equivalent up to a constant. If you cannot show this algebraically, you can graph the expressions to demonstrate their equivalence.

Substitution for definite integrals

We have two methods for dealing with definite integrals. The first is to return to the original variable (as we did with indefinite integrals). The second is to keep track of the change of variables in the limits of integration. We illustrate both of these methods by revisiting Example 4.

Method I: Return to the original variable.

images

Method II: Keep track of the change of variables in the limits of integration.

images

Consider the general case of this second method. Let u = g(x) and F(x) be an antiderivative of f(x). Then,

images

We summarize this observation in the following box.

Substitution with Definite Integrals

If g′(x) is a continuous function on [a, b] and f is continuous on the range of u = g(x), then

images

Example 7 United States Population Growth

The logistic formula

images

provides a reasonably good fit to the population of the United States (in millions) during the period 1790–1990, as illustrated in Figure 5.31. The variable t is the time (in decades) after 1790. Thus, t = 0 for 1790, t = 20 for 1990. Suppose that each person eats food at a rate of one annual-person-ration (APR) per year. Find the total number of APRs of food eaten in the United States between 1790 and 1990.

Solution Since each person consumes one APR per year, the rate at which food is being eaten in decade t is 10 P(t) APRs per decade. To find the number of APRs consumed over 20 decades, we integrate 10 P(t) from t = 0 to t = 20.

images

Figure 5.31 U.S. population growth (in millions)

images

Thus, between the years 1790 and 1990, 17,249 million APRs were consumed!

Example 8 Breathing

Breathing is a cyclic process, as illustrated in Figure 5.32. One cycle of breathing from the beginning of inhalation to the end of exhalation takes about five seconds. Since the maximum rate of airflow into the lungs during a typical breath is approximately images liter each second, we can model the rate of airflow into the lungs by the function

images

where t is the time in seconds. Find the total amount of air inhaled in one cycle. This volume of is air is known as the tidal volume.

images

Figure 5.32 Breathing cycle

Solution The time for inhalation is images seconds, so total amount of air inhaled in one cycle is images, in which case dt = images. When t = 0, u = 0, and when t = 5/2, u = π. Hence,

images

PROBLEM SET 5.5

Level 1 DRILL PROBLEMS

Problems 1 to 8 present pairs of integration problems; one of the pair will use substitution and one will not. As you are working these problems, think about when substitution may be appropriate.

images

Use substitution to find the indefinite integrals in Problems 9 to 16.

images

Use substitution to evaluate the definite integrals in Problems 17 to 24.

images

images

Level 2 APPLIED AND THEORY PROBLEMS

28. In Example 7, U.S. population growth was modeled by

images

where t is decades after 1790. If each person eats food at a rate of one ration per year, find the total number of rations of food eaten in the United States between 1800 and 1900.

29. Assume that a dust mite population starts with 10 dust mites and grows at a rate of 10e0.3t dust mites per hour. How many dust mites will there be one day from now?

30. Suppose an environmental study indicates that the ozone level, L, in the air above a major metropolitan center is changing at a rate modeled by the function

images

parts per million per hour (ppm/h) t hours after 7:00 A.M.

  1. Express the ozone level L(t) as a function of t if L is 4 ppm at 7:00 A.M.
  2. Use the graphing utility of your calculator to find the time between 7:00 A.M. and 7:00 P.M. when the highest level of ozone occurs. What is the highest level?

31. The Gompertz law of tumor growth is given by the equation

images

where N is the size of the tumor, t is time (measured in days), b is the asymptotic size of the tumor, and a is a measurement of the tumor growth rate. Assume a = 1 and b = 10. Integrate both sides of the Gompertz equation and solve for N in terms of t. To get rid of the integration constant, assume that N equals 5 at time t = 0.

32. In Example 6 of Section 1.5, we modeled the uptake of glucose by bacterial populations off of the coast of Peru by the function

images

where x is micrograms of glucose per liter. Suppose the concentration of glucose x is decaying exponentially in time: x(t) = 100e−0.01t micrograms per liter where t is measured in hours.

  1. Write a function U(t) that describes how the uptake rate is changing in time.
  2. Determine the net uptake of a cell from t = 0 to t = 6 hours.

33. In Example 5 of Section 2.4, we found that the rate at which wolves kill moose can be modeled by

images

where x is measured in number of moose per km2. Suppose that the density of moose is increasing exponentially according to the function x(t) = 0.1e0.2t moose per km2 where t is measured in hundreds of days. Determine the number of moose killed by a wolf from t = 0 to t = 3.

34. In Problem 42 in Problem Set 2.4, we examined how wolf densities in North America depend on moose densities. We found that the following function provides a good fit to the data:

images

where x is number of moose per km2. Assume the moose density is increasing exponentially according the function x(t) = 0.1e0.2t moose per km2 where t is measured in hundreds of days. Determine the change in the wolf density from t = 0 to t = 3.

5.6 Integration by Parts and Partial Fractions

In this section, we consider two important techniques of integration that will be useful in later chapters.

Integration by parts

Integration by parts is a procedure based on inverting the product rule for differentiation. To derive a formula for this procedure, we begin with the product rule for differentiating functions f(x) and g(x), assuming these derivatives exist.

images

If we let u = f(x) and v = g(x), then du = f′(x)dx, dv = g′(x)dx, and we obtain the following simplified formula.

Integration by Parts

images

To evaluate integrals using integration by parts, we want to choose u and dv so that the new integral is easier to integrate than the original integral.

Example 1 Integration by parts

Find

images

Solution For this example, there are two ways we can choose u and dv. Suppose we choose u = x and dv = ex dx. We differentiate u and integrate dv. Thus, du = dx, and v = ex. Now, substitute these values into the integration by parts formula.

images

We noted that there were two possible choices for u and dv in Example 1. The other choice is to let u = ex and dv = x dx. If we make this choice, and substitute into the formula for integration by parts, then we obtain

images

(Try this yourself to practice the technique.) In this case, instead of simplifying the problem and solving it, we just made it more complicated! Thus, it is important to try different substitutions to see which works best.

Example 2 When the differentiable part is the entire integrand

Find

images

assuming x > 0.

Solution Let u = ln x and dv = dx. Then, du = images, v = x, and

images

Example 3 Repeated use of integration by parts

Find

images

Solution Let u = x2 and d v = e2x dx. Then, du = 2xdx, v = imagese2x, and

images

To compute the rightmost integral, we need another application of integration by parts. Let u = x and d v = e2xdx. Then, du = dx, v = imagese2x, and

images

In the next example, it is necessary to apply integration by parts more than once. As you will see, though, when we do so a second time, we return to the original integral.

Example 4 There and back again

Find

images

Solution For this problem it will be useful to call the initial antiderivative I and ignore the constant of integration until the end of the calculation. That is, let

images

Hence, the general form of the antiderivative is

images

Sometimes you need to combine techniques to conquer an integral.

Example 5 Combining substitution and integration by parts

Find

images

The function x3ex2 arises in the context of considering the properties of the Gaussian or normal distribution, which is the most important distribution in statistics.

Solution

images

Integration by parts extends to definite integrals in a natural way.

Integration by Parts with Definite Integrals

If f(x) and g(x) are differentiable functions of x on the interval [a, b], then

images

Example 6 Integration by parts with a definite integral

Evaluate

images

Solution Let u = s and d v = es/2ds, so that du = ds and v = −2es/2.

images

Example 7 Survival to age t

Suppose a biologist found that for a particular population of monkeys, the proportion of individuals born each year who die before they are t years old is

images

  1. What proportion of individuals dies before the age of 3?
  2. What proportion of individuals dies between ages 3 and 4?
  3. What proportion of individuals lives to be at least age 6?
  4. At what rate is the proportion changing at age 1? age 4?

Solution From Example 6, we found that p(t) = −imageset/2(t + 2 − 2et/2). In particular, the proportion of individuals dying at age 0 equals p(0) = 0.

  1. p(3) = imagese−3/2(3 + 2 − 2e3/2) ≈ 0.442; so about 44% of the population will die before the age of 3.
  2. The proportion that will die before the age of 4 is

    images

    Thus, the proportion dying between ages 3 and 4 is

    images

    That is, about 15% of the population will die between the ages of 3 and 4.

  3. The proportion living to be at least age 6 is one minus the number that die before the age of 6. We find

    images

    Thus, the desired number is

    images

    Therefore, we would expect 20% of the individuals to live to at least the age of 6.

  4. Using properties of integrals and the fundamental theorem of calculus, we have that

    images

    Hence, the proportion is changing at an instantaneous rate of imagese1/2 ≈ 0.1516 at age 1 and images4e−2 ≈ 0.1353 at age 4.

Partial fractions

Partial fractions is an integration technique enable us to integrate any rational function; that is, functions of the form

images

Integration problems involving rational functions arise commonly in enzyme kinetics, evolutionary games, and population dynamics. For example, in describing the growth of a population of size N(t) with a growth that is positively impacted by its own size, we may encounter an integral of the form

images

The appropriate integration procedure is to write (expand) the rational function images into a sum of two simpler functions that we can directly integrate. More specifically, we try to find constants A and B such that

images

Equivalently, multiplying both sides by N(N − 1) gives

images

This equation holds for all N only if A = −1 and A + B = 0; in other words, B = −A = −(−1) = 1. Thus, we can write

images

and

images

Although it is possible to deal with all rational functions, we confine our discussion to rational functions f(x) = P(x)/Q(x), such that Q(x) can be expressed as a product of n distinct linear factors:

images

If the degree of P(x) is less than the degree of Q(x) (i.e., n > m), then we can always find constants A1, A2,..., An such that

images

For integration methods for more general rational functions, go online or read another calculus text.

Alternatively, if the degree of P(x) is greater than or equal to the degree of Q(x), then we can perform long division and factor the remainder term. When Q(x) can be decomposed into linear factors, there is a simple method to determine the coefficients. This method is the Heaviside “cover-up” method—named after Oliver Heaviside (see Problem 35, images, in Problem Set 5.6).

Example 8 Integrating a rational function

Find

images

Solution Since we can express x3x = x(x2 − 1) = x(x − 1)(x + 1) as a product of distinct linear factors, there exist constants A1, A2, and A3 such that

images

Multiplying both sides of the equation by x(x − 1)(x + 1) yields

images

Since this equation has to hold for all x, we solve for A1 by simply setting x = 0 to get 2 = −A1. Hence, A1 = −2. To solve for A2, we set x = 1 to get 3 = 2A2. Hence, A2 = 3/2. Finally, to solve for A3 we set x = −1, to get 1 = 2A3. Hence, A3 = 1/2. Notice that our choices of x ensured that all but one of the terms on the right-hand side of the equation vanished.

Thus, we can write

images

Example 9 Second-order chemical kinetics

Consider two compounds A and B that bind to form a third compound C. Assume a and b are the initial concentrations of A and B. If the rate at which C is produced is proportional to the product of concentrations of A and B, then it has been shown that the following integral equation holds when y is the concentration of C, k is a constant of proportionality, and t is time:

images

Integrate both sides of this equation and solve for y as a function of t, assuming that a = 2, b = 1, k = 3, and that y = 0 when t = 0. Sketch this function.

Solution Solve

images

To deal with the right-hand side, we need to solve for A1 and A2 such that

images

for all y. Multiplying both sides by (y − 2)(y − 1) yields

images

Setting y = 2 gives A1 = 1. Setting y = 1 gives A2 = −1. Therefore, we can rewrite the equation as

images

If t = 0, then y = 0; so we have

images

Hence, we need the positive solution +eC to equal 2. Thus,

images

images

Figure 5.33 Graph of a second-order chemical process

The graph is shown in Figure 5.33. To obtain this graph by hand, it is not hard to check that y′(t) > 0 and images y(t) = 1.

PROBLEM SET 5.6

Level 1 DRILL PROBLEMS

Find each integral in Problems 1 to 10.

images

Find the exact value of the definite integrals in Problems 11 to 16 using integration by parts. Check your answer by using a calculator to find an approximate answer correct to four decimal places.

images

Find the indicated integrals in Problems 17 to 22.

images

In Problems 23 to 28, first use an appropriate substitution and then use integration by parts or partial fractions to evaluate the integral. Remember to give your answers in terms of x.

images

30. a. Evaluate ∫ cos2x dx. Hint: Use the trigonometric identity: cos2x = images(1 + cos(2x)).

b. Use part a to evaluate ∫ x cos2x dx using integration by parts.

Level 2 APPLIED AND THEORY PROBLEMS

images

31. The 1988 film Stand and Deliver, starring Edward James Olmos pictured above, provides an alternative perspective—tabular integration—on integration by parts. This technique involves writing a table with two columns, one labeled D for differentiation and another labeled I for integration. The first row of the D column contains u, the part to be differentiated in the original integral. The second row in the D column contains images. The third row in the D column contains images. Proceed in this manner until the product of the functions in the last row either equals 0 or is a constant multiple of what you started with. The first row of the I column contains dv, the part to be integrated. For the second, third, and so on rows in the I column, place the successive integrals. For example, rework Example 1, namely ∫ xexdx, with u = x and d v = ex:

images

Now, draw diagonal lines from the first element of the D column to the second element of the I column, from the second element of D to the third element of I and so forth. Multiply the elements at the ends of each of the diagonal lines, take an alternating sum of these products, and add the integral of the product of terms in the last row. For Example 1, and the table shown here we have

images

which is the same result shown in Example 1. Use this method to find the integral in Example 2.

32. Use the tabular method from Stand and Deliver (Problem 31) on the integral in Example 3.

33. Use the tabular method from Stand and Deliver (Problem 31) on the integral in Example 4.

34. Contrast the method of integration by parts as illustrated by the examples in the text and the tabular method from Stand and Deliver (Problem 31).

35. images* Oliver Heaviside (1850–1925) was born in the same London slums as Charles Dickens. When young, Heaviside had scarlet fever, which left him partly deaf. He finished his schooling in 1865 at age 16. He was a top student but failed geometry. Heaviside went to work as a telegrapher, which drew him into the study of electricity. He read Maxwell's Treatise on Electricity and Magnetism and managed to reduce Maxwell's whole field theory into two equations.

Next Heaviside picked up the new ideas of vector analysis, which inspired him to develop a theory he called operational calculus, a powerful tool. However, its foundation was not rigorous, so his methods were not accepted by many mathematicians of his day. Only people like Kelvin, Rayleigh, and Hertz saw the brilliance that was driving Heaviside faster than method could follow. Heaviside knew what he was doing. He growled at his detractors, “Shall I refuse my dinner because I do not fully understand... digestion?”

Like vector analysis, Heaviside's calculus stood the test of time. So did the rest of his work. For example, he gave us the theory for long distance telephones. Ultimately, Heaviside grew sick of fighting his detractors and retired to Torquay in southwest England. A sketch of Oliver Heaviside is shown in Figure 5.34.

images

Figure 5.34 Oliver Heaviside (1850–1925)

Now let us consider Heaviside's so-called cover-up method for determining coefficients with partial fractions. Consider the antiderivative from Example 8. Find A1, A2, and A3 such that

images

The coefficients are found, one at a time, by “covering” that factor and evaluating the remaining expression by the value that causes the “covered” factor to be zero. That is, first cover x:

images

The covered factor is 0 when x = 0, so evaluate the noncovered portion at x = 0:

images

Thus, A1 = −2. Next, cover the factor under the A2 term:

images

Evaluate for x = −1:

images

Finally, cover the factor under the A3 term:

images

Evaluate for x = 1:

images

Thus,

images

Explain why this cover-up method of Heaviside works.

36. Assume that after t hours on the job, a factory worker can produce 100te−0.5t units per hour. How many units does the worker produce during the first three hours?

37. After t weeks, suppose that contributions in response to a fund-raising campaign were coming in at the rate of 2,000te−0.2t dollars per week. How much money was raised during the first five weeks?

38. An actuary measures the probability that a person in a certain population will die at age x by the formula

images

where λ is a parameter such that 0 < λ < e.

  1. For a given λ, find the maximum value of P(x).
  2. Sketch the graph of P(x).
  3. Find the area under the probability curve y = P(x) for 0 ≤ x ≤ 10, and interpret your result.

39. A population P, grows at the rate

images

thousand individuals per year at time t (in years). By how much does the population change during the eighth year?

40. Suppose that a drug is assimilated into a patient's bloodstream at a rate modeled by

images

where t is the number of minutes since the drug was taken. Find the total amount of drug assimilated into the patient's bloodstream during the second minute.

41. Recovering from an environmental perturbation, a (hypothetical) population exhibits dampened oscillations of the form

images

where t is measured in days. As a part of a sampling effort, a scientist captures and releases individuals from this population at a rate of 0.1N(t) individuals per acre per day. If the scientist is sampling one acre, determine the number of individuals she captures and releases in seven days.

42. Recovering from an environmental perturbation, a (hypothetical) population exhibits dampened oscillations of the form

images

where t is measured in days. As a part of a sampling effort, a scientist captures and releases individuals from this population at a rate of 0.01N(t) individuals per acre per day. If the scientist is sampling one acre, determine the number of individuals she captures and releases in ten days.

5.7 Numerical Integration

We have seen that integration is, in general, a more difficult task than differentiation. In differentiation, knowing the derivatives of several elementary functions (e.g., sin x, ex, and xn) and a set of basic rules (e.g., product rule, chain rule, and quotient rule) allows us to differentiate rather complex-looking functions. In contrast, integration is more complicated. The number of rules and special cases, and uncertainty about which rule to apply, make integration more of an art than a science. An optimist would argue that armed with enough rules, and a great deal of practice, we could express the integral of any reasonable continuous function in terms of familiar functions. Unfortunately, this is not true.

You may have encountered some of these functions. For instance, some calculators return the same integral when asked to compute ∫ ex2dx, while others return the answer images. This answer becomes circular when you find that the definition of the erf function, which is short for error function, is

images

Why is technology of no help when it comes to integrating this function analytically?

To answer this question, recall that an “elementary function” is a function that can be expressed in terms of power functions, exponentials, sines, cosines, and logarithms via the usual algebraic processes, including the solving (with or without radicals) of polynomials. Thus, elementary functions are all the “precalculus functions,” including polynomials and trigonometric and logarithmic functions. A theorem in mathematics asserts that there are an infinite number of elementary functions without an elementary antiderivative. Here are some examples of these functions:

images

The more common ones get their own names. For instance, we saw that up to some scaling factors, “erf” is the antiderivative of ex2. We can also find out using technology that “Si” is the antiderivative of sin x/x. Unfortunately, these functions are not exceptional.

What can we do when we need to integrate functions that are integrable but do not have elementary derivatives? If they are definite integrals, we could approximate them using Riemann sums, or we could use one of several other approximation schemes. In this section, we discuss five numerical schemes for approximating definite integrals. Three of these schemes, left endpoint rule, right endpoint rule, and midpoint rule, differ only in the manner that the sample points xi are chosen. A fourth scheme, the trapezoidal rule, and a fifth scheme, Simpson's rule, involve approximating the functions with piecewise linear and piecewise quadratic segments, respectively. These five schemes, three of which are depicted in Figure 5.35, differ in how rapidly they converge (as n → ∞) to the true value of the definite integral. These rates of convergence are described via error estimates.

images

Figure 5.35 Numerical estimates of the area under a curve on the interval [a, b] using the left endpoint, midpoint, and trapezoidal rules, as labeled.

Left and right endpoint approximations

We begin with the simplest approximation schemes, left endpoint approximation and right endpoint approximation. For presentation purposes, our discussion focuses on left endpoint approximation (see first panel in Figure 5.35). By analogy, similar statements apply to the right endpoint approximation.

Suppose f is a continuous function from x = a to x = b and we want to estimate images f(x)dx. By definition, images f(x)dx, is a limit of Riemann sums. Consequently, given n, we partition the interval [a, b] into n equal subintervals with endpoints:

images

where a1 = a + Δx, a2 = a + 2Δx,..., an = a +nΔx, and Δx = images. Taking the left endpoints, x1 = a0, x2 = a1,..., xn = an−1 as our sample points, we have

images

As a first example, we begin with a simple function, so that we can examine the error generated by taking an approximation.

Example 1 Using technology for a left endpoint approximation

We know

images

Use the left endpoint rule with n = 5, 10, 25, 50, and 100 to approximate this integral.

Solution We have f(x) = 3images/2, a = 1, and b = 4. We will show the detail for n = 5 and then use technology to generate other values. For n = 5, we have Δx = images = 0.6 and

images

Thus,

images

When we use technology to generate the approximation Ln, the other values for n are

images

Hence, the approximation given by technology appears to converging to the known value of 7.

We might have added another column to our answer for Example 1. What is the error of the approximation? That is,

images

We calculate the error for each of the n values in the 2nd table of Example 1 and find that the approximations are underestimates by 0.461, 0.228, 0.090, 0.045, and 0.023 for the respective entries in the table. As we would expect, the errors tend to decrease as n increases. This leads to two questions: First, how quickly do the errors decrease with n? Second, if we don't know the true value of the definite integral, how can we estimate the error?

To answer both questions requires introducing error bounds: upper bounds for the magnitude of the error. These upper bounds often involve understanding the derivatives of the integrand. For example, suppose we know for some constant K1 > 0 that |f′(x)| ≤ K1 for all x between a and b. The evaluation theorem implies that

images

for any point x in [a, a + Δx]. Since f′(u) ≤ K1, the dominance property of integrals implies that

images

for x between a and a + Δx. Equivalently,

images

Similarly, since f′(u) ≥ −K1, the dominance property of integrals implies

images

A graphical interpretation of these two inequalities is shown in Figure 5.36. The graph of f(x) above the interval [a, a + Δx] lies in a triangular wedge with area K1x)2. To estimate the error, we can consider three cases: the graph lies entirely in the upper half of triangular wedge (as illustrated in Figure 5.36), lies entirely in the lower half of the triangular wedge, or pass through both halves of the triangular wedge. When the graph lies entirely within one half of the wedge, it lies in a region with area Kx)2/2. Hence, the error in approximating imagesf(x)dx with f(x1x is less than or equal to K1x)2/2. On the other hand, if the graph of f(x) passes through both the upper and lower half of the triangular wedge, then over- and under-estimates partially cancel one another, and the error is less than if the graph lied entirely in either the upper or lower half of the triangular wedge.

images

Figure 5.36 Estimating errors for the left endpoint rule

Similarly, over any of the subintervals, the error in approximating the actual value with f(xkx is at most K1x)2/2. Let EL be the error by using a left endpoint approximation. Then, summing the error estimates over the n subintervals yields

images

where

images

is the error of the left endpoint approximation.

Example 2 Using the left endpoint rule

Consider

images

  1. Use technology to evaluate this integral.
  2. Use the left endpoint rule with n = 10 to approximate this integral.
  3. Give an error estimate for the approximation found for n = 10.
  4. Find the smallest value of n that ensures the left endpoint rule approximation has an error no larger than 0.001.

Solution

  1. Using technology, we get an estimate of 0.77265.
  2. Setting up a table of values for f(x) = sin(x2), a = 0, b = π, n = 10, and Δx = images (we leave the details for you), we obtain an estimate of 0.78997.
  3. To find an upper bound to the error, we need an upper bound to the derivative of f(x) = sin(x2) on the interval [0, π]. Since

    images

    and |cos(x2)| ≤ 1,

    images

    for x on [0, π]. Setting K1 = 2π, n = 10, a = 0, and b = π into the error bound yields

    images

    Hence, our estimate of 0.78997 is not guaranteed high-assured accuracy, despite it being reasonably close to the answer we found in part a.

  4. We want n such that EL ≤ 0.001:

    images

    Thus, we find n such that

    images

    Since n is an integer, we need to choose n = 31,007. Implementing n = 31,007 with technology, we obtain

    images

    Thus,

    images

    The original calculator answer is within this range of accuracy.

The preceding example illustrates several important points. First, even though we initially chose a reasonably large n (say, 10), the error bound was so large that we could not be certain about any of the digits. Second, an extremely large n is needed to ensure an estimate with accuracy to 0.001. Third, the approximation for the large n value was not very different from the estimate for the smaller n value of 10.

Midpoint rule

An alternative numerical approximation scheme is the midpoint rule in which the sample points chosen are the midpoints of each of the subintervals (see center panel in Figure 5.35). We begin, as we did with the left endpoints, by partitioning the interval [a, b] into n subintervals with endpoints:

images

where a1 = a + Δx, a2 = a + 2Δx,..., an = a + nΔx, and Δx = images. This time, we take the midpoints,

images

as our sample points to get

images

Associated with the midpoint rule is an error estimate. If |f″(x)| ≤ K2 for x on [a, b], then the midpoint error bound is

images

where

images

Example 3 Using the midpoint rule

Consider

images

  1. Use the midpoint rule with n = 10 to approximate this integral and compare it to the technology solution in the previous example.
  2. Provide an error estimate for the approximation found for n = 10.
  3. Find the smallest value of n that ensures the midpoint rule approximation will have an error no larger than 0.001.

Solution

  1. Setting up a table of values for the case f(x) = sin(x2), a = 0, b = π, n = 10, and Δx = images (we leave the details to you), we obtain the estimate 0.79918141, which is approximately 0.027 larger than the technology solution 0.77265 in the previous example.
  2. To find an upper bound to the error, we need an upper bound to the second derivative of f(x) = sin x2 on the interval [0, π]. Since

    images

    and |sin x2| ≤ 1 as well as |cos x2| ≤ 1,

    images

    for x on [0, π]. Setting K2 = 2 + 4π2, n = 10, a = 0, and b = π into the error bound yields

    images

    Hence, our estimate is accurate to within 0.54, which is substantially less than our error estimate of 6.2 with the left endpoint rule.

  3. We want n such that EM ≤ 0.001.

    images

    Thus, we find n such that

    images

    Since n is an integer, we need to choose n = 232. Compare this to the n = 31,007 that we needed for the left endpoint rule! Using midpoint rule with n = 232, we obtain

    images

    which we know is within 0.001 of the true answer.

If the midpoint rule differs from the left and right endpoint rules only by shifting the sample points by Δx/2, why does it do so much better? First, observe that both the midpoint rule and left endpoint rule integrate constant functions perfectly. Indeed, Ln and Mn for imagesc dx equal c(ba). Second, for the case of linear functions f(x) = cx + d on an interval [a, b], it is simple to show (verify this for yourselves!) that EM = 0 while EL = |c|(ba)/2. Hence, the midpoint rule integrates linear functions perfectly; consequently, it introduces error only for functions with nonzero, second-order derivatives. This explains why the error bound for the midpoint rule involves second-order terms (i.e., |f″(x)|and n2). In contrast, since the left endpoint rule introduces errors for functions whose derivatives are nonzero, the error bound for the left endpoint rule involves first-order terms (i.e.,|f′(x)|and n).

Another rule that perfectly estimates linear functions is the trapezoidal rule. This rule also has error bounds in terms of second-order rather than first-order derivatives. This rule is presented in the definition box at the end of this section and can be studied in more detail in Problem 38 of Problem Set 5.7.

Simpson's rule

As Example 3 illustrated, the midpoint rule was a significant improvement over the left endpoint rule because the error bound decreased like 1/n2 instead of like 1/n. The small price we had to pay for this improvement was that bounds are calculated in terms of the second rather than first derivative. With this sweet taste of success, can we do better? As it turns out, yes! There is another rule, Simpson's rule, for approximating integrals using parabolas to approximate the curve. This method requires breaking the interval into an even number n of subintervals. Let a = x0 < x1 = a + Δx < ··· < xn = b be the endpoints of these subintervals of width Δx = (ba)/n. The approximation is given by

images

Given |f(4)(x)| ≤ K4 for x on [a, b], the Simpson error bound is

images

where

images

Note that throughout this section we used the notation Kn to represent the bound of the nth derivative of f(x) for the cases n = 1, 2, and 4; that is,

images

This convention stresses the fact that each derivative has its own bound.

Example 4 How good is Simpson's rule?

Compare Simpson's rule with the left endpoint and midpoint rules in calculating the value of

images

Consider:

  1. The case where the interval is partitioned into n = 10 subintervals
  2. The smallest value of n that ensures the approximation is no larger than 0.001

From the results, what do you conclude?

Solution

  1. The approximation given by Simpson's rule for the case n = 10 can be calculated as outlined in the following table:

    images

    From technology, we know that the correct answer (to five decimal places) is I = 0.77265, and for n = 10:

    • The left endpoint rule approximation is I = 0.78997, yielding an error of 0.01732.
    • The midpoint rule approximation is I = 0.79918, yielding an error of 0.02653.
    • The Simpson's rule approximation is I = 0.79503, yielding an error of 0.02238.

    Thus, Simpson's rule, as expected, yields a smaller error than the midpoint rule. Rather surprisingly, however, both these rules are outperformed by the left endpoint rule, which in general is the least accurate of the three (see part b below)! This outcome is purely fortuitous in that positive and negative errors in the estimates obtained on each of the subintervals can sometimes cancel out to give a very good result for a generally inaccurate method.

  2. To find n such that ES ≤ 0.001, we need to solve for n in the inequality

    images

    where K4 is a bound on the fourth derivative of sin x2 on the interval [0, π]. Repeated differentiation yields

    images

    from which we conclude that K4 = (16π4 − 12) + 48π2 for x ranging over the interval [0, π]. Since in this case (ba) = π, it follows from the above inequality that the lower bound for n to ensure an error of less than 0.001 is

    images

    Solving this, we obtain n ≥ 43.05. Thus, selecting a value n ≥ 44 (remember for Simpson's rule n must be even) ensures that the accuracy of the estimate of I is better than 0.001.

    Again, from the two previous examples and the calculations here, we can see the requirements to ensure an accuracy of at least 0.001:

    • n ≥ 31,007 for the left endpoint rule
    • n ≥ 232 for the midpoint rule
    • n ≥ 44 for Simpson's rule

As the error estimate for Simpson's rule suggests from the foregoing example, Simpson's rule has much better convergence properties (i.e., the rate at which the error decreases with increasing n) than the midpoint rule. Why? Well, the Simpson's rule integrates a cubic function (i.e., third-order polynomials) perfectly. Hence, only nonzero fourth-order derivatives result in errors, and the error bound involves fourth-order terms (i.e., |f(4)(x)| and n4).

Example 5 Estimating crab harvests

The Dungeness crab (Cancer magister) is an important element of commercial fishing along the northern Pacific coast (California to Alaska). The data depicted in Figure 5.37 show the commercial harvest level of Dungeness crabs (in millions of pounds and excluding sport fishery and nontreaty landings) off the coast of Washington State for the period 1950 to 1999.

images

Figure 5.37 Crab harvests along the Pacific coast

A subset of these catch data are reported in the table below (unit of catch is in millions of pounds).

images

Use Simpson's rule to estimate the total amount of Dungeness crabs caught between 1950 and 1990.

Solution Applying Simpson's rule with n = 8 yields

images

For convenience, we next summarize the four methods of numerical integration described in this section. We include the trapezoidal rule (see the third panel in Figure 5.35 and Problem 38 in Problem Set 5.7).

Numerical Integration

Let f be continuous on [a, b]. Divide this interval into n equal parts:

images

Define Δx = images. The following rules provide estimates of images f(x)dx with error bounds given in terms of the bounding parameters |f(i)(x)| ≤ Ki for all axb.

LEFT ENDPOINT RULE

images

and

images

RIGHT ENDPOINT RULE

images

and

images

MIDPOINT RULE

images

and

images

TRAPEZOIDAL RULE

images

and

images

SIMPSON'S RULE (n is even)

images

and

images

A simple example reinforces how efficient Simpson's rule is compared with the left endpoint and midpoint rules in converging to a solution.

Example 6 Comparing the efficiency of the different methods

Consider

images

What is the smallest value of n needed to ensure that

images

Solution We have f(x) = x−1, f′(x) = −x−2, f″(x) = 2x−3, f″′(x) = −6x−4, and f(4) = 24x−5.

  1. The maximum value of |f′(x)| on [1, 3] is K1 = 1. Since |EL| ≤ images, we need

    images

    Hence, n = 20,000 will suffice.

  2. The maximum value of |f″(x)| on [1, 3] is K2 = 2. Since |EM| ≤ images, we need

    images

    Hence, n = 82 will suffice.

  3. The maximum value of |f(4)(x)| on [1, 3] is K4 = 24. Since |ES| ≤ images, we need

    images

    Since Simpson's rule requires an even number of intervals, n = 16 will suffice.

PROBLEM SET 5.7

Level 1 DRILL PROBLEMS

Approximate the integrals in Problems 1 to 12 using these four rules:

  1. left endpoint rule
  2. right endpoint rule
  3. midpoint rule
  4. Simpson's rule

images

Estimate the value of the integrals in Problems 13 to 22 to within the prescribed accuracy; use the rule indicated by the subscript on the error bound.

images

images

In Problems 23 to 28, determine how many subintervals are required to guarantee accuracy to within 0.00005 using these two rules:

  1. midpoint rule
  2. Simpson's rule

images

29. Estimate the area in the graph in Figure 5.38 using the left endpoint rule, right endpoint rule, and Simpson's rule with n = 6.

images

Figure 5.38 Estimate shaded area

30. Estimate the area in the graph in Figure 5.39 using the left endpoint rule, right endpoint rule, and Simpson's rule with n = 4.

images

Figure 5.39 Estimate shaded area

Level 2 APPLIED AND THEORY PROBLEMS

31. Governmental health agencies throughout the world monitor cases of human diseases. For example, the Office of Population Censuses and Surveys in the United Kingdom published weekly case reports about measles in England and Whales. Theoretical ecologist Ben Bolker published the data from 1948 to 1966 on the Web at http://ms.mcmaster.ca/bolker/measdata/ewmeas.dat The first 89 weeks at four week intervals is shown in the table below.

images

Use Simpson's rule and right endpoint rule to estimate the total number of cases over these 89 weeks.

32. Using inverse trigonometric functions, it can be shown that

images

Use this result to estimate π correct to four decimal places by applying Simpson's rule to this integral and using the appropriate error estimate.

33. Black Plague revisited Recall that for the outbreak of the plague in Bombay in 1905–1906, the mortality rate due to the plague was approximated by Kermack and McKendrick with the function

images

deaths per week.

  1. Write a definite integral that represents the number of deaths that accumulated from t = 0 to t = 30.
  2. Estimate the definite integral using Simpson's rule with n = 10.

34. The data set for 80 hours of the discharge (in m3/s) for the Raging River is shown here:

images

The data set for the first 24 hours is summarized in the following table:

images

Use Simpson's rule to estimate the total amount of discharge in the first 24 hours.

35. Sweet corn has a lower development threshold of 50°F and requires 1587 degree-days to complete development. On July 3, 2006, the temperatures in northern Illinois were as follows (measurements performed by the Northern Illinois Agronomy Research Center).

images

  1. Using the right endpoint rule, estimate the number of degree-days that elapsed on this summer day.
  2. If the temperatures on July 3rd typified the temperatures throughout the summer, estimate how many days it would take sweet corn to mature.

36. Repeat Problem 35 using Simpson's rule. Do you expect your answer to be more or less accurate than the answer to Problem 35?

37. The weekly rate of cases of influenza A (strain unknown) studied by WHO/NREVSS during the 2003–2004 season is plotted here:

images

Estimate the total number of cases (i.e., the area under the curve) from week 40 to week 56 using Simpson's rule on two-week intervals.

38. Trapezoidal rule for numerical integration: Assume that the area under a curve f(x) on an interval [a, b] is approximated by n trapezoids made up from rectangles topped by right-angle triangles so that the left- and right-hand sides of the (k + 1)th trapezoid over the interval [xk, xk+1], are of heights f(xk) and f(xk+1), respectively, as illustrated in Figure 5.35. Let a = x0 < x1 = a + Δx < ··· < xn = b and Δx =(ba)/n. Show that the approximate area is given by

images

For completeness, we note that the error

images

satisfies

images

where K2 satisfies |f″(x)| ≤ K2 on the interval [a, b].

39. images

images

Takakazu Seki Kōwa was born in Fujioka, Japan, the son of a samurai, but was adopted by a patriarch of the Seki family. Seki invented and used an early form of determinants for solving systems of equations. He also invented a method for approximating areas that is very similar to the rectangular method introduced in this section. His method, known as the yenri (circle principle), found the area of a circle by dividing the circle into small rectangles, as shown in Figure 5.40 (drawn by a student of Seki).

images

Figure 5.40 Early Asian calculus

Data Source: From Sawguchi Kazuyuki kokon Sampoki 1670. A History of Japanese Mathematics David Eugene Smith and Yoshio Mikami

Draw a circle with radius of 10 cm. Draw vertical chords through each centimeter on a diameter (you should have 18 rectangles). Measure the heights of the rectangles and approximate the area of the circle by adding the areas of the rectangles. Compare this method with the formula for the area of this circle.

40. images

images

Isaac Newton invented a preliminary version of Simpson's rule. In 1779, Newton wrote an article to an addendum to Methodus Differentials (1711) in which he gave the following example: If there are four ordinates at equal intervals, let A be the sum of the first and fourth, B the sum of the second and third, and R the interval between the first and fourth; then the area between the first and fourth ordinates is approximated by images(A + 3B)R. This is known today as the Newton-Cotes three-eighths rule, which can be expressed in the form

images

Roger Cotes and James Stirling (1692–1770) both knew this formula, as well as what we call in this section Simpson's rule. In 1743, this rule was rediscovered by Thomas Simpson (1710–1761).

Estimate the integral

images

using the Newton-Cotes three-eights rule, then compare with an approximation made using left endpoints (rectangles) with n = 8. Which of the rules gives the most accurate estimate?

5.8 Applications of Integration

In the preceding sections of this chapter, we motivated definite integration with area under a curve and accumulated change. In this section, we consider some additional applications that use Riemann sums to formulate integrals for survival and renewal processes, estimate cardiac output, and compute the amount of work required to perform a task.

Survival and renewal

Survival and renewal is the study of a population, or group of individuals, with the goal of predicting the size of the group at some future time. In the following example, a survival function gives the fraction of individuals in a group, or population, that can be expected to remain in the group for any specified period of time. In addition, a renewal function gives the rate at which new members arrive. Survival and renewal problems arise in many areas of study, including sociology, ecology, demography, and even finance—where the “population” is the number of dollars in an investment account, and “survival and renewal” refer to results of an investment strategy.

Example 1 Survival and renewal in a clinic

A new county mental health clinic has just opened. Statistics from similar facilities suggest that the fraction of patients who will still be receiving treatment at the clinic t months after their initial visit is given by the survival function s(t) = et/20. The clinic initially accepts 300 people for treatment and plans to accept new patients at the rate of 10 per month. Approximately how many people will be receiving treatment at the clinic 15 months from now?

Solution Since e−15/20 is the fraction of patients whose treatment we expect to continue for at least 15 months, it follows that of the current 300 patients, only 300e−15/20 ≈ 141.7 will still be receiving treatment 15 months from now.

Each month, however, 10 new patients enter, and some of these will also still be around at month t = 15. To account for this, we divide the 15-month time interval [0, 15] into n equal subintervals, each of length Δt = images months. Let tk = (k − 1)Δt denote the beginning of the kth subinterval for k = 1,..., n. Since new patients are accepted at the rate of 10 per month, the number of new patients accepted during the k th subinterval is 10Δt. When Δt is small, we can estimate 15 − tk to be the time that elapses for all of these patients by the 15th month. Consequently, approximately

images

of these patients will still be receiving treatment 15 months from now. Thus, the total number of patients arriving at times t1, t2,..., tn who are still receiving treatment at time t = 15 is approximated by the sum

images

As n → ∞, we obtain the integral

images

which is also referred to as a renewal function.

Adding this integral to the 141.7 original patients still receiving treatment after 15 months, the total number of patients receiving treatment at time t = 15 is

images

That is, 15 months from now, the clinic will be treating approximately 247 patients.

This example provides a guide to developing a more general formulation for survival and renewal processes. Suppose a population initially has N0 individuals and receives new individuals (renews) at a rate r(t), and suppose the fraction of individuals remaining (surviving) in the population after t units of time after entering the population is s(t). If we want to determine the number of individuals in the population at time T, we can divide the interval [0, T] into subintervals of width Δt = T/n. The number of individuals arriving into the population during the kth time interval is approximately r(tkt. The fraction of these r(tkt individuals surviving to time T is approximately s(Ttk). Hence, the number of individuals entering during the kth time interval and surviving to time T is approximately s(Ttk)r(tkt. Summing up over all these time intervals yields

images

Taking the limit as n → ∞ yields

images

Of the N0 individuals who were initially present, s(T)N0 of them survive to time T. Hence, the number of individuals in the population at time T is given by the following survival and renewal function.

Survival and Renewal Equation

Suppose there are N0 individuals initially present, a fraction of s(t) individuals survive a period of length t, and individuals arrive at a rate of r(t) individuals per unit time at time t. Then the total number of individuals present at time T is given by the survival renewal equation

images

Example 2 Fire ants

The imported fire ant (Solenopsis invecta) (Figure 5.41) is a pest in both urban and rural areas. Damage estimates for the United States range in the millions of dollars.

images

Figure 5.41 Fire ants were imported from South America

The fire ant has colonies in which workers live approximately 10 to 70 weeks and queens survive for about seven years. A single colony can have from 10 to 100 or more queens, each producing 1,000 to 1,500 eggs per year for 7 years. Suppose a colony is formed with 100 queens, in which each queen produces workers at a rate of 1,250 + 250 sin(2πt) per year and in which the fraction of workers living t years after their birth is given by s(t) = e−1.25t. Find the number of workers in the colony seven years from now, assuming all 100 queens survive the seven years under consideration.

Solution Initially there are no workers and N0 = 0. The rate at which workers are renewed is

images

The survival function is given by s(t) = e−1.25t. Setting T = 7 into the renewal equation

images

The renewal equation predicts that we should expect the colony to have around 100,000 workers seven years from now. It is possible to do the prior integral without technology: separate the integral into two integrals, use substitution with u = −1.25(7 − t) in the first integral, and use integration by parts (twice) for the second integral.

Another application of the renewal equation is in the area of finances, where we are concerned with the growth of economies or our personal fortunes. The key difference in this application is that instead of calculating how capital (population of dollars) decays, we are interested in how capital grows.

Although it is not common practice in banking systems to compound continuously, it is a reasonable approximation for daily compounding. Recall from Example 6 of Section 1.4 that compounding continuously at a rate of c% per year implies that if you put N dollars into an account, then t years later there will be ect/100N dollars in the account.

Example 3 Saving for retirement

Starting at age 20, Peggy Sue puts money into a retirement account at a rate of $2,000 per year. The money in this account is compounded continuously at a rate of 10% per year. How much money will be in her account when she turns 60? How much would she have if she starts at the age of 30?

Solution To determine the total amount in Peggy Sue's account, let us break up the time interval [0, 40] into n subintervals of width Δt = 40/n. The amount of money she puts into the account during the kth time interval [(k − 1)Δt, kΔt] is approximately 2,000Δt. Over the 40− kΔt year period, this money grows to approximately e0.1(40−tk) 2,000Δt where tk = kΔt. Hence, the total amount of money she has at age 60 is approximately

images

Taking the limit as n → ∞ yields

images

Integrating yields

images

Peggy Sue will be millionaire! Alternatively, if she started saving at age 30, then at age 60 she would have

images

Not even close to being a millionaire!

The solution to Example 3 shows that if money is being added to an account at a rate of r(t), the account is continually compounded at an interest rate of c%, and there is initially N0 dollars in the account, then the total amount in the account T years from now is

images

This is just another survival-renewal equation with s(t) = ecT/100.

Cardiac output

Cardiac output is the volume of blood pumped by the heart in a specified interval of time. Estimating cardiac output is important, because it is an indicator of certain heart diseases.

images

Figure 5.42 Schematic of the heart

Cardiac output can be measured using dye dilution. A known quantity of dye, say D milligrams (mg), is injected into a main vein near the heart. This dye circulates with the blood through the body (from the right ventricle of the heart to the lungs to the left ventricle and into the arteries) and returns to the left ventricle (see schematic in Figure 5.42). The concentration of the dye, c(t) milligrams per liter (mg/l), passing through an artery is monitored. To compute cardiac output from these recorded concentrations, assume that the cardiac output (i.e., blood flow) remains at a constant rate, F liters per second (l/s), during the experiment. The rate at which dye is passing through the artery at time t seconds is given by Fc(t) milligrams per second (mg/s). Notice how the units work out here: c(t) has units mg/L and F has units L/s. Hence c(t)F has units

images

Assume that the entire amount of dye passes through the artery between time t = 0 and t = T. The net amount of dye passing through the artery over the time interval from 0 to T is

images

By conservation of mass, the net amount of dye observed must equal the initial amount of dye, D; that is,

images

Solving for the cardiac output, F, yields

images

Example 4 Dye dilution

A (hypothetical) patient is given an injection of 5 mg of dye. The measured concentrations of dye are recorded in the following table.

images

images

Figure 5.43 Dye concentrations in the heart after an injection

A plot of these data is given in Figure 5.43. Use Simpson's rule to estimate the cardiac output of the patient.

Solution We want to use Simpson's rule with Δt = 1. However, we have 22 data points and consequently 21 (an odd number) of intervals. Since Simpson's rule requires an even number of intervals and c(21) = 0, we omit the last data point and make the following approximation:

images

Since the initial amount of dye is 5 mg, we get the following result for cardiac output:

images

Example 5 Thermodilution

Another approach to measuring cardiac output involves use of a pulmonary artery catheter that allows rapid, easy measurements of cardiac output using thermodilution. The principle of thermodilution is the same as that of dye dilution. Instead of injecting dye, doctors inject 10 ml of a cold dextrose solution. As the cold solution mixes with the blood in the heart, the temperature variations in the blood leaving the heart are measured. A hypothetical temperature variation curve may be described by the function

images

This curve is plotted in Figure 5.44, with this interpretation: for a cold solution, f(t) is the amount the temperature drops below normal body temperature.

Assuming that the temperature of a body is 37°C and the temperature of the dextrose solution is 0°C, estimate the cardiac output of a patient over a one-minute time interval.

images

Figure 5.44 Temperature variation in the heart resulting from an injection of cold dextrose (The interpretation here is that the peak represents 0.6°C below normal body temperature.)

Solution This example is just like the previous example but replacing dye concentration with temperature variation. The initial “amount of cold” (the equivalent of the initial amount of dye) is given by

images

If the cardiac output is F, then the rate of “cold” passing by at time t is

images

The accumulated change in cold is F × images(t)dt mL-°C which must equal 370 mL-°C. Hence,

images

Applying integration by parts (twice) and evaluating yields

images

We can convert units to get 49.95 ml/s ≈ 3.00 L/min.

Work

How much pasta should a person eat to dig a posthole? How much energy should a grizzly bear expend to dig out a pocket gopher? To answer these questions, we need to understand the relationship between work and energy. To complete any work, we need energy. A standard unit of energy is a calorie. A calorie is the amount of energy required to heat one gram of water one degree Celsius. What does this mean?

In the article “What is Energy?” in Exploring Food Magazine (vol. 14 #4 1990), Paul Doherty writes about his undergraduate days as a biophysics student at MIT and describes an experiment where the professor set a peanut on fire. The energy from this peanut was sufficient to heat a test tube filled with 10 grams of water from room temperature to 100° Celsius, and then boil away 2 grams of water.

Doherty calculated that amount of energy required to heat and then boil the water. To heat 10 grams of water from 20° C to boiling (100° C) requires (100 − 20) × 10 = 800 calories. Further, since 540 calories are needed to boil away a gram of water (at sea level), another 1080 calories were needed to vaporize 2 grams of water. Thus, in the demonstration, the burning peanut delivered 1880 calories of heat flow to the test tube of water. This might seem like a lot of calories for one peanut. The reason for this surprising number is that a single dietary calorie (i.e., the type of calorie that is reported with food), which is abbreviated Cal or simply C, is actually one kilocalorie, abbreviated kcal. Hence, a peanut typically represents 1.5 to 2 Cal of energy.

What can we do with all this energy once ingested? We can turn it into work.

Work Done by a Constant Force

If a body moves a distance d in the direction of an applied constant force F, the work, W, done is

images

From Newtonian physics we know that

images

which implies that units of force are mass × distance × time−2. Hence, the units of work are mass × distance2 × time−2. A standard work unit, called a Joule(J), is defined as follows:

images

The conversion between joules and dietary calories is given by the formula

images

Example 6 Calories used while working out

How much work is done lifting 30 kilograms 20 meters? (This is equivalent to 40 arm curls lifting about 67 pounds). Give your answer in joules and in calories.

Solution To solve this, we make use of the fact that on the surface of Earth, the force of gravity on 1 kilogram of mass is approximately 9.81 kg-m/s2. Thus,

images

All that work, and so little to show for it? Well, actually, we are not 100% efficient at translating calories from food to work. Roughly, humans have 10% efficiency (all that overhead from maintaining body temperature, etc.). Thus, we might estimate the number of calories being burnt off as 14 Cal. That is, we need about 7 to 10 peanuts to carry out this work.

Example 7 Climbing mountains with a candy bar

The article “What is Energy?” in Exploring Food Magazine (vol. 14 #4 1990) reveals that a Milky Way bar contains more energy than a stick of dynamite. The candy bar contains 270 Cal. If the energy from the Milky Way bar is used with 100% efficiency, determine how high (in meters) a 70 kg human could be lifted with the energy from the Milky Way bar.

Solution First find the number of joules:

images

The amount of work required to lift a 70 kg person x meters is

images

Thus,

images

This is almost twice the height of the cliff face of Yosemite's El Capitan (see Figure 5.45). No stick of dynamite can do that! In fact, an ounce of dynamite produces only one quarter as many calories when it explodes as an ounce of sugar does when it is burnt!

images

Figure 5.45 El Capitan, Yosemite

Often people try to achieve great things by moving huge loads of rocks and dirt, such as digging canals, cutting through mountains to build roads and railways, or hollowing out deep cellars to build skyscrapers. Calculus can be used to compute the amount of work required to accomplish such feats, as evidenced in the following example.

Example 8 Digging a cellar

Consider the problem of digging a cellar that it is 7 m deep, 100 m long, and 50 m wide. How much work, in theory, does it take just to remove all the dirt out of the hole being dug?

Solution To answer this question, we use the fact that the density of soil is approximately the same density as water. In this case, one cubic centimeter (i.e., milliliter) has a mass of one gram. Since our problem is phrased in meters and kilograms, we need to translate this statement into these units. Since one cubic meter equals 1003 cm3, one cubic meter of soil has a mass of approximately 1,000,000 g = 1,000 kg (one metric ton).

The amount of work required to lift one scoop of dirt to ground level depends on the depth of that scoop of dirt. Dirt at the bottom of the cellar has to be lifted higher than dirt at the top of the cellar. To find the amount of work, envision cutting the cellar up into n thin horizontal slices of thickness Δx = 7/n (see Figure 5.46). Let x denote the depth of a slice in meters.

images

Figure 5.46 Slicing a cellar into horizontal slices

The volume of a slice with thickness Δx is

images

The mass of this slice is given by

images

The weight of this slice is given by

images

If this slice is at depth x meters, then the work required to lift the slice is

images

If the depths of the slices are x1, x2,..., xn, then the work to dig the cellar is the sum of the work to lift all of the slices, approximately given by

images

Letting Δx get smaller and smaller should yield better and better approximation. Consequently, taking the limit as n → ∞ yields

images

This is equivalent 287,219 Cal—at last a way to burn off a considerable number of calories!

The next example computes the amount of work required to remove soil from a geometrically more complicated region.

Example 9 A hungry grizzly

Since the early 1980s, Steve and Marilynn French, founders of the Yellowstone Grizzly Foundation, have been watching grizzlies in Yellowstone National Park (see Figure 5.47). They report watching a bear digging a trench twenty feet long to get a little gopher as a tasty treat. Is the effort worth all the energy expended?

Assuming that the trench has a semicircular cross-section with radius 1 m and that the density of soil is 1,000 kg/m3, find the amount of work performed by the grizzly bear.

images

Figure 5.47 A gopher treat for Spirit? Spirit is the first Montana grizzly to reside at the Grizzly & Wolf Discovery Center.

Solution We convert 20 feet to meters to obtain ≈ 6.1 m. To determine the approximate amount of work, we slice the trench into n slices of thickness Δx meters (see Figure 5.48).

images

Figure 5.48 Grizzly's trench

To determine the width w of a slice at depth x meters, we use the fact that the cross-sectional profile of the trench is a semicircle of radius 1. Thus, (w/2)2 + x2 = 1 so that

images

The volume of a slice at depth x meters is approximately

images

The weight of the slice is approximately

images

The amount of work to lift a slice at depth x is

images

If x1, x2,..., xn are depths of the n slices, then the total work is approximately

images

Taking the limit as n → ∞ yields

images

The answer of nine to ten calories seems surprisingly few calories! Of course all this presumes that the grizzly was able to perform the work 100% efficiently, which is certainly not the case. For instance, if the grizzly worked with 5% efficiency, then the bear used about 200 Calories—the amount of Calories in one Twinkie.

PROBLEM SET 5.8

Level 1 DRILL PROBLEMS

After reviewing the mental health clinic in Example 1, for Problems 1 to 6 calculate the number of patients in the clinic after 15 months if the patient survival rate s(t) and the renewal rate r(t) are as given.

1. s(t) = et/20 and r(t) = 20 per month. How did doubling the renewal rate change the answer from what we found in Example 1?

2. s(t) = et/40 and r(t) = 10 per month. How did halving the survival rate change the answer from what we found in Example 1?

3. s(t) = et/10 and r(t) = 20 per month

4. s(t) = et/40 and r(t) = 20 per month

5. s(t) = et/20 and r(t) = 10 + t per month

6. s(t) = images and r(t) = 10 per month

After reviewing the fire ants in Example 2, for Problems 7 to 12 calculate the number of workers after five years if the worker survival rate s(t) is as given. Use technology to numerically evaluate the integrals.

7. s(t) = e−0.625t, that is, survival rate is doubled

8. s(t) = e−2.5t, that is, survival rate is halved

9. s(t) = images

10. s(t) = e−1.25t and, in addition, the proportion of queens alive at time t is q(t) = e−0.1t

11. s(t) = e−1.25t and, in addition, the proportion of queens alive at time t is q(t) = e−0.2t

12. s(t) = e−1.25t and, in addition, the proportion of queens alive at time t is q(t) = e−0.05t

For Problems 13 to 16, reconsider Example 3. Calculate the amount of money Peggy Sue will have in her account by age 60 if she adds A dollars per year to her account, she opens her account at age B years, and the annual interest rate is 10%.

13. A = 4,000 and B = 20

14. A = 4,000 and B = 30

15. A = 1,000 and B = 10 (she starts really young!)

16. A = 10,000 and B = 40 (she starts very late!)

Answer Problems 17 to 19 by finding the work done; express your answer using the unit of foot-pounds (ft-lb).

17. Lifting a 90 lb bag of concrete 3 ft

18. Lifting a 50 lb bag of salt 5 ft

19. Lifting a 850 lb billiard table 15 ft

Level 2 APPLIED AND THEORY PROBLEMS

20. Analysts speculate that patients will enter a new clinic at a rate of 300 + 100 sinimages individuals per month. Moreover, the likelihood an individual is in the clinic t months later is et. Find the number of patients in the clinic one year from now.

21. A patient receives a continuous drug infusion at a rate of 10 mg/h. Studies have shown that t hours after injection, the fraction of drug remaining in a patient's body is e−2t. If the patient initially has 5 mg of drug in her bloodstream, then what is the amount of drug in the patient's bloodstream 24 hours later?

22. Consider a mental health clinic that initially has 300 patients, and accepts 100 new patients per month; the fraction of patients receiving treatment for t or more months is given by f(t).

images

Using Riemann sums with left endpoints, estimate the number of patients in the clinic after 12 months.

23. Consider the following two scenarios involving an IRA account that yields 9% continuous interest.

  1. You graduate from college at age 22, get a job, and open an IRA account. You deposit $1,000 per year until age 65. How much money is the account at age 65? How much money did you pay into this account?
  2. You graduate from college at age 22 and do not bother to start an IRA account until you reach 32. Then you deposit $2,000 per year into the IRA account until you reach age 65. How much money is in your IRA account at age 65? How much money did you pay into this account?

24. The administrators of a town estimate that the fraction of people who will still be residing in the town t years from now is given by the function S(t) = e−0.04t. The current population is 20,000 people and new people are arriving at a rate of 500 per year.

  1. What will be the population size 10 years from now?
  2. What will be the population size 100 years from now?

25. After 5 mg of dye is injected into a vein, we obtain the concentration levels shown in the following table. The variable t is in seconds and c(t) is in mg/liter. Using Simpson's rule, compute the cardiac output.

images

26. Sediment flow.* Methods used in the measuring of cardiac output can be applied to other situations. For example, ecologists and scientists are interested in how much sediment is moved by a river. Data on the water flow and suspended sediment in the Des Moines River near Saylorville Lake, Iowa, is given in the following table. Using Simpson's rule, compute the total amount (kilograms) of suspended sediment that passed the measurement point for the period ending December 15, 1993.

Des Moines River Basin Water Discharge Records (December 1993)

images

27. Kety-Schmidt technique.* Seymour Kety and Carl Schmidt described a widely acknowledged and accurate method for the determination of cerebral blood flow and cerebral physiological activity (e.g., metabolic rate of oxygen). For example, a patient breathes 15% nitrous oxide (N2O). After the start of administration, the arterial concentration, A, is measured in the radial artery. This is the concentration before the blood enters the brain. The venous concentration, V, is measured at the base of the skull in the superior bulb of the internal jugular vein (at the point of exit of the jugular vein from the brain). This process for measuring the blood flow of cerebral physiological activity is commonly referred to as the Kety-Schmidt technique. A sample table is shown below.

images

  1. Initially, the rate of N2O flow into the brain is greater than the flow out of the brain. Moreover, after approximately ten minutes, the concentrations flowing into the brain and from the brain are approximately equal. The brain has become saturated with N2O. Assuming a constant cerebral blood flow rate, F, use Simpson's rule to estimate the total amount of N2O accumulated in the brain during ten minutes. Your answer will depend on F.
  2. Through other means, the maximum amount of N2O in the brain can be measured. Suppose that the maximum amount is determined to be 58.8 cc. Determine F.

28. A bucket weighing 75 lb when filled and 10 lb when empty is pulled up the side of a 100 ft building. How much more work in foot-pounds is done in pulling up the full bucket than the empty bucket?

29. A 20 ft rope weighing 0.4 lb/ft hangs over the edge of a building 100 ft high. How much work in foot-pounds is done in pulling the rope to the top of the building? Assume that the top of the rope is flush with the top of the building and that the lower end of the rope is swinging freely.

30. How much ice water do you need to ingest to burn off 300 calories? Assume your body temperature is 37°C and the energy required to digest ice water is the energy needed to raise the ice to body temperature.

31. A children's book, a steam shovel claims that she can do as much work in one day as 100 men can do in seven days. In a modern society we assume that the work needed to dig the cellar considered in Example 8 is done by a machine. But how many calories would 100 men produce in seven days, if we assume that each ate two pounds of pasta a day and worked with 10% efficiency? How does this compare with the work done (i.e., calories needed) to dig the cellar considered in Example 8? Assume a serving of pasta is two ounces and contains 200 Cal.

32. Determine the length of a rectangular trench you can dig with the energy gained from eating one Milky Way bar (270 Cal). Assume that you convert the energy gained from the food with 10% efficiency and that the trench is 1 meter wide and 1 meter deep. Assume the density of soil is 1,000 kg/m3.

In the next two problems, use the fact that a serving of pasta contains 200 Cal and the density of soil is 1,000 kg/m3.

33. How much work does it take to dig up a conical hole of depth 5 meters and diameter 6 meters? How many servings of pasta are required to complete this work, assuming the energy from the pasta is converted with 5% efficiency to work?

34. How much work does it take to dig a hemispherical pit with radius 10 meters? How many servings of pasta are required to complete this work, assuming the energy from the pasta is converted with 5% efficiency to work?

CHAPTER 5 REVIEW QUESTIONS

  1. Find the general antiderivative of f(x) = images.

    Evaluate the definite integrals in Problems 2 to 5.

  2. images
  3. images
  4. images
  5. images
  6. Find the area under the curve y = images over [1, 2].
  7. The slope F′(x) = images at each point is shown in Figure 5.49.

    images

    Figure 5.49 Slope field

    Find F passing through (1, −2) both graphically and analytically.

  8. The Royalty rose has a lower developmental threshold of 41.4°F and requires 473 degree-days for harvesting time. If the temperature were to remain a constant 72°F, how long would it take for this rose to mature?
  9. Find images where

    images

  10. Evaluate the following integral

    images

  11. Consider a mental health clinic that initially has 450 patients and accepts 150 new patients per month; the fraction of patients receiving treatment for t or more months is given by f(t):

    images

    Using Riemann sums with left endpoints, estimate the number of patients in the clinic after 12 months.

  12. Find an upper bound for

    images

  13. The rate of infection of a disease in a population of 10,000 is given by the function

    images

    where t is the time in months since the disease broke out.

    1. Use technology to plot R(t). Why is this a reasonable description of a disease spreading in a population?
    2. Compute the number of people infected by the disease by time T.
    3. Use technology to approximate the time when 50% of the population have the disease.
  14. In a wild week of temperature fluctuations, the temperature in Corvallis, Oregon, is given by

    images

    where t is measured in days. Find the number of degree-days that have elapsed for a beet armyworm over the first week. Note: The lower developmental threshold of a beet armyworm is 54°F.

  15. Express images tan x dx as the limit of a Riemann sum using right endpoints.
  16. Express images as a definite integral.
  17. Find images
  18. Suppose images
  19. Use the geometrical interpretation of the definite integral to find images(1 − |x|)dx. Be sure to provide a sketch.
  20. A stone was dropped off a tower and hit the ground at a speed of 200 ft/s. What was the height of the tower?

GROUP PROJECTS

Seeing a project through on your own, or working in a small group to complete a project, teaches important skills. The following projects provide opportunities to develop such skills.

Project 5A Physiological Time

Section 5.1 introduced the concept of developmental thresholds that defined the range of temperatures over which plants and poikilothermic animals (those without an internal mechanism for maintaining their body temperature within a narrow range of values, as in homeothermic birds and mammals) grow and develop. In Section 5.2 this idea was articulated further through the concept of physiological time, as measured through the accumulation of heat units called degree-days. The number of degree-days that accumulate over time for a given temperature profile is the area under the curve of this profile between the lower and upper thresholds, as illustrated in Figure 5.50. Note that a lower threshold is always needed, to bound the area from below, but the calculated area is either bounded above by the temperature curve itself or by an upper threshold, depending on which is the minimum for the time in question. Thus, a lower threshold is always needed, but an upper threshold is only included as a refinement of physiological time as a model for estimating the growth and developmental rates of plants and poikilotherms.

What if we could continuously measure the temperature in an orchard, for example, from the time of bud break (the first buds appear on otherwise bare trees), and we knew how many degree-days between minimum and maximum thresholds were needed until the trees come into blossom? If so, we could predict—using anticipated weather patterns—the expected date for the occurrence of blossoms and make sure we have honeybee hives in the orchard in sufficient time to anticipate this event. Thus, the calculation of degree-days can help growers optimize their use of honeybee pollinators, schedule harvest activities, and so on.

images

Figure 5.50 Solid line represents the continuous temperature that a plant or poikilotherm experiences, and shaded area represents the accumulated degree-days that the organism in question will experience subject to development being arrested above and below the upper and lower thresholds, respectively.

It is generally not possible, or even desirable, to continuously monitor the temperature of an orchard. Most growers have temperature gauges that record only the maximum (max) and minimum (min) temperatures each day. These data can be used to generate a degree-day calculation based on the assumption that the maximum and minimum temperatures occur 12 hours apart (as idealized in Figure 5.50) using an appropriate function (i.e., model) for interpolating the temperature between each consecutive pair of max-min and min-max temperatures. If a linear function is used, and only a lower threshold is assumed, the method is equivalent to constructing a sequence of right-angle triangles with either a rectangular piece added below when the minimum temperature is above the threshold; or, the base of the triangle is raised for the case when the minimum temperature is below a lower threshold (as illustrated in Figure 5.51).

images

Figure 5.51 The thick, irregular red line represents a hypothetical temperature profile that oscillates like a distorted sine wave, so that in every 24-hour period it has a maximum and a minimum value. The thin line is a linear interpolation between these maximum and minimum values. The shaded quadrilaterals, plus intervening nonshaded quadrilaterals—all with their bases defined by the lower threshold temperature (dotted line: note the upper threshold is above the max temperature in all cases and so does not apply)—are the accumulation of degree-days between consecutive min-max temperatures and max-min temperatures, respectively. This method of accumulating degree-days is called the double triangle method because two different triangular-looking quadrilaterals are used in every 24-hour cycle.

  1. Use the double triangle method illustrated in Fig. 5.51 to calculate the number of degree-days accumulated over a three-day period in which the minimum and maximum temperatures in degrees Celsius are T = {(5, 23), (7, 22), (4, 26), (5, not measured)} and the lower threshold is 0°C with no upper threshold assumed to exist.
  2. Using the double triangle method, recalculate the number of degree-days accumulated when the lower threshold is 5°C, first for the case when there is no upper threshold, then when the upper threshold is 30°C, and finally when the upper threshold is 25°C.
  3. Instead of using a line to interpolate between min and max temperatures, use the rising first quarter phase of a sine function to interpolate between the given min and max and the falling second quarter of a sine function to interpolate between the max and min temperatures. This method is referred to as the double sine method (Figure 5.52).

    images

    Figure 5.52 The thick irregular red line represents a hypothetical temperature profile. The thin line, instead of being a linear interpolation as depicted in the double triangle method, is a quarter sine wave interpolation (of 12 hour duration) between each min-max and max-min pair of temperatures. This method of accumulating degree-days, is called the double sine method.

  4. Repeat step 2 using the double sine method and compare your results with the double triangle method.
  5. Use your precalculus knowledge of algebra and geometry to write a general expression for the number of accumulated degree-days under the double triangle method when the temperature profile is T = {(m1, M1), (m2, M2),···, (mn, Mn)} and the minimum and maximum developmental thresholds are k and K, respectively.
  6. Use your knowledge of integral calculus to repeat the exercises and write a general expression for the double sine method.
  7. Find a real data set on the Web of daily max and min temperatures that spans a period of several months (if you find a longer data set, select a subset); use double triangle and double sine formulas, implemented in your favorite technology (e.g., a spreadsheet application, Mathematica, Maple, or some other programming language) to calculate the number of degree-days progressively accumulating each day from the start to end date of your data, if the lower and upper thresholds, respectively, are equal to the average min and average max over the data. Plot these results on a graph of accumulated degree-days to date to provide a visual sense of how much the two methods differ over time.

Project 5B Life Histories and Population Growth (challenging!)

Every biological species has a life history characterized by two functions: the mortality function l(x) and natality function b(x). The interpretation of the first function is that l(x) represents the proportion of individuals in a large population that survive to age x or, in a small population the probability that an individual survives to age x. As we will see in Chapter 7, when we look at the relationship between integration and probability theory, l(x) also represents the probability that any given individual will survive until age x (which can be a fractional number). Thus, l(30) = 0.2 implies that only 20% of individuals in a population will survive until age 30 or that any particular individual has a 20% chance of surviving until age 30. Note that we don't have to use years as our unit of time. In the case of fruit flies, for example, a more appropriate measure of age is weeks or days.

The function b(x) represents a force of natality which only has a clear meaning in terms of being integrated over some nonzero age interval x. For example, imagesb(x)dx = 3.5 implies that each individual in its third period of life (i.e., from age 2 to age 3) is expected (on average) to produce 3.5 offspring. If these are sexually reproducing organisms, then this implies that any male-female pair in its third year of life is expected to produce 7 offspring.

The theory we are about to explore assumes that either the species is clonal, or males and females have the same life histories, or males and females have different gender-specific life histories but only females are considered. In the latter case, b(x) is interpreted as the force of natality of female progeny per reproducing female of age x—that is, the statement imagesb(x)dx = 3.5 implies that each female in her third period of life is expected to have 3.5 daughters. Of course some might have 0 and others might have 10, but the average for the age range in question is 3.5.

Demographers have shown, under assumptions of stationarity (a technical term that requires more advanced concepts than we have to define it, but can be loosely thought of as a population that has an unchanging age-structure over time), that the quantity

images

represents the number of individuals being born for every individual that dies, given that no individual lives beyond age xmax. This implies that the population is growing if R0 > 1 and declining if R0 < 1. Further, demographers have shown that this rate of growth or decline is equivalent to the mathematical statement that Nt+G = R0Nt, where G is the length of a generation which is given by the integral

images

Thus, if we rescale time so G = 1, then this model implies that the population will have grown from an initial size N0 to a size Nm = images after m generations.

  1. If the proportion of individuals that die each time period in an age-specific cohort (i.e., a group of individuals of the same age) is independent of their age, then the mortality schedule (curve, function) for the species in question is said to be Type II. Demonstrate that the form

    images

    is a Type II mortality curve on [0, xmax] for some constant r > 0.

  2. Species are said to have Type I mortality schedules if mortality rates are much higher in immature than mature individuals (except, of course, for the very old). Type III is the reverse case. By searching the Internet or other reference sources, identify three to five species conforming to each of the three mortality schedule types.
  3. Over long periods of time, ecological processes ensure that most populations either stay the same size or go extinct, since the finiteness of our world does not permit them to grow without bound. In the former case, we expect in the long run (i.e., on average over time) that R0 = 1, which implies:

    images

    If a species has the mortality schedule given in part 1, and a natality schedule b(x) of the form

    images

    where m < xmax (i.e., individuals start reproducing at age m beyond which reproduction is the age-independent rate of b progeny per time period), then for xmax = 100 explore the trade-offs in the values of r, b, and m that correspond to a stable population (i.e., R0 = 1) and provide an expression for the corresponding generation time. (Hint: Use the condition R0 = 1 to get a relationship among r, b, and m; then express one of the parameters in terms of the other two. For selected values of one of the parameters, you can then graph relationships between the other two. What general statements can you make about these relationships?)

  4. The mortality schedule

    images

    is of Type III on [0, xmax], provided d > xmax, because mortality rates are relatively low until individuals approach age d, around which mortality rates increase strongly. Repeat the previous exercise with this mortality schedule instead of the Type II schedule; look at trade-offs in the values of d, b, and m.

* K. E. Godfrey and L. W. J. Anderson, “Developmental Rates of Bagous affinis at Constant Temperatures,” Florida Entomologist 77 (1994); 516–519.

* W. O. Kermack and A. G. McKendrick, “A Contribution to the Mathematical Theory of Epidemics,” Proceedings of the Royal Statistical Society 115 (1927): 700–721.

* Jon Jorgenson et al., “Effects of Population Density on Horn Development in Bighorn Rams,” Journal of Wildlife Management 62 (1998): 1011–1020.

* Data from The Engines of Our Ingenuity public radio program produced by KUHF-FM Houston, Episode No. 426.

* Problem 26 and 27 were found at http://illuminations.nctm.org/LessonDetail.aspx?ID=L461 on February 25, 2013.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset