Solution Manual For Intermediate Microeconomics and Its Application 13e Walter Nicholson, Christophe

Page 1


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Solution Manual For Intermediate Microeconomics and Its Application 13e Walter Nicholson, Christopher Snyder Chapter 1-18

Chapter 1: Two Basic Economic Models Purpose and Organization of the Chapter This chapter provides an introduction to the book by showing why economists use simplified models. The chapter begins with a few definitions of economics and then turns to a discussion of such models. Development of Marshall's analysis of supply and demand is the main example used here, and this provides a review for students of what they learned in introductory economics. The notion of how shifts in supply or demand curves affect equilibrium prices is highlighted. The chapter also reminds students of the production possibility frontier concept and shows how it illustrates opportunity costs. The chapter concludes with a discussion of how economic models might be verified. A brief description of the distinction between positive and normative analysis is also presented.

Lecture and Discussion Suggestions We have found that a useful way to start the course is with one (or perhaps two) lectures on the historical development of microeconomics together with some current examples. For example, many students find economic applications to the natural world fascinating and some of the economics behind Application 1.1, might be examined. Application 1.6: Economic Confusion provides normative distinction and to tell a few economic jokes (if your supply of such jokes is running low – see Additional Resources). In terms of explicit content, some time should be spent on reminding students about how supply and demand curves work since these concepts underlie most of microeconomics. Especially important is to make sure that students understand that these curves show firms’ and consumers’ reactions to all possible prices. That is, the independent variable is on the vertical axis. Far more on the problems raised by this approach is provided in Chapter 2.

Cengage Supplements The following product-level supplements provide additional information that may help you in preparing your course. They are available in the Instructor Resource Center.  

Test Bank PowerPoint slides

Chapter Objectives The following objectives are addressed in this chapter:

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

1


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

01.01 Develop and analyze two basic economic models: 

The Production Possibility Frontier (PPF)

The Supply-Demand Model

01.02 Explain how to use the PPF to break down six basic economic principles. 01.03 Understand how you can apply microeconomics to analyze all types of problems. 01.04 Explain how the interaction of buyers and sellers determines a good’s price. 01.05 Explain different ways in which economists verify theoretical models.

[return to top]

What's New in This Chapter The following elements are improvements in this chapter from the previous edition: 

Chapter 1 is now a standalone chapter. The mathematical material from what was previously an appendix to the chapter, now forms the basis for Chapter 2.

An extended Application 1.3 examines video streaming and cord-cutting, suggesting the importance of dynamism in the economy.

A new Application 1.4 shows how a simple supply and demand model can explain pricing of eggs during the COVID-19 lockdowns.

A revised set of Review Questions focus more explicitly on the basic two models introduced in the chapter.

[return to top]

Additional Resources

The most accessible introduction to the history of economics remains Robert Heilbroner’s The Worldly Philosophers (Seventh Edition) Touchstone, 1999. This one-minute video does a nice job of introducing most elements of supply and demand analysis in a cartoon format: https://www.youtube.com/watch?v=720uyg0Dd_M There are many websites featuring economic jokes. A good one is https://upjoke.com/economist-jokes. [return to top]

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

2


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Solutions to End of Chapter Problems Chapter 1 has no end of chapter problems. Problems involving both the production possibility frontier and simple supply and demand curves can be found at the end of Chapter 2.

Chapter 2: Some Useful Math Purpose and Organization of the Chapter Chapter 2 is new to this edition. It draws together the material that was previously in the appendix to chapter 1 and adds a considerable amount of new material to fill out a complete chapter. Many of the concepts here are drawn from a course in algebra. These include concepts of the slope and intercept of a linear graph and some details on the solving of simultaneous equations. Topics that are specifically oriented toward the use of algebra in microeconomics include the importance of defining units for specific economic relationships and how functions with two independent variables can be represented by their contour lines. The chapter concludes with a very brief introduction to some of the statistical problems encountered in estimating microeconomic models using real world data.

Lecture and Discussion Suggestions Lecturing on this material is a good way to turn off most students. Hence, we believe the chapter should be used primarily as a reference, urging students with poorer math preparation to use it as needed. It is likely, however, that all students could benefit from a reminder about how units of measurement affect linear equations and some of the material on contour lines (since these will be encountered in the next chapter. Whether mathematics should play a prominent role in microeconomics is a good topic for discussion. As shown by this text (and by our more advanced one) we are firmly in the camp of stressing the value of mathematics to the subject. But there are a variety of contrary views that might be brought up. Whether economics should be viewed as a ―science‖ or as philosophy is a good place to start. Two videos on the topic are listed below.

Cengage Supplements The following product-level supplements provide additional information that may help you in preparing your course. They are available in the Instructor Resource Center.  

Test Bank PowerPoint Slides

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

3


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Chapter Objectives The following objectives are addressed in this chapter: 02.01 Write and graph an equation of a function of one variable. 02.02

Write and graph an equation of a function of two or more variables.

02.03

Solve systems of simultaneous equations.

02.04

Understand the basics of testing economic models with real world data.

[return to top]

What's New in This Chapter The most important new material in this chapter includes the following. 

An extended discussion of ―Marshall’s Trap‖ mitigates the confusion caused by Alfred Marshall’s decision to put an independent variable (price) on the vertical axis in graphs.

An update of the text’s simple model of the world oil market incorporates the effects of COVID-19 (Application 2.3). Somewhat surprisingly, the model continues to perform fairly well.

A new Application 2.4 introduces the ―identification problem‖ as it relates to using actual data to derive supply and demand curves. The proposed solution of Working in 1927 remains the key insight.

An entire series of new review questions stresses a variety of algebraic topics.

[return to top]

Additional Resources Videos from Kahn Academy provide extensive review material on all topics related to algebra and the representation of functions with graphs. Students with poor math preparation should be directed to this excellent source. The Foundation for Economic Education has a nice essay on the possible overuse of mathematics in economics : https://fee.org/articles/the-overuse-of-mathematics-ineconomics/. Several years ago, Dani Rodrik’s website had a very nice discussion of why math is important to economics: https://rodrik.typepad.com/dani_rodriks_weblog/2007/09/why-weuse-math.html. [return to top]

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

4


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Solutions to End of Chapter Problems Students have access to solutions for the odd-numbered problems as well as video problem walkthroughs for problems 3 and 7.

2.1

a.

b.

Yes, the points seem to be on straight lines. For the demand curve: P = 1 Q = –100

Q 100 at P  1, Q  700, so a  8 and Q P 8 or Q  800  100 P 100 Pa

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

5


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

For the supply curve, the points also seem to be on a straight line: P 1  Q 200 If P  a  bQ  a 

Q 200

at P = 2, Q = 300, 2 = – a + 1.5, or a = 0.5. Hence the equation is P  0.5  c,d

Q or Q  200 P  100 200

For supply Q = 200P – 100 If P = 0, Q = –100 = 0 (since negative supply is impossible). If P = 6, Q = 1100. For demand Q=800 ‒ 100P When P = 0, Q = 800. When P = 6, Q = 200. Excess Demand at P = 0 is 800. Excess supply at P = 6 is 1,100 – 200 = 900.

2.2

a.

Supply: Q = 200P – 100 Demand: Q = –100P + 800 Supply = Demand: 200P – 100 = –100P + 800 300P = 900 or P = 3 When P = 3, Q = 500.

b.

At P = 2, Demand = 600 and Supply = 300. At P = 4, Demand = 400 and Supply = 700.

c.

d.

New demand is Q = –100P + 1100.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

6


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

e.

Supply = Demand: 200P – 100 = –100P + 1100. 300 P = 1200 P = 4, Q = 700.

f.

Supply is now Q = 200P – 400.

g.

Supply = Demand when QS  QD 200P – 400 = –100P + 800 300P = 1200 P = 4, Q = 400

h.

At P = 3, QS  200, QD  500 ; this is not an equilibrium price. Participants would know this is not an equilibrium price because there would be a shortage of orange juice.

i.

2.3

a. Excess Demand is the following at the various prices

P  1 ED  700  100  600 P  2 ED  600  300  300 P  3 ED  500  500  0 P  4 ED  400  700  300 P  5 ED  300  900  600 The auctioneer found the equilibrium price where ED = 0. b. Here is the information the auctioneer gathers from calling quantities:

Q  300 PS  2 PD  5 Q  500 PS  3 PD  3 Q  700 PS  4 PD  1 So, the auctioneer knows that Q = 500 is an equilibrium.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

7


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

c. Many callout auctions operate this way – though usually quantity supplied is a fixed amount. Many financial markets operate with ―bid‖ and ―asked‖ prices which approximate the procedure in part b. 2.4

The complaint is essentially correct – in many economic models price is the independent variable and quantity is the dependent variable. Marshall originally chose this approach because he found it easier to draw cost curves (an essential element of supply theory) with quantity on the horizontal axis. In that case, quantity can legitimately be treated as the independent variable. a. The restrictions on P are necessary with linear functions to ensure that quantities do not turn negative. b. The following graph has P on the vertical axis. Equilibrium P is found by  P  10  P  2  2 P  12  P  6, Q  4 .

c. The following figure graphs the demand and supply curves with P on the horizontal axis. Solution proceeds as in Part b.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

8


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

d.

Graphed either way, the equations yield the same intersection point.

f.

Both graphs yield the same solution

e.

Reasons for preferring one over the other are not readily apparent in these drawings. As we shall see, however, developing demand and supply curves from their underlying theoretical foundations does provide some rationale for Marshall’s choice.

2.5

The algebraic solution proceeds as follows: a.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

9


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

QD  2 P  20 QS  2 P  4 QD  QS  4 P  24 P  6, QD  QS  8 b.

QD  2 P  24 QD  QS  4 P  28 P  7, QD  QS  10.

2.6

c.

P = 8, Q = 8 (see graph)

a.

T = .01 I

2 2

I = 10, T = 0.01(10) = 1 2

I = 30, T = 0.01(30) = 9

Taxes = $1,000 Taxes = $9,000

2

I = 50, T = 0.01(50) = 25 Taxes = $25,000 I = 100, T = 100. b.

Average Rate

Marginal Rate

I = 10,000

10%

20%

I = 30,000

30%

60%

I = 50,000

50%

100%

I

T

Marginal Tax Rate

10,000

1,000

c.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

10


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

2.7

2.8

10,001

1,000.20

30,000

9,000

30,001

9,000.60

50,000

25,000

50,001

25,001

0.20 0.60 1.00

a.

b.

Both these points lie below the frontier.

c.

This point lies beyond the frontier.

d.

Opportunity cost of 1Y is 2X independent of production levels.

a.

If Y = 0, X = 10 If X = 0, Y = 5

b.

X

Y

2

24

4

21

6

4

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

11


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

X2 Y2   1 is a quarter of an ellipse since both X and Y are positive. 100 25 c.

The opportunity cost depends on the levels of output because the slope of a frontier is not constant.

d.

The opportunity cost of X is the change in Y when one more unit of X is produced. Example: X0 = 3, X1 = 4 When X0 = 3, Y0 =

When X1 = 4, Y1 =

21

[Y1 – Y0] = 0.187. 0.187 units of Y are "given up" to produce one more unit of X at X = 3. 2.9

a.

2

2

X + 4Y = 100 2

If X = Y, then 5X = 100 and X = 20 and Y =

20 .

b.

X = 10, can consume where any X, Y combination such that X + Y = 10.

c.

Since prefers X = Y, will choose X = Y = 5.

d.

The cost of forgone trade is 5 –

20 = 5 – 2

5 = 1.52 units of both X and Y.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

12


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

2.10

This problem provides practice with contour lines. a.

If Y  X  Z the Y = 4 is the same line as ―Y = 2‖ in Figure 1A.5.

b.

If X  8  4Z , Y  X  Z  8Z  4Z 2  4. This has a solution of Z = 1, X = 4.

c.

None of the other points on the Y = 4 contour line obey the linear equation. This is so because the contour line is convex and hits the straight line at only a single tangency.

d.

If X  10  4Z , Y  10Z  4Z 2  4 or 4Z 2  10 Z  4  0 . Using the quadratic formula yields Z  (10  100  64) / 8 or Z  2, 0.5 . Hence the line intersects the contour in two places. These points of intersection are Z = 2, X = 2, and Z = 0.5, X = 8.

e.

Yes, many points on the line X  4Z  10 provide a higher value for Y (any points between the two identified in part d do). The largest value for Y is at the point X = 5, Z = 5/4. In this case Y = 25/4 = 6.25.

f.

As we shall see in Chapter 3, this problem is formally equivalent to utility maximization in which utility is given by U ( X , Z )  X  Z , the price of good X is 1, the price of good Z is 4, and income is either 8 or 10.

Chapter 3: Utility and Choice Purpose and Organization of the Chapter Since this chapter introduces the student to many new concepts, it is one of the more difficult chapters in the text. The central concept of the chapter is the indifference curve and its slope, the Marginal Rate of Substitution (MRS). The MRS formalizes the notion of trade-off and is (in principle) measurable. For those reasons it is superior to an approach to consumer theory involving ―marginal utility.‖ The definition provided for the MRS in this chapter needs to be approached carefully. Here the concept is defined as the Marginal Rate of Substitution of ―X for Y,‖ by which is meant X is being substituted for Y. In graphic terms the individual is moving counterclockwise along an indifference curve and the MRS measures how much Y will be willingly given up if one more X becomes available. The pedagogic convention of always using counterclockwise movements along an indifference curve is helpful because the MRS does indeed diminish for movements in that direction. Students’ primary difficulty with the material in this chapter tends to be confusing the MRS (a slope concept) with the ratio of the amounts of two goods. Unfortunately, that confusion is

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

13


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

increased by some examples based on the Cobb-Douglas utility function, which make it appear that the two concepts are interchangeable. To avoid this confusion, some instructors may wish to give further emphasis to the marginal utility definition of MRS, which is presented in footnote 2 of the chapter. This might be followed by greater use of the utility maximization principle (the ―equi-marginal principle‖) from footnote 5. The soft drink-hamburger example that runs throughout the chapter is intended to provide an easy, mildly amusing introduction to the subject for students. In general, the example seems to work well and is, we believe, definitely superior to introducing the concepts through general goods X and Y. Note also that this chapter includes analyses of four specific kinds of goods (useless goods, economic bads, perfect substitutes, and perfect complements). Examining the utility maximizing conditions in these cases (Figures 3.5 and 3.9) should help students to visualize what the conditions mean in cases where the results should be obvious.

Lecture and Discussion Suggestions The challenge in lecturing on this chapter is to avoid mere repetition of the text. One way to do that is to offer a somewhat more mathematical treatment. The use of calculus involved in such a treatment may, however, prove too difficult for students to grasp, especially if it involves introducing the Lagrangian technique. An alternative approach would be to start from one point in the X-Y plane and ask how an indifference curve might look. Proceeding from one point to the next in this way reinforces the concept of the trade-off and (on a more sophisticated level) demonstrates Samuelson’s integrability problems. Once a single indifference curve has been traced out, a second can be constructed to the northeast of the first by using the ―more is better‖ assumption and proceeding with an identical construction. Utility maximization can be approached in the same way by starting at the Y-intercept on the budget constraint and inquiring whether the individual would make various trades along the constraint. Discussions of the chapter material might focus on real world illustrations of both economic and non-economic choices that people make. To approach these, students might be asked to theorize what budget constraint faces people in unusual situations (e.g., what is the cost of shopping for bargains or for wearing seat belts). The instructor can then ask whether there is evidence that individuals respond to changes in the relative costs associated with such activities (that is, do they search more for bargains in high priced items, or are certain types of people less likely to wear seat belts). Application 3.6 Loyalty Programs also offers a number of discussion possibilities that would help to illustrate the actual shape of budget constraints.

Cengage Supplements The following product-level supplements provide additional information that may help you in preparing your course. They are available in the Instructor Resource Center.  

Test Bank PowerPoint Slides

Chapter Objectives The following objectives are addressed in this chapter: 03.01 Develop a theory of choice. How do people make choices or decision?

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

14


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

03.02

Explain how to represent a consumer’s preferences.

03.03

Understand how income and prices constrain a consumer’s choices.

03.04

Determine how consumers maximize utility given their budget constraint and preferences.

[return to top]

What's New in This Chapter There are relatively few new elements to this chapter. The most important changes from the previous edition are the following. 

More material has been included on neuroscience and economics including a discussion of how findings in this area are used to develop consumer regulations.

An increased set of numerical examples shows how utility-maximization problems are solved.

An explicit color convention is maintained for graphs with indifference curve maps being shown in red and budget constraints in blue. In Chapter 4, the red color continues to be used for demand curves (which are related since they are derived from indifference curve maps).

[return to top]

Additional Resources The text focuses on the Marginal Rate of Substitution in its discussion of utility maximization. The notion of ―equal marginal utility per dollar spent may be more intuitive for some students. The Kahn Academy video provides an excellent discussion of this approach: https://www.khanacademy.org/economics-finance-domain/apmicroeconomics/basic-economic-concepts/16/v/equalizing-marginal-utility-perdollar-spent. The famous paper by Gary Becker and George Stigler ―De Gustibus Non Est Disputandum‖ (American Economic Review, March 1977) contains a wide-ranging and fascinating discussion of the origins of ―tastes‖ and the relationship of this to utility theory. The Wikipedia entry on the court case of Hamer v. Sidway, https://en.wikipedia.org/wiki/Hamer_v._Sidway , offers a great deal more historical information about this case from Application 3.5.

[return to top]

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

15


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Solutions to End of Chapter Problems Students have access to solutions for the odd-numbered problems as well as video problem walkthroughs for problems 3 and 7.

3.1

a.

$8.00 = 20 apples can be bought. $0.40/apple

b.

$8.00 = 80 bananas can be bought. $0.10/banana

c.

10 apples cost: 10 apples × $0.40/apple = $4.00, so there is $8.00 – $4.00 = $4.00 left to spend on bananas which means

$4.00 = 40 bananas can be bought. $0.10/banana d.

One less apple frees $0.40 to be spent on bananas, so

$0.40 = 4 more bananas can be bought. $0.10/banana

3.2

e.

$8.00 = $0.40  number of apples + $0.10  number of bananas = 0.40A + 0.10B.

a.

U=

b.

U = 20 = 10B so 400 = 10 B, implying

A B =

5  80 =

400 = 20.

40 = B. c.

U = 20 =

20 B , so 400 = 20 B, implying 20 = B.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

16


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

d.

e.

From the budget part d, an individual can buy 10 apples and 40 oranges.

f.

One less apple: U =

9  44 =

396 <

One more apple: U = 11  36 =

400 = 20.

396 <

400 .

At both endpoints of the budget constraint: U = 0 =

20  0 =

0  80 .

Graph shown in d. 3.3

To graph the indifference curves, use U 2 instead of U. U = 10 means U 2  100  C  D . Hence, indifference curves are hyperbolas.

3.4

a.

See Figure in Solutions to odd-numbered problems.

b.

See graph.

c.

D = 10, U  10  0  0

d.

If, say, spent half of income on D, half on C, would buy D = 5, C = 20. Utility would be U  5  20  10 which is less than 20. Trial and error shows that any other budgetary allocation provides even less utility than this.

e.

As in part d, Paul can buy 20 C and utility will be 10.

f.

Any other allocation yields less utility (see graph).

a.

Tangency is the same in either case.

b.

Costs are: i. $520 ii. $290 iii. $205 iv. $200 v. $250 vi. $425

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

17


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

c.

The bundle C = 20, D = 5 (option iv in part b) is the least costly of those that provide utility of 10. This is the same solution as in problem 3.3.

3.5

a. The indifference curves here are straight lines with slope –4/3. Hence, the MRS is a constant 4/3. The goods are perfect substitutes. b. Because one unit of tea provides more utility than a unit of coffee, she will spend all of her income on tea when the prices are equal: T = 4, C = 0. c. The graph shows that the indifference curves (which have a constant slope) are always steeper than the budget constraint, so maximum occurs on the T axis. d. With more income she would continue to buy only tea. If coffee prices fall to $2, coffee is now a cheaper way to obtain utility: one unit of coffee yields 3 units of utility at a cost of $2 so utility costs $2/3 per unit of utility. With tea, utility costs $3/4 per unit of utility. 3.6

a.

Each meal consists of PB = 2, C = 1. This costs 4×2 + 2×1 = 10. With an income of $100 she can buy 10 meals per month, or PB = 20, C = 10.

b.

Now each meal costs 5×2 + 2×1 = 12. She can buy 100/12 = 8.33 meals.

c.

To restore Vera’s ability to buy 10 meals she would need Food Stamps to buy 1.67 meals. These would cost 1.67 × 12 = 20.

d.

These preferences allow no substitution of PB for C in response to changing prices. A graph of this utility function would resemble that shown for Right Shoes and Left Shoes in Figure 3.5d.

3.7

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

18


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Income subsidy is cheaper since AB < A´B´. This result occurs because the housing subsidy encourages people to buy more housing though housing is not really cheaper. 3.8

This person will participate in the Food Stamp program if (as in graph) they can reach a utility level higher than U 0 by doing so. With cash, the post-transfer constraint would extend the line to the nonfood axis, making it desirable for all to participate. 3.9

a, b.

The figure shows that an unconstrained choice will yield utility level U1 with choices of C*, H*. If the government requires purchase of H**, utility would fall to U0. Low-income consumers are most likely to be constrained by H  H**. c.

To restore this person to U1 would require extra income to shift the budget constraint outward to I′.

d.

A housing subsidy would permit this person to reach U1 with budget constraint I′′.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

19


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

3.10

a.

In problems 2.2 and 2.3     0.5 .

b.

Utility maximization requires PX PY  MRS  Y  X  Y (1   ) X . Some algebraic manipulation yields: (1   ) PX X   PY Y . Substituting this into the budget constraint yields: PX X  (1   ) PX X   I or PX X   I .

c.

Because this person spends  I on good X, this amount does not change unless I changes.

d.

Because spending on X is given by  I , changes in the price of Y will not affect this spending.

e.

If income doubles, spending on both X and Y must double because income is split evenly between the two goods. But prices have not changed, so the quantities of X and Y must double.

Chapter 4: Demand Curves Purpose and Organization of the Chapter This chapter provides a complete development of the demand curve concept. It begins with the traditional analysis of the effects of changes in income and prices on the quantities of goods one person demands. Most of the analysis deals with reactions to price changes: income and substitution effects are stressed. Considerable emphasis is given to investigating reasons why these individual demand curves might shift. The purpose of such a detailed investigation is to plant firmly in students’ minds the distinction between movement along a demand curve and shifts in a demand curve. Only by understanding the way in which demand curves are constructed and the ceteris paribus assumptions that are implied is it possible to grasp this distinction completely. Consumer surplus is shown using the usual (Marshallian) demand curve rather than with a compensated demand curve. The compensated demand curve notion is mentioned only in the problems. Market demand curves are developed in the second half of this chapter by summing the individual curves. This construction demonstrates the notion of price-taking behavior that lies behind such demand curves. The summing technique is also used to demonstrate how shifts in market curves are brought about by shifts in individuals’ curves. Finally, the chapter introduces the general concept of elasticity and shows its application to demand theory. Only point elasticity is mentioned. This raises some problems in providing a precise definition without using calculus, but the approach seems preferable to introducing all the algebra that arc elasticity requires. An extended section seeks to clarify the relationship between slope and elasticity for a linear demand curve. Throughout the discussion linear demand is represented by the expression Q = a + bP (where b<0). Some authors prefer to write this expression as Q = a – bP (where b > 0), but we believe this leads to considerable confusion, especially when deriving elasticities. We believe students also should learn about log-linear (constant elasticity) demand curves too, but these are covered only in a footnote (footnote 8). The relationship between total expenditures and price elasticity is analyzed in the chapter, but the concept of marginal revenue is not explicitly introduced until Chapter 9. The reason for this

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

20


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

is that marginal revenue is not relevant to demanders—it is a concept that should be encountered in connection with the discussion of firms’ goals.

Lecture and Discussion Suggestions Comparative statics methodology should be the principal focus of the lecture for this chapter. It is essential that students understand why one compares equilibrium (utility-maximizing) positions to analyze behavior. In addition to repeating the discussion of income and substitution effects (including perhaps the Slutsky equation) there are two other approaches to this chapter that might also make the point. First, one could introduce the revealed preference concept (through a two-good graphical approach) and show that the axioms of rationality require that the substitution effect be negative. This proof would take about one class and would be a useful supplement to material in the text. A second approach to teaching comparative statics would be to offer an extended example in lecture. Going over the Lump Sum Principle (Figure 4.6) should reinforce the distinction between income and substitution effects. Policy applications of the elasticity concept provide the most interesting points of departure for discussions based on this chapter. The health-insurance example (Application 4.7) raises a number of issues about whether elasticities can provide a guide for policy actions (whether services with high demand elasticities should be covered by insurance is an important issue for all health-reform plans, for example). The housing and electricity estimates offer the opportunity to develop a similar set of questions. For students who have had a fairly broad exposure to economics, the elasticity estimates might also be used as a way to introduce notions of ―optimal‖ (non-distorting) taxes and subsidies—that is, one might ask how tax- and subsidy-induced price effects might be minimized and whether this would make sense from an overall perspective.

Cengage Supplements The following product-level supplements provide additional information that may help you in preparing your course. They are available in the Instructor Resource Center.  

Test Bank PowerPoint Slides

Chapter Objectives The following objectives are addressed in this chapter: 04.01

Derive an individual’s demand curve.

04.02

Show how changes in income, price, and prices of related goods affect a consumer’s optimal choice.

04.02

Understand the difference between the substitution effect and income effect.

04.03

Explain what factors cause an individual’s demand curve to shift inward (left) or outward (right).

04.04

Explain how price changes affect people’s welfare or consumer surplus.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

21


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

04.05 04.06

Derive a market demand curve from individual demand curves. Measure how changes in one variable affect some other variable: by how much does quantity demanded change when price changes?

[return to top]

What's New in This Chapter Most of the new material in this chapter is in the applications. Some examples include the following. 

Consumption data has been updated in the Application 4.1 Engel’s Law. The new data show that savings are highly income elastic. This could be the basis for discussions of individual’s retirement needs.

The application on valuing new goods has been expanded to include an analysis of how the arrival of new goods affects other markets. The welfare analysis of such effects is complex (Application 4.4).

The discussion of brand loyalty has been expanded to discuss some of the research on the effects of social media (Application 4.5)

The other major addition to the chapter is a greater focus on linear demand functions of the form Q = a + bP (where b < 0). The goals here are to stress that, for the individual, price is the independent variable. Once again Marshall’s choice of putting this variable on the vertical axis is discussed and some intuitive material about how to think about what ―slope‖ means in tis context is provided. This linear form is also used to motivate a formal definition of price elasticity.

[return to top]

Additional Resources How the traditional model of choice relates to addictive type behavior has been a focus of economic research for many years. Perhaps the most famous article is Gary Becker and Kevin Murphy ―A Theory of Rational addiction‖ (Journal of Political Economy, August 1988). A more up-to-date application is provided by Rachel Griffith’s work on whether taxes on sugary drinks affect consumption. Her work is summarized in the video https://voxeu.org/content/dosoda-taxes-work. Wikipedia has an extensive discussion of brand loyalty, mainly from a marketing perspective. Students might be asked to read this and then explain how all of these insights relate to the simple economic model of choice. See https://en.wikipedia.org/wiki/Brand_loyalty. An example of economic education in popular culture is from the HBO series The Wire. The drug dealer Stringer Bell (Idris Elba) attends an economics class (misleadingly called

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

22


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

―macroeconomics‖) and learns about elasticity. He then conveys what he has learned to his dealer team. One location for the clip is https://www.youtube.com/watch?v=COf2bQEQ7Zw.

[return to top]

Solutions to End of Chapter Problems Students have access to solutions for the odd-numbered problems as well as video problem walkthroughs for problems 3 and 10. 4.1

a.

I = $200, S = J. Ps S + PJ J = 20, implying 20S + 20S = 200, implying 40S = 200. S = 5, J = 5.

b.

PS.S + PJ.J = I, implying 20S + 30S = 200, implying 50S = 200. S = 4, J = 4

c.

Elizabeth’s indifference curves are L-shaped since she gains utility only when shoes and jeans are purchased in a one-to-one proportion. 10 shoes and 5 pairs of jeans yield the same utility as 5 sweaters and 5 pairs of jeans. d.

The change from U 2 to U1 is entirely attributable to the income effect. There is no substitution effect due to Elizabeth’s insistence on a fixed proportion of jeans and shoes.

e.

S = J throughout because of her preferences. 20S + PJ S = 200 SJ 

200 20  PJ

The following choices will be made: PJ S=J 30 20

4 5

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

23


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

62/3 8

10 5 f.

g.

Now S = J =

300 20  PJ

PJ

S=J

30 20 10 5

6 7.5 10 12

More J is demanded at each price (see graph in part f). h.

Now: S  J 

200 30  PJ

This will shift both demand curves inward.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

24


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

4.2

4.3

a.

b.

These price changes still allow Paula to afford her initial choices. Hence the budget constraint rotates around this point (T  5, L  4) .

c.

Because the new budget constraint is no longer tangent to the indifference curve, Paula can make choices that improve utility. The figure shows why this choice will involve more L and less T.

d.

This effect allows utility to increase whereas we have defined the substitution effect as constituting a move along a single indifference curve.

e.

If the substitution effect is defined as the result of a rotation around the initial consumption bundle, the ―income effect‖ would be measured by the effects of parallel shifts in the budget constraint from this point. The end result would be the same under either disaggregation.

a.

PB = 2J and 0.05PB + 0.1J = 3 5PB + 10J = 300 PB + 2J = 60 4J = 60

b.

c.

J = 15,

PB = 30

PJ  $0.15

PB = 2J

0.05PB + 0.15J = 3

5PB + 15J = 300

25J = 300

J = 12, PB = 24

To continue buying J = 15, PB = 30, David would need to buy 3 more ounces of jelly and 6 more ounces of peanut butter. Should increase his allowance by 3(0.15) + 6(0.05) = $.75.

d.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

25


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

e.

Since David uses only PB and J to make sandwiches (in fixed proportions), and because bread is free, it is just as though he buys sandwiches where Psandwich  2 PPB  PJ . In part a, PS = .20, QS = 15. In part b, PS = 0.25, QS = 12. In general, QS  3 PS

f.

There is no substitution effect due to the fixed proportion nature of David’s preferences. A change in price results only in an income effect.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

26


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

4.4

a. b.

c.

See graph.

d.

Since Q = 0 for P > 10, CS = 0. Equation gives same result.

e.

If P = 3, Q = 20 and CS = 180 – 60lnP = 180 – 60 = 120. This is the amount that Irene would pay for the right to buy pizza at a price of 3. With P = 4 Q = 15 and CS = 174 – 60ln(4) = 174 – 83 = 91. She would be willing to pay 39 less for the right to buy pizza at 4.

f.

4.5

4.6

Function is homogeneous because a doubling of I and P leaves Q unchanged 60 Graph of Q  P

a.

This is true by definition. The person starts from the same place under either concept.

b.

The compensated demand curve incorporates only substitution effects. Because Marshallian demand also incorporates income effects, demand will generally be more price-responsive under the Marshall concept.

c.

Because utility varies along the Marshallian demand curve, each point provides a new utility level from which to construct a different compensated demand curve.

d.

There are no substitution effects in this case so the compensated demand curve will be vertical. The Marshallian demand curve will be sloped, however, because of income effects.

a.

U2020 =

40  40 = 40

U2021 =

20  80 = 40

b.

(

)

(

)

(

)

(

)

.

―Real‖ income has risen.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

27


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

c.

( (

) )

( (

) )

( (

) )

.

―Real‖ income has fallen.

4.7

d.

Results of calculations depend on which prices are used. It may be necessary to use some combination of the two indices to conclude (correctly) that utility (real income) has not changed. Notice that the product of the real income ratios does 100 160   1.0 . give the correct solution (though this is a special case). 80 200

a.

Q = 20

b.

Q = 0 when P = 20

c.

P=1

Q = 19 P  Q = 19

P=2

Q = 18 P  Q = 36

P=3

Q = 17 P  Q = 51

P=9

Q = 11 P  Q = 99

P = 10

Q = 10 P  Q = 100

P = 11

Q=9

P  Q = 99

P = 19

Q=1

P  Q = 19

d.

Highest total expenditures are 100 when P = 10.

e.

Since 40 – 2P = 2(20 – P), Q will be twice as large at each price. Total expenditures are still as large as possible when P = 10.

4.8

a.

Yakko Wakko

Dot

Total

P = 50

0

0

0

0

= 35

30

20

0

50

= 25

50

60

25

135

= 10

80

120

100

300

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

28


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

= 0 b.

100

160

150

410

―Total‖ column in part a.

c.

4.9

d.

Above graph.

a.

Because the market demand curve is the horizontal sum of each individual’s demand curve, the total area of consumer surplus triangles for each person will equal the area of the consumer surplus triangle in the market. This is easiest to show for a small price increase of P . Let initial quantities be denoted by asterisks, post-change quantities by primes. Then, for each person the loss of consumer surplus is Pxi'  0.5P( xi*  xii ) . Summing over all individuals yields

PX '  0.5P( X *  X ' ) which is what one would get from the market demand curve. b.

The loss of consumer surplus is larger in the inelastic case because consumers do not reduce quantity purchased by very much in response to the price increase. With small substitution effects consumers cannot ―get out of the way‖ of the price increase whereas with larger ones they can.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

29


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

4.10

c.

This would be literally true only if demand were completely inelastic. With a more elastic demand total spending total spending may even fall in response to a price increase though there will still be a loss of consumer surplus.

a.

P

a when Q = 0. b

b.

Y

a a Q * a Q *  P*    . b b b b

c.

Q P P* because the demand curve is linear with slope b (note this is   b P Q Q* a negative number). eQ , P 

d. Now use the result that X  P* , Y 

Q* b

eQ , P 

P* X  * Q Y b

e.

For movements downward along a linear demand curve, distance X falls and distance Y increases. Hence the demand curve becomes less elastic.

f.

One could draw a linear demand curve tangent to any demand curve at a specific point. This tangent demand curve would have the same elasticity as the original one at this point.

Chapter 5: Uncertainty Purpose and Organization of the Chapter Chapter 5 provides a foundation for students to help understand the important role that uncertainty and information theory has come to play in microeconomics. This material appears early in the text for two reasons: first, so that it can be used occasionally in subsequent chapters (including for example the chapter on game theory, where uncertainty comes up in the discussion of mixed strategies), and second, because it can be viewed as an extension of the basic theory of consumer choice to environments involving uncertainty, it should naturally appear right after the chapters on utility and choice. The basic goal of the chapter is to show why individuals are generally risk-averse and are therefore willing to pay something to reduce the risks they face. That point is made initially using the Friedman-Savage utility of wealth analysis and followed up with a discussing of various methods for reducing risk and uncertainty including insurance, diversification, options, and information. Financial applications are highlighted in a separate section. The appendix introduces a new model—a two-state model based on Rothschild-Stiglitz—which we show can be applied to understand all the previous concepts. Making use of the ―certainty line‖ in this model of consumer choice is an especially intuitive way of illustrating risk aversion. Most of the material on asymmetric information is a bit more advanced and so is provided in a later chapter (Chapter 16).

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

30


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Lecture and Discussion Suggestions With the huge volume of material that needs to be covered in an intermediate microeconomics course, the instructor needs to pare down what is covered to fit into a sensible one-semester course. There is a great temptation to omit this chapter, but we would urge the instructor to reconsider this choice. Uncertainty is an extremely important topic and may be covered in no other course that the undergraduate takes. If time constraints are severe, the instructor could cover the very basics, say on risk aversion and insurance (Section 5.2 and the first entry in Section 5.3). With more time, the instructor could include coverage of diversification, options, and information. Of course, a second semester course covering advanced topics in microeconomics could spend a lot more time on uncertainty and could have a whole unit on uncertainty, including a deeper treatment of Chapter 5 combined with the material on asymmetric information in Chapter 16. Courses in business schools may want to include coverage of Section 5.4 on financial applications. The appendix model essentially goes through all the same material a second time using a different model. Many instructors will choose to omit the appendix entirely. Another possibility is just to introduce the bare essentials (say just the text surrounding Figures 5A.1 to 5A.3) to supplement the earlier material. A third option, for instructors who particularly favor the twostate model, is to focus their lectures entirely around the appendix and have the students read the earlier material as background. The concepts in this chapter are perhaps a bit more difficult than in some other chapters. More repetition of material in the book may be re-quired in lectures to help guide students to an understanding (for example, Figure 5.3 is very complex, as are some of the graphs for the twostate model in the appendix), or certain issues can simply be omitted. Discussion topics on uncertainty and information are virtually unlimited. Issues about the stock market (efficient market theory, the role of in-vestment advisors, and so forth), the reform of insurance for healthcare in the U.S., and crime and punishment (with imperfect and so uncertain en-forcement) are all of great interest to students.

Cengage Supplements The following product-level supplements provide additional information that may help you in preparing your course. They are available in the Instructor Resource Center.  

Test Bank PowerPoint Slides

Chapter Objectives The following objectives are addressed in this chapter: 05.01

Calculate the expected value of an uncertain outcome.

05.02

Understand how risk aversion influences the choices people make.

05.03

Explain how risk averse people can purchase insurance to avoid risky outcomes.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

31


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

05.04

Understand different methods for reducing risk and uncertainty.

05.05

Understand the choices made by risk averse investors.

[return to top]

What's New in This Chapter The chapter has remained substantially the same as in the previous edition. The applications to financial markets have been updated to reflect current data. 

Application 4.4: Puts, Calls, and Black-Scholes continues to use the example of Microsoft stock, but the stock price has been increased from around $30 to around $230, closer to the market price as of this writing. Problem 5.8 has been revised to reflect higher stock prices.

Application 4.6: The Equity Premium Puzzle now incorporates market data through 2019, the most current available from the referenced data source.

[return to top]

Additional Resources The American Economic Association posts videos of a diverse set of academic economists discussing their relevant research on the website http://diversifyingecon.org/index.php/Videos_on_economists_and_their_research. In the video of a TED talk linked below, Eleni Gabre-Madhin describes her work creating a commodity market in Ethiopia to reduce risk faced by farmers: https://www.ted.com/talks/eleni_gabre_madhin_a_commodities_exchange_for_e thiopia.

[return to top]

Solutions to End of Chapter Problems Students have access to solutions for the odd-numbered problems as well as video problem walkthroughs for problems 1 and 6.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

32


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

5.1

a.

Given that these are actual gambles offered in Las Vegas, you shouldn’t be surprised to learn that they are unfair in the casino’s favor. To verify this claim, we can compute the expected payoff from gamble 1,

2,000  18   20   38   1,000    38   1,000   38  52.6,     and from gamble 2,

1,000  1   37   38  17,500    38   500   38  26.3.     They are both negative, not zero as required of fair gambles. b.

To determine which gamble Wen would take, compute the expected utility from gamble 1,

 18   20   38  10,000  1,000   38  10,000  1,000  99.61,     and from gamble 2,

 1   37   38  10,000  17,500   38  10,000  500  99.27.     The first is higher, so Wen would choose gamble 1. c.

5.2

a.

The expected utility from not taking either gamble is the same as the utility from current income ($10,000) with certainty: 10,000  100 . This is higher than the expected utility from either gamble, verifying that Wen, who is risk averse, wouldn’t want to take fair gambles, let alone the unfair gambles offered in Las Vegas.

E(1) = 0.50(100) + 0.50(–100) = 0 E(2) = 0.75(100) + 0.25(–300) = 0 E(3) = 0.90(100) + 0.10(–900) = 0

b.

Assume current income is $1,000. Then utility of income graph is:

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

33


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

5.3

c.

Bet 1 will be preferred since it has smaller variability.

a.

Expected utility without insurance is 0.75 log(10,000) + 0.25 log(9,000) = 3.9886.

b.

Expected utility with insurance is log (9750) = 3.9890. This is greater than the expected utility from part a.

c.

The individual will pay up to point where expected utility with insurance equals the expected utility without. The expected utility with insurance having premium cost P is log (10,000 – P). Therefore, we set log (10,000 – P) = 3.9886 to find the highest premium Mr. Fogg is willing to pay. Raising both sides to the power of 10, 103.9886 = 10,000 – P = 9.741, implying the maximum premium is $259.

d.

5.4

a. b.

The fair insurance premium equals the expected loss: E(L) = .30  1,000 = $300. Since $300 > $259, he will not buy this insurance even though it is fair. This is an example of moral hazard.

I  0.5 45,000  0.5 55,000  446.65I  49,875 . ln( I )  0.5 ln( 45,000)  0.5 ln(55,000)  10.815.

I  e10.815  49762. c.

1 0.5 0.5 105    2.02 105 I   49,504 . I 45,000 55,000 2.02 The functions exhibit increasingly greater risk aversion. This is an illustration of the general principle that the degree of (relative) risk aversion for the utility function  I R / R U (I )    ln( I )

R 1 R0

is given by 1 – R. 5.5

a.

U = ln($18,000) = 9.798.

b.

U = ln($18,300) = 9.815.

c.

If Molly invests $100 in the trip, she will have a wealth of $17,900 if Crazy Eddie does not have the set and $18,200 if he does. E(U) = 0.5 ln(17,900) + 0.5 ln(18,200) = 9.801. Since this exceeds the utility from part a, it is worth the trip.

5.6

a.

Strategy one: Outcome

Probability

12 eggs

0.5

0 eggs

0.5

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

34


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Expected value = (0.5  12) + (0.5  0) = 6.

Strategy two: Outcome

Probability

12 eggs

0.25

6 eggs

0.5

0 eggs

0.25

Expected Value = (0.25  12) + (0.5  6) + (0.25  0) = 3 + 3 = 6. b.

5.7

5.8

c.

Gains from diversification are offset by costs of extra trips, so there may be an optimal number of such trips.

a.

The expected value of the prize is $7,500. The value of the option is (0.5  $0) + (0.5  $8,000) = $4,000. So the option is not worth what is being asked.

b.

The option promises income of 10,500 if the ring is behind the door and 3,500 if the goat is behind the door. Hence, as shown in part a, expected income is lower if the option is purchased. However, the variability of income is lower with the option (ranging only between 3,500 and 10,500 rather than between 0 and 15,000), so a particularly risk-averse contestant may choose the option.

a.

Now Equations (1) in Application 5.4 are k(200) – L = 0 and k(250) – L = 10. The solution is k = 0.2, L = 40. Net cost for this purchase is 0.2 × 220 – 40 = $4 (you need to buy 0.2 of a Microsoft share at $220 a share, but can use the $40 loan to pay for part of it). That $4 is now the value of the option, which is lower than in the application because the strike price is higher.

b.

Now the net cost is (0.2  225) – 40 = 5. This is the value of the option. This is an increase from the original cost because the replicating portfolio is now more expensive. Notice that this case is a bit artificial because the rise in Microsoft price does not affect expected future prices.

c.

Now the replicating portfolio is found by solving the equations k(190) – L = 0 and k(260) – L = 10. The solution is k = 0.4, L = 8. The cost of this portfolio is $4.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

35


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

This is an increase from the value in the application because of the increased volatility in the share price.

5.9

d.

Interest payments can be treated as creating a difference between the amount borrowed and the amount repaid. If, for example, the interest rate were 5 percent per period (paid at the end of the period) the amount repaid in the replicating portfolio would be 1.05L. For the example in the application the replicating k would remain at 0.4, but the loan amount would fall to 80/1.05 = 76.2. Hence the cost of the portfolio would rise to 220 × 0.4 – 76.2 = 11.8, which would be the value of the option now.

a.

With the first utility function, we have I  0.5 116,000  0.5 98,000 , implying I = 106,800, or a certainty equivalent yield of 6.8%. With the second utility function, we have ln I  0.5 ln(116,000)  0.5 ln( 98,000) , implying I = 106,600, or a certainty equivalent yield of 6.6%. Finally, with the third utility function, we have

1 0.5 0.5   , I 116,000 98,000 implying I = 106,200, or a certainty equivalent yield of 6.2%. With any of these utility functions, stocks offer a much higher certainty equivalent yield than do bonds. b.

With this extreme utility function,

I 10  0.5(116,000) 10  0.5(98,000) 10 implying I = 103,200, or a certainty equivalent yield of 3.2%. Even with this extreme risk-aversion stocks have a certainty equivalent yield much higher than bonds. Hence, the high yield for stocks is a paradox. 5.10

a.

Leah’s initial situation without insurance is shown as point A in the graph. Full, fair insurance moves her to point B. Full insurance at unfair terms moves her to a point such as C. Point B is on a higher indifference curve than A, so full, fair insurance certainly makes her better off. As long as the terms of unfair insurance aren’t too unfair in the insurance company’s favor, C will also be on a higher indifference curve than A, as shown.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

36


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

b.

If Leah is risk neutral, her indifference curves are straight lines. Full, fair insurance moves her from uninsured point A to point B. She is indifferent between the two outcomes because the insurance line has the same slope as her indifference curve. Unfair insurance in her favor would move her to point C, which is on a higher indifference curve than A, so she would accept it. Unfair insurance in Gecko’s favor would move her to point D, which is worse for her than A.

c.

Return to the graph from part a, but imagine that point A is quite close to B. Next imagine magnifying the graph, shown below. The bend in Leah’s indifference curve has been ―ironed out‖ in the magnification, making her look almost risk neutral, with indifference curves that look almost like straight lines. She would reject even moderately unfair insurance such as indicated by point C. Indeed, full, fair insurance shown by point B is hardly better for her than she is without insurance at point A.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

37


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Chapter 6: Game Theory Purpose and Organization of the Chapter This chapter provides an introduction to game theory, providing students with tools to analyze simple games. We have placed this chapter so early in the text for several reasons. Of course, it is useful to have this material covered before it is used in later chapters, most importantly in studying the strategic interaction between oligopoly firms in Chapter 13. But placing it up front in the text in Part 3 on uncertainty and strategy drives home the point that game theory is not just for oligopolies. We view game theory as the natural generalization of the maximizing decision maker (the foundation of our analysis of consumer and producer theory) to settings with two or more decision makers whose decisions interact. Game theory applies to the interactions ranging from that between individual people (say neighbors regarding how they maintain their yards) to that between nations (say countries deciding whether or not to sign a new climate treaty). The chapter begins with a description of the components of a game: players, strategies, and payoffs. Then it turns to the fundamental equilibrium concept for simultaneous games, Nash equilibrium. Then it turns to sequential games and the equilibrium concept for these, subgameperfect equilibrium. It ends with more advanced topics including repeated games, games with continuous actions, and so forth. The focus throughout is on methods to solve for equilibrium. The chapter deals mostly with classic, abstract games such as the Pris-oners’ Dilemma and the Battle of the Sexes. As mentioned above, applications of game theory to analyze imperfect competition have been moved forward to Chapter 13. We only briefly touch on games of asymmetric information, games in which one player has better information about the world than another. Such games are more complex than those

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

38


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

studied in the chapter. A whole later chapter, Chapter 16, is devoted to the study of games of asymmetric information.

Lecture and Discussion Suggestions The various game theory concepts introduced in this chapter are, we believe, best illustrated with specific games in lecture. Three simple ones to begin with might be the Prisoner’s Dilemma, Matching Pennies, and the Battle of the Sexes. The various concepts can be introduced one after the other in analyzing these simple games, as demonstrated in the text. One way to inject fun into the lecture is to have students play games in class. Avinash Dixit, ―Restoring Fun to Game Theory,‖ Journal of Economic Education 36: 205-219 (Summer 2005) provides a handful of tested classroom experiments that are both entertaining and illustrative of important concepts. He advises using monetary payoffs and advises providing a brief discussion immediately after the game of the theoretical points raised. Students often raise the issue that the equilibrium predictions in, for example, the Prisoners’ Dilemma and the Battle of the Sexes are off target because of dynamic aspects of the game, possibilities of threats and other communications, and the possibility of irrational play. One could respond that any of these additional factors would be worthwhile to consider, but it should be handled by posing a different game that can be analyzed in its own right. For example, the text shows how the outcome in the Prisoners’ Dilemma changes when the game is repeated. Regarding the possibility of irrational play, one could acknowledge the new research on behavioral economics that is receiving a great deal of attention recently and point the student to Chapter 18. Some practitioners contend that the crucial skill is less the ability to solve a game with a sophisticated equilibrium concept than the ability to boil down a particular economic situation into a simple game that can then be analyzed. One way to have students practice modeling economic situations as games is to assign a project that has students take a situation from student life, from current events in the newspapers, or from movies (Dixit 2005 provides some nice examples of movies along these lines), and model it as a game, and then solve for equilibrium. There is a lot of material in the chapter. To pare it down to something more manageable, we would suggest omitting the sections on repeated games and games with continuous actions. It would be nice to cover mixed strategies and sequential games, but if needed one of these topics could also be omitted. Drawing the best-response-function diagrams is somewhat difficult, so we would recommend omitting this topic for less analytical courses. We warn you that the diagrams are particularly tricky to draw for games with two actions. We included them in the text so that the student could see the connection between them and the diagrams for games with continuous actions (where the diagrams are easier to draw and are quite helpful for the analysis). We would recommend omitting them in all but the most advanced courses.

Cengage Supplements The following product-level supplements provide additional information that may help you in preparing your course. They are available in the Instructor Resource Center.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

39


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

 

Test Bank PowerPoint Slides

Chapter Objectives The following objectives are addressed in this chapter: 06.01 Illustrate a game using an outcome matrix and a game tree. 06.02 Derive the Nash equilibrium using the concept of best responses. 06.03 Understand choices that involve randomization. 06.04 Find solutions for games in which players take turns. 06.05 Solve games that players play over and over. 06.06

Solve games in which players have continuous action.

[return to top]

What's New in This Chapter We retained all the concepts from before but made some major updates to the exposition and applications. 

We changed the game used extensively throughout the chapter as an illustration of multiple equilibria and sequential games from the Battle of the Sexes to the GroupScience Project. This is a big change, requiring all of the text and figures in Sections 6.6 and 6.7 to be revised. The change is worth it, as the Group-Science Project is better in at least three ways. First, it avoids archaic gender roles. Second, the application is less artificial and indeed may be something students have experience with. The game also has the nontrivial pedagogical advantage of making it obvious which coordination outcome each player prefers (they prefer the subject starting with the same letter as their names). Problem 6.6 has been updated to explore the new game.

Application 6.1 used to feature the movie A Beautiful Mind. The movie is becoming a bit dated and the dating game might strike students as sexist. We replaced the application with Application 6.1: Game Theory on the Screen, covering a recent, popular miniseries about chess (The Queen’s Gambit) as well as other movie applications students might find interesting (the Battle of Wits in The Princess Bride and duels in movie history).

We replaced the dated application to the standard war between HD-DVD and Bu-Ray in Application 6.3 with a discussion of big-tech dominance. We discuss the important idea of how a platform market can come to be dominated by a single firm, and explain how this idea might apply to Google, Amazon, Apple, Facebook, and Microsoft, firms that students interact with daily and read about in the newspapers (at least those who follow recent privacy and antitrust cases). Problem 6.4 has been updated to reflect the new concepts and firms in the application.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

40


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

[return to top]

Additional Resources The American Economic Association posts videos of a diverse set of academic economists discussing their relevant research on the website http://diversifyingecon.org/index.php/Videos_on_economists_and_their_research. In the video linked below, Juan Camilo Cárdenas describes laboratory studies that may help us better understand environmental problems such as overfishing and carbon pollution, related to the Tragedy of the Commons studied in this chapter. https://www.core-econ.org/the-economy/book/text/04.html#BqEOGDX766Q.

[return to top]

Solutions to End of Chapter Problems Students have access to solutions for the odd-numbered problems as well as video problem walkthroughs for problems 1 and 10.

6.1

6.2

a.

A plays Up; B plays Left.

b.

A’s dominant strategy is Up. B does not have a dominant strategy.

c.

If v = $3, RTS = 1/2 > w/v = 1/3, the manufacturer will use only L. For q = 20, L = 20; q = 40, L = 40; q = 60, L = 60. Now the manufacturer’s expansion path is the L axis.

a.

A plays Down; B plays Right.

b.

A’s dominant strategy is Down. B does not have a dominant strategy.

c.

Yes. A’s equilibrium payoff increases from 3 to 4. A comparison of the games in Problems 6.1 and 6.2 suggests the possibility that ―burning money‖ can be beneficial in a strategic setting.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

41


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

6.3

a. A Up

Down

B

B Right

Left

3, 3

Right

Left

5, 1

4, 4

2, 2

b. A Up

Down

B

B Right

Left

3, 3

5, 1

Right

Left

4, 4

2, 2

c. B

Up

Left | Up Left | Down

Left | Up Right | Down

Right | Up Left | Down

Right | Up Right | Down

3, 3

3, 3

5, 1

5, 1

2, 2

4, 4

2, 2

4, 4

A

Down

There are two Nash equilibria: first, A plays Up, and B plays ―Left | Up, Left | Down‖; second, A plays Down, and B plays ―Left | Up, Right | Down.‖ The second is a subgame-perfect equilibrium. 6.4

a.

One pure-strategy Nash equilibrium is for Facebook to Invest Heavily and MySpace to Slacken. The other is the reverse (Facebook Slackens and MySpace Invests Heavily).

b.

Let a be the probability that Facebook Invests Heavily and 1 – a that it Slackens. Given Facebook’s mixed strategy,

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

42


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

if MySpace Invests Heavily, its expected payoff = (a)(0) + (1–a)(3) = 3 – 3a; if MySpace Slackens, its expected payoff = (a)(1) + (1–a)(2) = 2 – a. Equating the two expected payoffs, 3 – 3a = 2 – a, implies a = ½. Letting b be the probability that MySpace Invests Heavily and 1 - b that it Slackens, calculations similar to the preceding show that b = 1/2 in the mixed-strategy Nash equilibrium.

c.

Let I stand for Invest Heavily and S for Slacken. Toshiba’s contingent strategies are in the column headings of the following normal form: MySpace I|I I|S

I|I S|S

S|I I|S

S|I S|S

0, 0

0, 0

3, 1

3, 1

1, 3

2, 2

1, 3

2, 2

I Facebook

S

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

43


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

d.

Refer to the underlining method in the normal form in part c. There are three Nash equilibria indicated by the boxes with both payoffs underlined.

e.

Proper subgames circled below.

The subgame-perfect equilibrium is for Facebook to play I and for MySpace to play ―S | I, I | S.‖ In the other Nash equilibria, either MySpace irrationally plays S after Facebook plays S or MySpace irrationally plays I after Facebook plays I. 6.5

a. B Shirk

Work

Shirk

0, 0

4, -2

Work

-2, 4

1, 1

A

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

44


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

6.6

b.

Both shirk.

c.

Shirking is a dominant strategy for both. Game resembles the Prisoners’ Dilemma.

a.

The normal form becomes B Astro

Bio

Astro

4, 2

0, 0

Bio

0, 0

2, 4

A

The mixed strategies do not change, and the best-response-function diagram does not change from Figure 6.4 in the text. b.

The normal form becomes B Astro

Bio

Astro

4, 1

0, 0

Bio

0, 0

1, 4

A

The mixed-strategy Nash equilibrium is for A to play Astro and Bio with probabilities 4/5 and 1/5, respectively, and for B to play Astro and Bio with probabilities 1/5 and 4/5, respectively.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

45


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

c.

The normal form becomes B Astro

Bio

Astro

2, 1

1/2, 1/2

Bio

0, 0

1, 2

A

The mixed-strategy Nash equilibrium turns out to be the same as in part b. 6.7

a.

Using the underlining method shows that playing Rat is a dominant strategy for both and that both Ratting is a Nash equilibrium.

b.

Expected payoff in equilibrium is 1  ( g )(1)  ( g 2 )(1)  ( g 3 )(1)    (1)(1  g  g 2  g 3  )  1 /(1  g ).

If a player deviates to Rat in the first period, his or her payoff is 3 in the first period and 0 from then on. For the trigger strategies to be an equilibrium, 1/(1 – g)  3, implying g  2/3. c.

The expected equilibrium payoff is the same as in part b, 1/(1 – g). If a player deviates from tit-for-tat, he or she earns 3 in the first period, 0 in the second, and then the players return to the original equilibrium for an expected payoff of

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

46


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

3  ( g )(0)  ( g 2 )(1)  ( g 3 )(1)    2  1  g (1  1)  (1)( g 2  g 3  )  2  g  (1)(1  g  g 2  g 3  )  2  g  1 /(1  g ).

For this payoff from deviating to be less than the equilibrium payoff, 2 – g ≤ 0, implying g  2. This is impossible since g is a probability. So players cannot sustain cooperation on Silent using tit-for-tat. 6.8

The pure-strategy Nash equilibrium is for A to play Up and B to play Left.

6.9

a.

There are four pure-strategy Nash equilibria, one in which none of the three locate in the mall and three different ones in which two locate in the mall and the third does not (so three different ones, one for each different left-out store A, B, and C).

b.

Playing cooperatively, they might reach one of the three outcomes in which two of the stores locate in the mall and the third does not. The sum of the payoffs is the highest in these outcomes, 4. The stores locating in the mall may pay the left-out one for not locating there, perhaps each paying 2/3 so that total surplus is split evenly.

a.

Following the logic of equation (6.6), the marginal benefit of an additional sheep for A is

6.10

300 – 2sA – sB. Setting the marginal benefit equal to the marginal cost 0 gives sA = 150 – sB/2. Similarly, sB = 150 – sA/2. Solving simultaneously shows that the Nash equilibrium is s *A  s *B  100.

b.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

47


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

c.

The marginal benefit of an additional sheep for A is 330 – 2sA – sB. Setting the marginal benefit equal to the marginal cost 0 gives sA = 165 – sB/2. As before, sB = 150 – sA/2. Solving simultaneously shows that the Nash equilibrium is s *A  120, s *B  90.

Chapter 7: Production Purpose and Organization of the Chapter This chapter describes production functions. Returns to scale and substitution possibilities are stressed as descriptive and analytical concepts for studying production in the real world. Each of these concepts offers difficulties to students. Returns to scale is often confused with returns to a single factor. That confusion is easily remedied by examining constant returns with fixed proportions production functions since these require the simultaneous increase in inputs specified in the returns to scale measure. The simplified isoquant maps in Figure 7.3 are also helpful in making this point as do Review Questions 5-7. Clarifying importance of input substitution can probably best be accomplished by spending some time on the fixed-proportions production function. That presentation might also be supplemented by reading the Cobb-Douglas numerical example at the end of the chapter. This extended example can be skipped if time is short. We think it is probably best not to introduce the elasticity of substitution formally. Problem 7.6 provides a nice intuitive illustration of

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

48


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

substitution possibilities arising from when there are several production techniques available, however.

Lecture and Discussion Suggestions Chapter 7 will generally require two lectures: one on theory, one on policy or empirical applications. For the theory lecture, one could proceed in much the same way that consumer theory was developed (by asking about trade-offs for example). The important difference in production theory is that production functions are measurable, and one is therefore more interested in their specific shapes. Both returns to scale and substitution possibilities should probably be covered in a lecture to reinforce those concepts in students’ minds and to plant the notion that they are not simply useless baggage from the text. An empirical lecture might discuss both traditional production function examples (i.e., the everpopular case of beer—see Application 7.4) and nontraditional examples (a school, a doctor’s office, etc.). The purpose of all the examples should be to demonstrate the importance of questions of substitution and of scale. Any of these illustrations could then be pushed a bit further by asking about how firms should be regulated (if at all) when they experience economies of scale over a broad range of output. The theory and applications of productivity change also provide a useful starting point for discussions on the subject matter of Chapter 7. Although discussions of the production function concept per se can be excruciatingly dull, students are very interested in the productivity question and that question offers the possibility for wide-range discussion. We particularly like questions about how computer technology may have affected overall measures of productivity growth (see Application 7.5). Students could be asked to speculate on how the emerging technology of today (robots, cloud computing, big data, smartphones) might show up in tomorrow’s productivity statistics.

Cengage Supplements The following product-level supplements provide additional information that may help you in preparing your course. They are available in the Instructor Resource Center.  

Test Bank PowerPoint Slides

Chapter Objectives The following objectives are addressed in this chapter: 07.01

Describe the relationship between a firm’s inputs and outputs using a production function.

07.02

Understand the importance of marginal product and why it diminishes.

07.03

Illustrate the relationship between inputs and output using an isoquant diagram.

07.05

Explain the difference between constant, increasing, and decreasing returns to scale.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

49


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

07.06

Understand the difference between input substitution and technological progress.

[return to top]

What's New in This Chapter The chapter is substantively the same as in the previous edition. Some examples or data has been updated. 

We updated some of the peripheral examples in Applications 6.1, 6.3, and 6.5.

We updated the dates and data used in Section 7-6b on multifactor productivity.

[return to top]

Additional Resources The American Economic Association posts videos of a diverse set of academic economists discussing their relevant research on the website http://diversifyingecon.org/index.php/Videos_on_economists_and_their_research. Several videos touch on issues related to production. In the video linked below, Lucy Wang discusses how advances in information technology in healthcare can avoid opiod-related deaths. https://vimeo.com/250465659. In this second video, Steven Bednar notes how allocating scarce time to physical education can boost student academic achievement. https://vimeo.com/310389478.

[return to top]

Solutions to End of Chapter Problems Students have access to solutions for the odd-numbered problems as well as video problem walkthroughs for problems 7 and 9.

7.1

a.

K = 6, q = 6K + 4L = 6(6) + 4L = 36 + 4L. If q = 60, 4L = 60 – 36 = 24, L = 6. If q = 100, 4L = 100 – 36 = 64, L = 16.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

50


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

b.

K = 8, q = 6K + 4L = 6(8) + 4L = 48 + 4L. If q = 60, 4L = 60 – 48 = 12, L = 3. If q = 100, 4L = 100 – 48 = 52, L = 13.

7.2

c.

RTS = 2/3: If L increases by 1 unit, can keep q constant by decreasing K by 2/3 units.

a.

When K = 10, the production function is q = 2K + L = 2(10) + L = 20 + L. If q = 100, L = 100 – 20 = 80.

b.

When K = 25, the production function is q = 2(25) + L = 50 + L. If q = 100, L = 50.

c.

RTS 

25  10 1  . 50  80 2

Since the RTS is the slope of the isoquant and the isoquant is linear, the RTS (slope) is the same at every point. d.

See graph in part c. This production function has linear isoquants.

e.

K = 10

q = 3(10) + 1.5L = 30 + 1.5L.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

51


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

q  100  30  1.5L so L  46.67. K  25 q  75  1.5L q  100  75  1.5L so L  16.67. The isoquants are still straight lines with slope –1/2, but any particular combination of inputs now represents a larger q than before. 7.3

a.

q 100  . L L

b.

AP 

c.

Graph in part b. Since the APL is everywhere decreasing, then each additional worker must be contributing less than the average of the existing workers, bringing the average down. Therefore, the marginal productivity must be lower than the average. Here MPL  APL 2.

L

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

52


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

7.4

a.

b.

2,000 = 20 101K or K = 10,000/101 – 99.01

RTS  c.

 K 1. L

2,000 = 20 401K or K = 10,000/401 = 24.94 RTS = 0.06.

d.

In b K/L = 100/100 = 1. In c K/L= 25/400 = 0.06 If L = 201, K = 10,000/201 = 49.75 RTS = 0.25 = K/L = 50/200.

e. 7.5

With q = 40 KL , none of the RTS values calculated before changes.

a.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

53


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Will operate at the vertex of the isoquants b.

Hire 20 workers, q = 1,000.

c.

Depends on whether grapes can be sold for a price exceeding average cost.

d.

Choice would depend on clipper costs and wages for ambidextrous workers. 7.6

a., b.

function 1: use 10K, 5L function 2: use 8K, 8L c.

5K, 2.5L, and 4K, 4L so this 50–50 mix requires 9K, 6.5L. A 75–25 mix would need 7.5K, 3.75L, and 2K, 2L for a total of 9.5K, 5.75L. Here, fractions of K and L represent fractions of hours using whole units of capital and labor.

d.

A plot of the points yields a linear q = 40,000 isoquant.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

54


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

7.7

a.

In Equation 6.7, A  10 and a = b = 1/2.

b.

If we use 2K, 2L, have q  A  2 K   2 L   2a b AK a Lb  2a b q. a

b

Then if a + b = 1, this is twice q . c., d. From part b, it follows that output will less than double or more than double if a + b < 1 or a + b > 1.

7.8

e.

Function can exhibit any returns to scale desired depending on the values of a and b.

a.

If a + b = 1, MPK = aA(L/K)b = aA(K/L)-b and MPL = bA(K/L)a . Hence, each declines as its input is increased.

a K  . b L

b.

RTS  MPL / MPK 

c.

It is obvious from part b that as K/L falls, RTS falls.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

55


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

7.9

a.

b.

.

K = 10, L = 10 APL = q/L = 1,000/10 = 100 boxes per hour per worker.

c.

If q = 200

KL = 1000

KL = 5, or, KL = 25.

Isoquant shifts to q0'. Now, if K = 10 L = 2.5 APL = q/L = 1,000/2.5 = 400 boxes per hour per worker. d.

Now

⁄ ⁄ ⁄( ) √ This last equation for the isoquant shifts toward the origin as time passes. More output can be produced with the same inputs as technology progresses. To solve part b for this new production function, substitute into one of the ⁄ ⁄ equations, , just derived: ⁄ . Less labor is required over time to produce a given output with fixed capital. 7.10

a.

The function exhibits constant returns to scale because its exponents sum to one.

b.

Let X denote the proportional change in the ―generic‖ variable X. That is,

dX dt X  . X Then the equation in the text can be written q  A  aK  (1  a) L . So technical change can be estimated by A  q  aK  (1  a) L . Notice that all of the terms on the right of this equation can be measured. c.

Use the math facts that if z = xy z  x  y and if z 

x y

z  x  y . Hence

K q q  L     A  a  K  L   A  a   . Hence, the change in q/L is a good L L

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

56


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

measure of changes in total factor productivity if a is small or if the capital/labor ratio is not changing by very much.

Chapter 8: Costs Purpose and Organization of the Chapter This chapter defines input costs and relates these cost concepts to the production function. In the initial section on definitions, some care is taken to distinguish between economic and accounting costs and to show why economic concepts are more appropriate for theoretical investigations. The concept of opportunity cost for the firm is stressed throughout this discussion. The bulk of the chapter concerns the relationship between input costs and the production function. The firm’s expansion path is introduced to show that relationship in the long run. Long-run total, average, and marginal cost curves are derived from the expansion path. These cost curves represent the primary tools introduced in this chapter. The final set of ideas introduced in the chapter concern the distinction between the short and long run by holding one input (capital) constant in the short run. That technique permits the usual set of U-shaped cost curves to be derived. To avoid unnecessarily cumbersome graphs, the average variable cost curve is not introduced explicitly, though the notion of dividing shortrun costs into fixed and variable components is mentioned as is the role of variable costs in defining the shut-down point. The chapter concludes with an extension of the numerical example from Chapter 7. This example is helpful both in illustrating the duality between production and cost functions and in showing the short run-long run distinction.

Lecture and Discussion Suggestions There seems no ready escape from a flood of cost curves when lecturing on this chapter. The curves are both important and relatively difficult to derive so it is probably impossible to avoid repeating the text to some degree. Over the years we have shortened the material on differences between the short and long runs, and we believe the profession has been moving that way, too (especially in the theory of industrial organization). For lectures, therefore, we would suggest focusing on long-run average and marginal cost concepts and have students pick up most of the information on the short run on their own. The main idea to get across, if the topic is covered at all, is that the firm will be able to do better (in this case attain the same or lower costs) if it has more margins of adjustment (in this case able to adjust capital as well as labor). Two applications may help make the lectures more entertaining and enlightening. If you leaf all the way back to the first chapter, you will find Application 1.2. This removes some of the mystery behind opportunity cost by showing how it arises in a decision central to students’ lives, that of attending college. The salary foregone over one’s college years is a real economic cost, along with tuition, room, board, etc. Application 8.3 should interest students because it provides cost estimates for real-world industries.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

57


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Semantic Issues We think it may help instructors to discuss some semantic issues that took us a while to sort out but think we finally have a good handle on. We would not recommend presenting these ideas in this way to your students. We provide this background to resolve some inconsistencies that would only occur to experts and may help these experts answer some thorny student questions should they arise.

Modeling the Short and Long Run The textbook model of fixed costs in the short run is a very standard one used in the profession. Still, there are some apparent inconsistencies among definitions that can be resolved if the model is thought about in the right way. The textbook says that fixed costs equal expenditures vK1 on capital K1 that cannot be adjusted in the short run. But isn’t vK1 a sunk cost which should be excluded from economic costs in the short run? Yes and no. By definition, economic costs are those relevant for an economic decision. Thus, they will vary depending on the decision under consideration and the decision maker’s perspective. Everything can be sorted out using a three-stage model which is too complicated to share with students but may help clarify the instructor’s understanding. Here’s the model in a nutshell. It involves some uncertainty about market conditions. Let’s capture that uncertainty by assuming that the firm starts out with some forecast of what market price P will be but doesn’t know it exactly. Stage 1: The firm builds capacity K1. Stage 2: The firm learns the market price P. It chooses output to maximize profits knowing P. It is free to adjust labor but must stick with capital K1. This is the short run. Stage 3: The market conditions stay the same, so the price is still P but now the firm can produce output by adjusting both capital and labor. This is the long run. Let K2 be the capital it ends up choosing. The following figure gives the timeline.

If we put the firm in stage 2 and look at its shut-down decision having already invested in its capital, then vK1 is indeed sunk and should not factor into economic costs going forward. This is exactly why the firm compares revenue from continued operation only to variable costs in Chapter 9; if variable costs can be covered, the firm should operate; if not the firm should shut

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

58


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

down. The firm may end up operating even if it can’t also cover because that, being sunk, is an accounting but not a true economic cost at that point. However, when the firm is allowed to vary its capital in stage 3, the long run, capital costs vK2 are true economic costs, so should be added in. If we don’t also add capital costs into the short-run cost function used for comparison, this leads to the absurd situation where the inability to adjust capital somehow helps the firm by lowering its costs. To avoid this absurdity, to compute short-run costs, we consider the perspective of the firm in stage 1, when it is choosing the level of K1 to invest in. Then vK1 is not sunk yet, so is a true economic cost. Yet it is fixed in the sense that it cannot be adjusted until stage 3 so is fixed during stage 2. It isn’t important to keep this complicated structure in mind, but it should at least provide some comfort that all the definitions can, with some effort, be made consistent with each other. It also reinforces the point that economic cost is a relativistic concept, depending on who the decision maker is (individual? firm? social planner?) as well as which decision it is making (matriculating in college or dropping out? firm output or entry?).

Returns to Scale vs. Economies of Scale Recent editions introduced the definitions of economies and diseconomies of scale. Earlier editions just talked about returns to scale. The two sets of concepts are related. For certain classes of production functions, they are synonymous. For example, a production function that is everywhere in-creasing returns to scale will have a downward sloping AC curve, thus exhibiting economies of scale; a production function that is everywhere constant returns to scale will have a flat AC curve. For production functions that have different sorts of returns to scale over different input regions, the relationship between the two sets of concepts is more complicated. Published scholars differed on the true nature of the relationship. Witness the series of articles in the Journal of Economic Education in the 1980s and 1990s that went back and forth, sometimes contradicting each other. The instructor can be forgiven for being confused. Lila J. Truett and Dale B. Truett (1990) ―Regions of the Production Function, Returns, and Economies of Scale: Further Considerations,‖ Journal of Economic Education 21: 411-419 does perhaps the best job helping the instructor wade through the confusion, showing that some of the difference of opinion depends on whether one is looking at a large (arc) change in inputs or a small (point) change.

Cengage Supplements The following product-level supplements provide additional information that may help you in preparing your course. They are available in the Instructor Resource Center.  

Test Bank PowerPoint Slides

Chapter Objectives The following objectives are addressed in this chapter: 08.01 Understand basic cost concepts such as economic and accounting costs.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

59


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

08.02 Explain how a firm determines its cost minimizing mix of inputs. 08.03

Understand the relationship between output levels and total costs, average costs, and marginal costs, both in the long run and short run.

08.04

Understand what factors cause a firm’s production costs to change.

[return to top]

What's New in This Chapter The chapter is substantively the same as in the previous edition. The exposition has been streamlined in a few places. Data and examples have been updated in text and applications. 

The data on licensing fees in the text has been updated.

The concluding passage on the future of deregulation in Application 7.1 has been rewritten to update it.

The passage on congestion tolls in Application 7.4 has been updated to mention the most sophisticated current systems that can monitor the extent of congestion and condition the fees on that.

[return to top]

Additional Resources For a view of how a project manager thinks about estimating costs, perhaps as the basis for a bid on a construction project, see the following website. https://www.cmu.edu/cee/projects/PMbook/05_Cost_Estimation.html.

[return to top]

Solutions to End of Chapter Problems Students have access to solutions for the odd-numbered problems as well as video problem walkthroughs for problems 1 and 8.

8.1

a.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

60


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

RTS = 1/2 since if L is increased by one, K can be reduced by 1/2 while holding q constant.

8.2

8.3

b.

Since RTS = 1/2 < w/v = 1, the manufacturer will use only K. For q = 20, K = 10; q = 40, K = 20; q = 60, K = 30. The manufacturer’s expansion path is simply the K axis.

c.

If v = $3, RTS = 1/2 > w/v = 1/3, the manufacturer will use only L. For q = 20, L = 20; q = 40, L = 40; q = 60, L = 60. Now the manufacturer’s expansion path is the L axis.

a.

Because the manufacturer does not change its input mix in response to changing input prices, the cost of producing 1,000 gumballs will always be the cost of one worker and two presses: 2v  w .

b.

Added units of 1,000 gumballs can be produced by just replicating the underlying technology q times.

c.

Average and marginal cost are both 2v  w .

d.

When v  3, w  5 : AC  MC  2v  w  11.

e.

Now AC  MC  2v  w  17.

a.

This is a cubic cost curve, resembling Figure 8.3(d).

b.

AC = TC/q = q – 30q + 350.

2

This is a parabola. It reaches a minimum at the axis of symmetry: q = –(–30)/2 = 15. At q = 15, AC = 225 – 450 + 350 = 125.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

61


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

c.

At q = 15, MC = 3(225) – 900 + 350 = 125.

d.

8.4

a.

q = 2 H q/2 =

H

H

q2 4

TC = wage rate × H = 2q2 AC = TC/q = 2q b.

c.

q=4

TC = 2(4)2 = 32

q=6

TC = 2(6) 2 = 72

q=8

TC = 2(8) 2 = 128

q=4

AC = 2(4) = 8

q=6

AC = 2(6) = 12

q=8

AC = 2(8) = 16

The TC and AC curves are shown in the graph. Notice that the convex shape of TC implies that AC is always increasing.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

62


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

8.5

a.

K = 100, q = 2 100L . q = 20 L.

q = 2 KL

L

q 20

so L 

q2 400

STC  vK  wL  100 

q2 100

b.

q 50 q  25, STC  106.25, SAC  4.25 SMC  0.5 q  50, STC  125, SAC  2.5, SMC  1 q  100, STC  200, SAC  2, SMC  2 q  200, STC  500, SAC  2.5, SMC  4

SMC  If If If If c.

8.6

d.

The curves intersect at q = 100. As long as the marginal cost of producing one more unit is below the average cost curve, average costs will be falling. Similarly, if the marginal cost of producing one more unit is higher than the average cost, then average costs will be rising. Therefore, the SMC curve must intersect the SAC curve at its lowest point.

a.

It is a constant returns to scale production function. The average cost function , which is flat (that is, independent of ). So this is a case on the boundary between economies and diseconomies of scale (you might say constant economies of scale).

b.

(

)

c.

(

) , which is increasing in . So exhibits diseconomies of scale.

d.

(

)

, which is falling in . So exhibits economies of scale.

implies ( (

For economies of scale, .

(

)

.

) )

⁄ , implying

. For diseconomies of scale,

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

63


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

8.7

To minimize, costs should equate the marginal productivities of labor in each plant. If labor were more productive in one plant than another, costs could be lowered by moving workers. a.

MPL1 = MRL2.

5/2 L1 = 5/ L2 .

2 L1 = L2 .

L2 = 4 L1.

q1 = 5 L1 ; q2 = 10 L2 = 10 4L1 = 20 L1 . Hence q2  4q1 b. STC (plant 1)  25  wL1  25  q12 25 STC (plant 2)  100  q22 / 100

STC = STC (plant 1) + STC (plant 2)

Since q1 

q 5

and q2 

4q 5

Substitution yields:

q12 q22 0.8q 2 q2   125   125  25 100 100 125 125 q AC   q 125 2q MC  125 STC  125 

MC(100) = $1.60 MC (125) = $2.00 c.

MC (200) = $3.20.

In the long run because of constant returns to scale, can change K so it doesn’t really matter where production occurs. Could split evenly or produce all output in one plant. TC = K + L = 2q. AC = 2 = MC

8.8

d.

If there were decreasing returns to scale, then should let each firm have equal share of production. AC and MC, not constant anymore, are increasing functions of q so do not want either plant to be too large.

a.

With a  b  0.5, TC  Bqv 0.5 w0.5 . The associated is flat, so we are right between economies and diseconomies of scale. Input prices have equal exponents so are in a sense equally important.

b.

Returns to scale for this production function are measured by , the reciprocal of the exponent on q in the function. If , implying increasing returns to scale, the exponent on will be less than 1. This means costs rise less than proportionately with output. On the other hand, if , implying decreasing returns to scale, the exponent on in will be greater than 1. This means costs rise more than proportionately with output. Finally, if , implying constant returns to scale, the exponent on in will be 1, just as seen in part a.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

64


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

This means costs rise in exact proportion with output and the cost function will be a line through the origin as in Figure 8.3(a). c.

⁄(

)

⁄(

)

⁄(

)

⁄(

) ⁄(

. Substituting, ⁄( )

) ⁄(

⁄( )

) ⁄(

)

If (implying we have increasing returns to scale), then (implying we have economies of scale). If (implying we have decreasing returns to scale), then (implying we have diseconomies of scale). If (implying we have constant returns to scale), then (implying we have neutral economies of scale, on the boundary between economies and diseconomies of scale). This proves that returns to scale concepts have a one-to-one relationship with corresponding economies of scale concepts for Cobb-Douglas production functions.

8.9

d.

The greater is either one of the exponents the greater will be the exponent for that input’s unit cost in the total cost function.

e.

This function is linear in the logs of the various variables. It is therefore a good form for linear regression techniques. Note that the coefficient of (actually the reciprocal of this coefficient) reflects returns to scale and economies of scale, whereas the coefficients of v,w reflect the relative importance of the inputs.

f.

To find this you will have to understand why the Cobb-Douglas cost function is a special case of the Translog. The Translog adds interaction terms, which multiply the log of various combinations of variables. Setting the coefficients on these interactions to 0 allows one to recover the Cobb-Douglas

a.

Now K = L so q = 20 L. TC = vK + wL = 5K + 5L = 10L. So TC = 0.5q AC = TC/q = 0.50 MC = TC/q = 0.50. These costs are half what they were before.

8.10

b.

All costs will fall at the rate of r per year.

a.

Since w/v = 10/10 = 1, the expansion path would be unchanged. All costs would be twice what they were before: TC = 2q, AC = MC = 2.

b.

If w = 20, v = 5, w/v = 4 and the firm will operate on a new expansion path. Since cost minimization requires RTS = w/v = 4 = K/L, K = 4L and q = 10 KL = 20L = 5K. TC = 5K + 20L = q + q = 2q. Hence, AC = MC = 2. Multiplication of the wage by 4 only doubles costs because the firm substitutes K for L.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

65


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

c.

Now q  20L  20K : TC  vK  wL  v(q 20)  w(q 20)  (w  v)(q 20). With v = w = 10, TC = q. AC = MC = 1. The technical change has totally offset the price rise in the input prices.

d.

Now K = 4L,

q  20(2L)  40L  10K , TC  vK  wL  v(q 10)  w(q 40)  (q 40)(4v  w). With v = 5, w = 20, TC = q. AC = MC = 1. Again, technical progress has fully offset the rise in the price of L.

Chapter 9: Profit Maximization and Supply Purpose and Organization of the Chapter The chapter examines models of firms’ output decisions. Primary emphasis is placed on the consequences of the profit-maximization hypothesis. The chapter begins by analyzing the marginal decisions that accompany the profit-maximization hypothesis. It is at this point that marginal revenue is introduced and contrasted to the earlier concept of the price elasticity of demand. Notice that the elasticity of the market-demand function, eQ,P, is denoted differently from the elasticity of demand facing the individual firm, eq,P. This distinction is used to motivate the price-taking assumption (see also Application 9.4). The final sections of the chapter develop the short-run supply curve for a price-taking firm. This supply curve then provides the starting point for perfectly competitive price determination in Part 5. As in Chapter 8, the average variable cost curve is not shown explicitly, but its role in the shutdown decision is briefly discussed. Application 9.5 gives a nice illustration of how the shutdown decision is explained with the profit maximization model.

Lecture and Discussion Suggestions Some students have difficulty with the marginal revenue concept and therefore that concept should be featured in lectures on this chapter. One way to approach the subject is to repeat that the demand curve is an ―average revenue‖ curve with the MR curve ―marginal‖ to it. This approach creates an analogy between revenue and cost curves that may otherwise escape many students. Repeating the numerical example based on a linear demand curve can be especially helpful in this regard. Empirical material might again be appropriate for filling out lectures on this chapter. ―Incremental‖ thinking on the part of firm managers could be stressed together with some discussion of ―rules-of-thumb‖ (such as markup pricing or revenue maximization). Many firm decisions can be explained in this way as profit maximization when the firm faces uncertain-ties about the demand curve facing it. Discussions of the chapter material might also focus on profit maximization in the real world. Students might be asked to explain how some local business sets the prices of its products (or how much they choose to supply if they can be regarded as pure price takers). Trying to tie actual observed behavior to the theoretical models of the chapter can be quite challenging for students—especially since most actual behavior consists of price setting whereas the chapter is written mainly in terms of output choice. Often students will not see that there are simply alternative ways of looking at the same problem.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

66


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Cengage Supplements The following product-level supplements provide additional information that may help you in preparing your course. They are available in the Instructor Resource Center.  

Test Bank PowerPoint Slides

Chapter Objectives The following objectives are addressed in this chapter: 09.01

Understand why firms exist.

09.02 Explain how a firm decides how much to produce in order to maximize its profits. 09.03 Understand the relationship between marginal revenue and the demand curve. 09.04

Understand the supply behavior of a price taking firm.

[return to top]

What's New in This Chapter Because this brief chapter is very straight-forward, few changes have been made to its content. These include: 

Somewhat greater detail on how the marginal revenue curve is derived from a linear demand curve. The key step is to rewrite the demand function with price as a function of quantity. With this representation, the slope of the marginal revenue curve is indeed twice that of the original demand curve.

There is a bit more here on price-taking behavior (Application 9.4) – especially in relationship to relatively concentrated industries.

The data on oil drilling has been updated to capture more recent trends.

[return to top]

Additional Resources The treatment of firms here, although standard, is very simple. Because there have been many major advances in the theory of the firm in recent years, students might be encouraged to read more widely. Of course, a good starting point is the classic paper by Ronald Coase (see footnote 1). A very readable more recent summary is Oliver Williamson ―The Theory of the Firm as Governance Structure: From Choice to Contract‖ Journal of Economic Perspectives, Summer, 2002.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

67


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

The Summer 2014 issue of The Journal of Economic Perspectives has a number of articles about entrepreneurship. Some recent research on ―crowdfunding‖ is provided by Erin McGuire in ―How Crowdfunding can Open Doors for Female Entrepreneurs‖: https://vimeo.com/199841699. Economists who look in detail about price setting in various market settings often conclude that the situation is not so simple as suggested by the price-taking assumption. Katheryn Graddy’s study of the Fulton Fish Market provides some interesting examples: https://www.coreecon.org/the-economy/book/text/08.html#5hJF8zNJg5I.

[return to top]

Solutions to End of Chapter Problems Students have access to solutions for the odd-numbered problems as well as video problem walkthroughs for problems 1 and 7.

9.1

a.

Set P = MC, 20 = 0.2q + 10. q = 50.

b.

Maximum Profits = TR – TC = (50  20) – [0.1(50)2 + 10(50) + 50] = 1000 – 800 = 200.

c.

9.2

a.

This charge is a fixed cost of $100 per week. Will lower profits to $100, but will not affect output.

b.

Still maximize profits, so Beth earns $200, gets to keep $100.

c.

Now, MC = 0.2q + 12 since the fee increases Beth’s marginal costs. To maximize profits, Set P = MC, 20 = 0.2q + 12 q = 40 Father gets $80 and Beth’s profits are 800 – [160 + 400 + 50] – 80 = 110

d.

Net revenue per acre is now 18.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

68


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

P = MC yields 18 = 0.2q + 10 Again, q = 40. Change in MR here is same as change in MC in part c, so profitmaximizing output is the same in these two cases. 9.3 a.

Assume that the demand curve has the linear form P  c  dQ . Then marginal revenue is given by MR  c  2dQ . Solving for the Q-intercept of the demand curve yields Q  0  c  dQ*  Q*  c d . Making the same calculations for MR yields: MR  0  c  2dQ**  Q**  c 2d as was to be shown.

b. Total spending is maximized when MR = 0. c. If demand were inelastic raising price would increase spending, if demand were elastic lowering price would increase spending. Neither of these can happen because total spending is at a maximum. d. First, solve for P : P  48  Q 2  MR  48  Q. If P  0 , Q*  96. MR = 0 when Q**  48. With Q  48, P  24 and total spending is $1,152. This is the maximum spending with this demand curve. 9.4

a.

The graph shows that the demand curve has a convex shape, whereas the MC curve is linear.

Here the demand curve has a constant elasticity of –2. Hence 1 P MR  P (1  )  e 2

This is also shown in the graph. For profit-maximization need to show MR in terms of q, not P.

P

16 q

so MR 

8 q

Setting MR = MC yields

8 q  or q 1000

3

q 2  8000 so q  400 P 

16  0.8 20

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

69


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Here MC = P/2 = 0.40. 9.5

a.

Let AC  MC  c and suppose demand is given by Q  a  bP. The firm should now charge P  c and total quantity sold will be Q  a  bc .

b.

Since MR 

c.

This is the same analysis as in case a except now it must be the case that

a 2Q a a , MR = 0 when Q  hence, P  .  b b 2 2b

PQ  cQ  0.01PQ. Hence d.

Pc  0.01. P

It should sell just one unit. Price would be

a  1  bc . b

9.6

a 1 and unit profits would be b

e.

The solution in part a is the competitive solution and profits will be zero. Profits are not maximized in part b because MR  0  c . In part c profits are also probably not at a maximum because MR  c . In part d profits are clearly not at a maximum because the firm could profitably produce a second unit at a lower perunit profit.

a.

P  SMC  1  0.2q

b.

Variable costs are q  0.1q2

q  5P  5

Average variable costs are 1  0.1q . Hence SMC is always greater than average variable cost. There is no shutdown price.

9.7

10  1  0.1q . Hence this is the minimum for SAC. q

c.

At q = 10 SMC  1  0.2q  3 

d.

Yes, any value for P of less than 3 will cause price to fall short of SAC.

e.

No. With P = 2 the firm will produce q = 5. Average variable cost will be 1.5. Hence, price will exceed average variable cost. Total revenues will be 10 and total variable costs will be 7.5. Hence the firm will cover all its variable costs and have 2.5 to contribute to fixed costs (which here are 10).

a.

Beth’s supply function is q = 5P – 50. If P = 15, q = 25. If P = 25, q = 75.

b.

When P = 15,  = 15 × 25 – 362.5 = 375 – 362.5 = 12.5. When P = 25,  = 25 × 75 – 1,362.5 = 1,875 – 1,362.5 = 512.5. Average  = (512.5 + 12.5) ÷ 2 = 262.5.

c.

If P = 20, q = 50, = 1,000 – 800 = 200. The father’s deal makes Beth worse off.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

70


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

d.

Since high profits are associated with the high P, q combination, it’s more profitable to let price fluctuate. 9.8

a.

With a flat grant of $200 per week, Beth will end up with a total of $400 per week. The grant itself will not change the profit maximizing choice from Problem 9.1. With a subsidy of $4 per acre, net price rises to $24 per acre. Now profit maximization requires P = MC or 24 = 0.2q + 10. Hence, 0.2q = 14, q = 70. TR = 24  70 = 1,680 TC = 490 + 700 + 50 = 1,240 Total profits = 1,680 – 1,240 = 440, an improvement over the lump sum grant. The problem assumes that the MC curve correctly reflects Beth’s attitudes toward mowing 70 rather than 50 acres.

9.9

b.

With q = 70, at $4 per acre, the subsidy will cost the government $280 per week. Notice this is larger than would have been predicted if Beth’s output were assumed to be unchanged.

a.

STC = vK + wL = 10  100 + wL = 1,000 + 5L but q = 10 L so L =

q2 . 100 2

Hence, STC = 1,000 + q /20. b.

Use P = MC. 20 = 0.1q

so q = 200.

2

L = q /100 so L = 400. c.

If P = 15, P = MC implies 15 = 0.1q or q = 150, L = 225.

d.

Cost will be 175 to reduce L from 400 to 225. With q = 150, 2 Profits = TR – TC = 15(150) – (1,000 + 0.05q ) = 2,250 – (1,000 + 1,125) = 125. After paying severance cost of 175 the firm will incur a loss of 50. Note that if the firm continues to hire 400 workers it will have no severance costs and profits of TR 2 – TC = 15(200) – (1,000 + 0.05(200) = 3,000 – (1,000 + 2,000) = 0, which is

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

71


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

better than in part d. An output level of 180 (L = 324) would yield an overall profit for the firm. 9.10

a.

Using the profit-maximizing condition that P = MC yields q = 5 on Wednesdays and q = 10 on Saturdays.

b.

On Wednesdays profits are ( P  AC )q  (10  7)  5  15. On Saturdays profits are

(20  11) 10  90. c.

With P = 15, She should produce 7.5 each day. Profits on each day will be (15  8.33)  7.5  50.03 Hence, weekly profits will be about 100. With the variable price policy profits are 105. Hence she should not join.

d.

The claim makes no sense. Although the prices do fluctuate, those fluctuations add no uncertainty to Abby’s wealth because they are fully anticipated.

Chapter 10: Perfect Competition in a Single Market Purpose and Organization of the Chapter This chapter develops the familiar ―Marshall Cross‖ analysis of perfectly competitive pricing. By assuming that each firm takes market price as given, the short-run market supply curve is shown to be the horizontal sum of each firm’s short-run marginal cost curve. This market supply curve then interacts with market demand to determine equilibrium price and quantity in the short run. Long-run supply responses in perfectly competitive markets are the primary focus of this chapter. Emphasis is placed on the free entry assumption and on the way in which such entry will assure that economic profits are forced to zero. Considerable care is taken to develop longrun supply curves in the proper way. By focusing on average (rather than marginal) cost in the long run, the shape of the curve is shown to depend on the effect of entry on input costs. This in turn implies that the shape of long-run supply curves depends ultimately on the shape of the supply curves for factor inputs. That observation provides the basis for the rather extended discussion of producer surplus later in the chapter. The chapter also provides a number of illustrations of how the competitive model can be used. Consumer and producer surplus measures are used extensively to determine the welfare consequences of various actions. Special attention is devoted to the notion of producer surplus in the long run since there is considerable confusion about this concept. Because long-run competitive equilibrium always involves zero economic profits, producer surplus is not measured by profits in the long run (unlike the short run where producer surplus is the sum of short-run profits and fixed costs). Instead, changes in producer surplus arise because of changes in the rents earned by inputs to the market. This point is made using a simple model of Ricardian rent.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

72


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

The basic insight is then used throughout the applications to describe which economic actors actually experience the welfare gains or losses being experienced by ―producers.‖

Lecture and Discussion Suggestions The derivation of short-run supply curves in the chapter is relatively simple and may be familiar to students from previous economic courses. For that reason, emphasis in lecture might more properly be put on long-run analysis—a topic not usually well-covered in introductory economics. For some reason, students seem to have trouble seeing why entry or exit shifts the short run supply curve, so perhaps a numerical example would be helpful to get the dynamics right. The notion that shifting of the short run supply curve ceases when price reaches minimum average cost (because the number of firms has reached the optimal level) need not be shown explicitly as that does involve a number of diagrams. Instead, one can show a single, long-term equilibrium and then describe comparative statics analysis solely in a short-run context. There is, of course, no shortage of discussion material for this chapter. We believe it is especially important to stress the long-run producer surplus notion because this is the correct way to integrate input and output markets in a partial equilibrium context. The tax-incidence issue is especially revealing in this regard—it may come as a shock to students that only people pay taxes.

Cengage Supplements The following product-level supplements provide additional information that may help you in preparing your course. They are available in the Instructor Resource Center.  

Test Bank PowerPoint Slides

Chapter Objectives The following objectives are addressed in this chapter: 10.01

Understand how prices are determined in a perfectly competitive market.

10.02

Explain what happens in a competitive market as the result of the entry and exit of firms.

10.03

Show how perfect competition leads to an efficient allocation of resources.

10.04

Explain why the party who is legally obligated to pay a tax is not necessarily the one who bears the burden of the tax.

10.05

Understand the welfare consequences of international trade.

[return to top]

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

73


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

What's New in This Chapter Here are a few of the changes to this chapter in the current edition: 

The most important change in this chapter is the transition to multi-colored graphs. This permits both a clear distinction between supply and demand concepts, but also shading is used to indicate initial (shaded color) versus final (full color) positions. The more complicated graphs (such as Figure 10.8) in the chapter are now a bit easier for students to understand.

The chapter also includes more on expressing supply and demand models with algebra. This material should be helpful to students in attempting the end-of-chapter problems (which in some cases are quite complicated—see Problem 10.9, for example).

Several of the applications have significant updates including those on Network Externalities (Application 10.3), Internet Commerce (Application 10.4) and Dumping Claims (Application 10.6).

[return to top]

Additional Resources The possible taxation of Ricardian Rents is an important topic in applied economics, largely dating from Henry George’s advocacy of the ―single tax‖ on them. For a nice summary see: https://onlinelibrary.wiley.com/doi/full/10.1111/joes.12340. It is somewhat depressing to realize that our application on steel tariffs (Application 10.6) has been revised many times to account for new tariffs throughout this text’s thirteen editions. A good examination of the effects of the most recent round of such tariffs is provided by Amiti, Redding, and Weinstein ―The Impact of the 2018 Tariffs on Prices and Welfare‖, Journal of Economic Perspectives, Fall, 2019. A particularly nice illustration of the need to consider the actual incidence of taxes concerns the effect of a tax imposed of American-built yachts in the early 1990s. It was ultimately American employees of yacht-building firms who paid the tax. See: https://www.washingtonpost.com/archive/business/1993/07/16/how-to-sink-anindustry-and-not-soak-the-rich/08ea5310-4a4b-4674-ab88-fad8c42cf55b/.

[return to top]

Solutions to End of Chapter Problems Students have access to solutions for the odd numbered problems as well as video problem walkthroughs for problems 1 and 9.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

74


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

10.1

a.

Set supply equal to demand to find equilibrium price: QS = 1,000 = QD = 1600 – 600P. 1,000 = 1,600 – 600P. 600 = 600P P = 1/pound b.

QS = 400 = 1,600 – 600P. 600P = 1,200. P = 2/pound

c.

QS = 1,000 = 2,200 – 600P. 1,200 = 600P. P = 2/pound QS = 400 = 2,200 – 600P. 600P = 1,800. P = 3/pound

d.

QS = 0 = –1,000 + 2,000P 1,000 = 2,000P P = 1/2 per lb. Price will have to be greater than 1/2 per lb. for flounder to be supplied in Cape May.

e.

At equilibrium, QD = QS: –1,000 + 2,000P = 1,600 – 600P 2,600P = 2,600 P = 1/lb. Equilibrium occurs at P = 1/lb.

f.

–1,000 + 2,000P = 2,200 – 600P 2,600P = 3,200

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

75


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

P = 32/26 = 16/13 per lb. g.

Price will rise by less because in c there can be a supply response. The increased demand does not lead only to a price increase, but also an increase in the quantity supplied. The graph shows these various equilibria.

10.2

a.

Supply = 100,000. In equilibrium, 100, 000  QS  QD  160, 000  10, 000 P or P = 6.

b.

For any one firm, quantity supplied by other firms is fixed at 99,900. Demand Curve is

qd  160, 000  10, 000 P  99,900  60,100  10, 000 P . If quantity supplied is 0, qs  0  60,100  10, 000 P  P  6.01 . If quantity supplied is 200, qs  200  60,100  10, 000 P  P  5.99. Elasticity = Slope of demand × P/Q for market.

eQ, P  10,000 

6  0.6 100,000

For a single firm, demand is much more elastic:

eq , P  10,000 

6  600 100

A change in quantity supplied by one firm does not affect price very much (as shown in the numerical example in part b). Now qi  200  50 P c.

If there are 1,000 firms

QS  1, 000qi  200, 000  50, 000 P . For equilibrium –200,000 + 50,000P = 160,000 – 10,000P

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

76


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

60,000P = 360,000 or

P=6

For any one firm,

qd  160, 000  10, 000 P  (199,800  49,950P) = 359,800 – 59,950P. If qs  0

P

359,800  6.002 59,950

If qs  200

P

359,600  5.998 59,950

Demand curve facing the firm is even more elastic than in the fixed supply case because of the potential supply response by other firms. 10.3

3

2

STC = 1/300q + 0.2q + 4q + 10 a.

2

MC = 0.01q + 0.4q + 4

Short run profit maximization requires P = SMC. 2

P = 0.01q + 0.4q + 4 2

2

100P = q + 40q + 400 = (q + 20) = 100P q + 20 = 10 P q = 10 P – 20 b.

Industry with 100 firms has supply curve of Q = 1,000

c.

P – 2,000

QD = –200P + 8,000 For equilibrium, set demand = supply: –200P + 8,000 = 1000 P – 2,000 1,000 P + 200P = 10,000 5 P + P = 50,

d.

P = 25,

Q = 3,000

–200P + 11,200 = 1,000 P – 2,000  P  36 .

Q  4, 000 each firm produces q  40 . For each firm, total revenue is 1,440. Short-run total costs are 703. Profits are 737.

10.4

a.

2

If w = 10, STC = q + 10q. SMC = 2q + 10 = P. Hence, q = P/2 – 5.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

77


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Industry Supply: 1,000

Q   q  500 P  5,000 1

at P = 20, Q = 5,000; at P = 21, Q = 5,500. b.

Here MC = 2q + 0.002Q. Set = P for profit maximization. Hence, q = P/2 – 0.001Q. Supply for industry as a whole is 1,000

Q   q  500 P  Q 1

Therefore, Q = 250P. P = 20, Q = 5,000. P = 21, Q = 5,250. Supply is more steeply sloped in this case of cost interactions—increasing production bids up the wages of diamond cutters. 10.5

a.

In long-run equilibrium, AC = P and MC = P, so AC = MC. 100  0.02q  1 or q 2  10,000 q q  100 gallons.

0.01q  1 

b.

In the long-run, P = MC, implying P = $1. QD = 2,500,000 – 500,000(1) = 2,000,000 gallons. The market supplies 2,000,000 gallons so

2,000,000 gallons  20,000 stations 100 gallons per station c.

In the long run, P = $1 still since the AC curve has not changed. QD = 2,000,000 – 1,000,000(1) = 1,000,000 gallons Now there are only 10,000 stations.

10.6

a.

LR supply horizontal at P = MC = AC = 10.

b.

Q* = 1,500 – 50P* = 1,000. Each firm produces q* = 20,  = 0. There are 50 firms.

c.

SMC = q – 10, AC = 0.5q – 10 + 200/q. AC = min when AC = MC. 0.5q = 200/q, q = 20.

d.

P = MC = q – 10. So q = P + 10. For the entire industry 50

Q   q  50P  500 1

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

78


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

e.

Q = 2,000 – 50P. If Q = 1,000, P = 20. Each firm produces q = 20,  = 400 – 200 = 200.

f.

50P + 500 = 2,000 – 50P

P = 15, Q = 1,250.

Each firm produces q = 25,  – 25(15 – AC)  = 25 (15 – 10.5) = 112.5

10.7

g.

P = 10 again, Q = 1,500, 75 firms produce 20 each.

  = 0.

a.

With Q = 400, demand curve yields 400 = 1000 – 5P or P = 120. For supply, 400 = 4P – 80 or P = 120. Hence, P is an equilibrium price. Total spending on broccoli is 400  120 = 48,000. On the demand curve when Q = 0, P = 200. Hence, area of the consumer surplus triangle is 0.5(200 – 120)(400) = 16,000. On the supply curve, P = 20 when Q = 0. Producer surplus is then 0.5(120 – 20)(400) = 20,000.

b.

With Q = 300, the total loss of surplus would be given by the area of the triangle between the demand and supply curves which is 0.5(140 – 95)(100) = 2,250.

c.

With P = 140, consumer surplus is 0.5(200 – 140)(300) = 9,000. Producer surplus is 0.5(95 – 20)(300) + 45(300) = 24,750. Consumers lose 7,000, producers gain 4,750; net loss is 2,250. With P = 95, consumer surplus is 0.5(200 – 140)(300) + 45(300) = 22,500. Producer surplus is 0.5(95 – 20)(300) = 11,250. Consumers gain 6,500, producers lose 8,750; again, net loss is 2,250.

d.

With Q = 450, demand price would be 110, supply price is 132.50. Total loss of surplus is 0.5(132.5 – 110)(5) = 562.50. Net loss is shared depending where price falls between 110 and 132.5.

e.

10.8

a.

For supply, set P = SMC.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

79


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

P = q + 10 q = P – 10 100 firms in industry, so industry supply is Q = 100q = 100P – 1,000. b.

For equilibrium, 1,100 – 50P = 100P – 1,000 P = 14, Q = 400, P = TR – TC = 4(14) – (8 + 40 + 5) = 3.

c.

The graph shows supply-demand equilibrium. Consumer surplus is 0.5(22 – 14)(400) = 1,600. Producer surplus is 0.5 (14 – 10)(400) = 800. Total surplus is 2,400.

d.

Since profits for a single firm are 3, total industry profits are 300. Short-run fixed costs are 5 for each firm, or a total of 500. Hence, short-run profits plus fixed costs is 800, which equals producer surplus.

e.

New equilibrium is found as: Q  100 P  1,000  1,100  50( P  3) P  13 P  3  16 Q  300

Total taxes are 900 f.

Consumers pay (16 – 14)(300) = 600 Producers pay (14 – 13)(300) = 300

g.

Producer surplus is now 0.5(13 – 10)(300) = 450 -- a decline of 350 from problem 11.2c. Now profits for each firm are Pq  STC (3)  39  39.5  0.5

Total profits are –50 – a decline from +300 in problem 11.2d. Hence, the decline in profits precisely matches the decline in producer surplus. Fixed costs (of 500) do not change throughout the problem. Calculated another way, short-run producer surplus is now profits (of –50) plus short-run fixed costs (of 500) for a total of 450. 10.9

a.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

80


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

b.

Since P = AC = 10 + r = 10 + 0.002Q, substitute this into demand: Q = 1,050 – 50P = 1,050 – 500 – 0.1Q or 1.1Q = 550, Q = 500. Since each firm produces 5 bobbleheads, there will be 100 firms. Royalty is r = 0.002(500) = 1 so P = 11.

c.

With Q = 1,600 – 50P, same substitution gives Q = 1,600 – 500 – 0.1Q or

1.1Q = 1,100, Q = 1,000.

So now there are 200 firms and r = 0.002(1,000) = 2 so P = 12. d.

Producer surplus when P = 11 is 0.5(11 – 10)(500) = 250. When P = 12, it is 0.5(12 – 10)(1,000) = 1,000. e.

Royalties when Q = 500 are 500. Increment when Q rises from 500 to 1,000 is (2 – 1)(500) + 0.5(2 – 1)(1,000 – 500) = 500 + 250 = 750 which is precisely the increase in producer surplus in part d.

f.

With the tax demand is now Q = 1,050 – 50(P + 5.5). Since P = 10 + 0.002Q, this means Q = 1,050 – 500 – 0.1Q – 275 or 1.1Q = 825, Q = 750, P = 11.5.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

81


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Price to consumers is 17. g.

Total tax collections are 5.5(750) = 4,125. Consumers pay (17 – 12)(750) = 3,750 Producers pay (12 – 11.5)(750) = 375. Consumer surplus is now 0.5(32 – 17)(750) = 5,625 whereas previously it was 0.5(32 – 12)(1,000) = 10,000, so the loss is 4,375: 3,750 of tax revenue and 625 from foregone transactions. Producer surplus was 1,000; now it is 0.5(11.5 – 10)(750) = 562.5 a loss of 437.5.

h.

All of the lost producer surplus is a loss of royalties. Now r = 0.002(750) = 1.5 whereas previously r = 2. Loss is (2 – 1.5)(750) + 0.5(2 – 1.5)(250) = 375 + 62.5 = 437.5.

10.10

a.

Set quantity supplied equal to quantity demanded 150P = 5,000 – 100P; P = 20, Q = 3,000.

b.

P will fall to 10. QD = 4,000, QS = 1,500. 2,500 radios will be imported.

c.

Price would now rise to 15. QD = 3,500, QS = 2,250. Imports are now 1,250. Tariff revenue is 5(1,250) = 6,250. With free trade, consumer surplus is 0.5(50 – 10)(4,000) = 80,000. Domestic producer surplus is 0.5(10)(1,500) = 7,500. With the tariff, consumer surplus is 0.5(50 – 15)(3,500) = 61,250, a loss of 18,750. Producer surplus is now 0.5(15)(2,250) = 16,875, a gain of 9,375. Deadweight loss is 18,750 – 6,250 – 9,375 = 3,125, as can be found by measuring the triangles.

d.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

82


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

[return to top]

Chapter 11: General Equilibrium and Welfare Purpose and Organization of the Chapter This chapter provides a very elementary introduction to general equilibrium theory. It begins by showing why taking a general equilibrium approach may be necessary to address some important economic questions and then proceeds to build a simple model of two markets. That model (drawn primarily from the graphical approach to international trade theory) generalizes ―supply‖ by using the production possibility frontier and ―demand‖ by using a typical person’s indifference curve. An advantage of this approach is to stress that the economic ―problem‖ is how to make the best (utility-maximizing) use of scarce resources. The middle portion of the chapter is devoted to showing the ―first theorem of welfare economics‖ (that perfectly competitive prices, under certain circumstances, yield economic efficiency). Again, this is done using the production possibility frontier and indifference curves to show how the operations of markets cause the economy to settle at an efficient allocation. Reasons why the first theorem may fail are discussed in the third section of the chapter. Subjects given very brief treatment include: (1) Imperfect competition; (2) Externalities; (3) Public goods; and (4) Imperfect in-formation. Each of these topics is covered in considerable detail in later chapters. The discussion here also includes a brief discussion of equity and of how goals of equity and efficiency may sometime (but by no means al-ways) be in conflict. The Edgeworth Box Diagram is the primary tool used for this purpose. The chapter concludes with a brief discussion of how money enters into general equilibrium models. The main goals here are: (1) to introduce the ―classical dichotomy‖ between monetary and real sectors; and (2) to illustrate the notion of fiat money and why this innovation has important eco-nomic implications.

Lecture and Discussion Suggestions Repeating the development of the general equilibrium model in this chapter in lecture would probably be quite dull. Hence, it may better to assume that students have understood the development in the text and just use the model to illustrate some results. One approach that seems to work well is to use separate supply and demand curves for goods X and Y together with the general equilibrium model to show how both approaches to equilibrium are getting at the same sort of thing. Reasons for the superiority of general equilibrium should become readily apparent in this comparison. Having an operational, simple GE model can also provide students with a lot of insights about how these models work in practice. The model described in W. Nicholson and F. Westhoff, ―General Equilibrium Models: Improving the Microeconomics Classroom‖ (Journal of Economic Education, Summer, 2009, pages 297-314) provides a nice such introduction. But there are many other possibilities that could be used.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

83


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Discussions of general equilibrium might focus on ―what more did you learn by using these models?‖ For example, students may find that tax incidence questions are much more complicated than they at first thought. Especially interesting are discussions of the role of capital taxation and how theoretical insights might shed light on real world issues about, say, the incidence of the corporate tax. Use of general equilibrium models to look at trade issues also provides a number of good discussion questions. For example, students may have rather simplistic views about how the NAFTA may have affected the welfare of low-income workers and it may be useful to show them how complex answering this question actually is.

Cengage Supplements The following product-level supplements provide additional information that may help you in preparing your course. They are available in the Instructor Resource Center.  

Test Bank PowerPoint Slides

Chapter Objectives The following objectives are addressed in this chapter: 11.01

Understand how disturbances in one market affect equilibrium prices in related markets.

11.02

Explain why a system of perfectly competitive markets leads to economic efficiency – the best use of societies resources.

11.03

Know some of the main reasons why competitive markets fail to achieve economic efficiency.

11.04

Understand how trade in competitive markets will lead to an outcome where no one person can be made better off without making someone worse off.

11.05

Understand the tradeoffs between efficiency and equity.

11.06

Explain how including money in our model is necessary to show how nominal prices are determined, but that money may have no effect on relative prices.

[return to top]

What's New in This Chapter There are only a few modest changes to this chapter from the previous edition. These include the following. 

An extended discussion of the Edgeworth Box Diagram shows how it can be used to illustrate efficiency and equity issues. The graphical analysis using the Edgeworth Box is significantly enhanced through the use of multiple colors.

A new application (Application 11.3) describes some of the general equilibrium attempts to model the COVID-19 lockdowns. Since this analysis is in its infancy, students might be asked how they would proceed to develop their own lockdown model. A good starting

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

84


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

point would be to ask whether this should be treated as affecting demand or supply and why getting the approach right matters. 

The two-page application on Trade Agreements (Application 11.2) has been expanded to include some recent material on the USMCA.

Although modeling climate change is not treated until Chapter 17, students may notice that this problem is also ideal for general equilibrium modeling. They might therefore be directed to the revised Application 17.4 on this topic.

[return to top]

Additional Resources Understanding how modern general equilibrium models work is a complex topic. The Paper by Nicholson and Westhoff in the Journal of Economic Education (see above) may be helpful in understanding. The Wikipedia entry on Computable General Equilibrium Models also offers a nice intuitive discussion. The introduction of general equilibrium modeling provides a good opportunity for microeconomists to needle their macroeconomic colleagues. A nice litmus test is how economists feel about the current wave of Dynamic Stochastic General Equilibrium (DSGE) models. Sbordone, Tambalotti, Rao, and Walsh have a useful review of the use of such models in macroeconomics in the Federal Reserve Bank of New York Economic Policy Review October, 2010. This might be contrasted to Robert Solow’s famous negative review of the models before Congress on July 10, 2010. The Environmental Protection Agency provides an insightful analysis on how computable general equilibrium models can be used for regulatory purposes: https://www.epa.gov/environmental-economics/cge-modeling-regulatoryanalysis#:~:text=These%20analyses%20quantify%20the%20expected,to%20occ ur%20absent%20the%20regulation. [return to top]

Solutions to End of Chapter Problems Students have access to solutions for the odd numbered problems as well as video problem walkthroughs for problems 1 and 5.

11.1

a.

The production possibility frontier for M and C is shown as:

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

85


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

b.

If people want M = ½ C and technology requires C + 2M = 600, then C + 2(1/2C) = 600. 2C = 600 or C = 300. M = 150.

c.

For efficiency RPT = MRS = 1/2, so

RPT  MRS  11.2

PC 1  . PM 2

a.

See graph.

b.

See graph.

c.

The production possibility frontier is the set of food and cloth outputs that satisfy both constraints (see graph).

d.

The frontier is concave because the two goods use differing factor proportions. The slope changes as a different input becomes the binding constraint.

e.

The constraints intersect at F = 50. For F < 50 the slope of the frontier is –1. P Hence, in this range, F  1 . For 50 < F < 75 the slope of the frontier is –2 PC P (because land is the binding constraint). In this range therefore F  2 . PC

f.

With these preferences,

g.

Any price ratio between 1.0 and 2.0 will cause production to occur at the kink in the frontier.

PF 5  . PC 4

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

86


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

11.3

h.

This capital constraint lies always outside the previous production possibility frontier. It will not therefore affect any of the calculations earlier in this problem.

a.

The frontier is a quarter ellipse:

b.

If Y  2 X , X 2  2(2 X ) 2  900 2

9X = 900; X = 10, Y = 20. This point is shown on the frontier in part a. c.

If X = 9 on the production possibility frontier,

Y  819/ 2  20.24 If X = 11, Y  779/ 2  19.75 Hence, RPT = 0.49/2 = 0.245 . This is the ratio of prices that will cause production to occur at X = 10, Y = 20. d. 11.4

See graph in part a. 2

2

Since LF  LC  8 . the production possibility frontier is F + C = 8 Given H = 16, U = 4F¼ C¼ and we know that optimality will require C = F since the goods enter both the utility function and the production possibility frontier symmetrically. 2 Since C = F, have 2C = 8 or C = F = 2. Utility = 4 2.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

87


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

11.5

a.

Given the production conditions, the production possibility frontier will be a straight line with slope - 3/2. Hence the price ratio in this economy must be PX 3 X Y   20 .  . The equation for the frontier is 2 3 PY 2

b.

Using the hint, X S 

3 PX

XJ 

5 PX

XT 

8 PX

12 . Substituting these into the equation for the frontier and using PY 2P 4 4 10 1 1    20 PX  ; PY  . Notice how the fact that PY  X yields 3 Px PY PX 2 3 setting the wage here also sets the absolute price level.

Similarly YT 

11.6

c.

With these prices, total demand for X is 16, total demand for Y is 36. Hence 12 hours of labor must be devoted to Y production, 8 hours to X.

a.

For region A the production possibility frontier is X A2  YA2  100 . For region B it is X B2  YB2  25 . Hence the frontiers are concentric circles with radius 10 for A and 5 for B.

b.

Production in both regions must have the same slope of the production possibility frontier. In this case that means that the ratio X/Y must be the same in both regions – production must take place along a ray through the origin.

c.

The geometry of this situation suggests that for efficiency X A  2 X B

YA  2YB .

Hence X T  3 X B

YT  3YB and the frontier is given by X  Y  9( X  YB2 )  225 . If X T  12 YT  9 . 2 T

11.7

2 T

2 B

a.

U1  10 U 2  5 .

b.

F1 

c.

The allocation in part a achieves this result -- F1  F2  100  U1  10 U 2  5 .

d.

A natural suggestion would be to maximize the sum of utilities. This would require 1 1 MU 2  that marginal utilities be equal. Because MU1  equality of 2 F1 4 F2

F2 which implies F1  40 F2  160 . 4

marginal utilities requires F1  4F2 F1  160; F2  40 -- a rather unequal distribution. Still the sum of utilities is 15.8 – the largest possible. With an equal allocation the sum of utilities, for example, is 15.0. 11.8

a.

The total value of transactions is 20w. So, money supply = 60 = money demand = 5w. So w = 12 (earlier we assumed w  10 ) With this wage, the nominal prices 1 12 1 12 should be changed as:. PX    0.6 PY    0.4 . 2 10 3 10

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

88


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

b.

11.9

If the money supply increases to 90, all wages and prices increase by 50 percent: w  18, PX  0.9, PY  0.6 . Relative prices and the overall allocation of resources remain the same. Yes, this economy exhibits the classical dichotomy.

a-c. See Graph

d.

As before, efficient points are the tangencies of the isoquants.

e.

The production possibility frontier shows the maximum amount of Y that can be produced for any fixed amount of X. Any point off the contract curve has the property that Y can be increased even if X is held constant.

f. (i)

The production possibility frontier is a single point where X gets all labor input, Y gets all capital input.

(ii)

The frontier would be a straight line

(iii) Again, the frontier would be a straight line. Only with differing factor intensities would the frontier have a concave shape. (iv)

The frontier would be convex.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

89


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

11.10

a.

The preferences of Smith and Jones are shown in the figure. The only exchange ratio that can prevail is set by Jones’ preferences – 1C must trade for 0.75H. On the other hand, all efficient allocations must lie along the main diagonal of the box where, because of Smith’s preferences, C = 2H.

b.

This is an equilibrium – the allocation lies on the contract curve and any trade would make at least one person worse off.

c.

Now the initial position is off the contract curve. Smith has 20―extra‖ H. If Jones gets all the gains from trade because Smith gives these to him/her, utility will increase from U J  4(40)  3(120)  520 to U J  4(60)  3(120)  600 . If Smith gets all the gains from trade, the new equilibrium requires 4 H  3C  520 and C  2 H . Hence, the equilibrium requires Jones to get H = 52, C = 104. Smith gets H = 48, C = 96 and is much better off than at the initial allocation. Smith may be able to enforce this equilibrium or, if he/she is especially strong may in fact take everything.

[return to top]

Chapter 12: Monopoly Purpose and Organization of the Chapter This chapter surveys the traditional theory of monopoly behavior. The implications of monopolists’ market power for the allocation of resources are stressed: the deadweight loss of reduced output and the redistribution of surplus from consumers to the firm. Two extensions of monopoly theory are analyzed in the chapter: price discrimination and regulation of monopoly. The price discrimination section distinguishes between two forms: discriminating by separating different markets and by using a nonlinear pricing scheme. (We © 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

90


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

avoid the terminology of ―degrees‖ of price discrimination because that seems to be more confusing than clarifying. In this terminology, discriminating by separating markets is called third-degree price discrimination and by using nonlinear prices is called second-degree price discrimination. Perfect price discrimination is sometimes called first-degree price discrimination.) The analysis of price discrimination is initially motivated by showing that the traditional monopoly solution still leaves unexploited gains for the monopoly. Price discrimination schemes are intended to convert these opportunities into profits. Chapter 16 on asymmetric information goes into the problem of nonlinear pricing in much more detail (indeed, this is the motivating ap-plication in the adverse selection section there), but the basic ideas are covered in Chapter 12. The discussion of regulation has as its principal concern the problems raised by marginal cost pricing for a firm with declining average costs. The ―natural monopoly dilemma‖ is illustrated and a few solutions are examined.

Lecture and Discussion Suggestions A theoretical lecture on this chapter should make clear why monopolies and perfectly competitive industries behave differently. One way to make that distinction is to analyze Figure 12.3 more thoroughly in lecture. Notice that the point of comparison here is with a perfectly competitive industry with an infinitely elastic long-run supply curve. We believe that is a more correct comparison than, say, to a single competitive firm because it permits illustrating welfare consequences for the market as a whole. Another way to show the difference between monopoly and perfect competition is to contrast the comparative statics analysis of the response to a shift in demand. Problems 12.3 and 12.4 provide illustrations (perhaps to the point of tedium) of such shifts that show why the monopoly situation is more complex than the price-taker case. Perhaps the most complex topic in the chapter is pricing discrimination, nonlinear pricing in particular. One possibility for covering this topic is to organize the lecture around how Disneyland sets its prices. Fairly detailed and up-to-date facts on its pricing strategies are provided in the revised Ap-plication 12.4: A Mickey Mouse Monopoly. What is nice about this application is that this one single firm uses such a variety of different schemes. The theory of regulation offers a number of empirical topics that can provide interesting material for both lectures and discussion based on this chapter. The natural monopoly pricing dilemma can be succinctly covered by reviewing Figure 12.8. Once this is done, the instructor can raise the discussion point that industries that were once natural monopolies may become more competitive when the technology changes, using the example of telephones (the technology change being the introduction of mobile phones and internet telephony) and television (with cable still being regulated, but facing increasing competition from satellite and internet technologies). Still, regulation is a perennial policy debate, the currently ―hot‖ areas being regulation of access prices charged by internet providers to con-tent providers and of interchange fees by credit- and debit-card associations. Discussions of the politics of regulatory activity can also be interesting both in terms of economic effects (of, say, regulatory uncertainty or regulatory lag) and in terms of issues in public choice theory (e.g., regulatory capture).

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

91


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Cengage Supplements The following product-level supplements provide additional information that may help you in preparing your course. They are available in the Instructor Resource Center.  

Test Bank PowerPoint Slides

Chapter Objectives The following objectives are addressed in this chapter: 12.01

Understand the reasons monopoly markets exist.

12.02

Illustrate how a monopolist sets its price and output.

12.03

Explain some of the problems with monopoly.

12.04

Understand alternative pricing strategies a monopolist may use.

12.05

Explain three different ways governments use to regulate a natural monopoly’s prices.

[return to top]

What's New in This Chapter The chapter is similar to that in the previous edition. 

Facts in Applications 12.1 through 12.5 have been updated.

Application 11.6 on telephone pricing has been deleted. Most current students use smartphones rather than fixed lines, so may not find the historical example of regulated telephone prices that compelling.

The product in the extended numerical example in Section 12-3c has been changed from the antiquated CD example to mugs.

[return to top]

Additional Resources The American Economic Association posts videos of a diverse set of academic economists discussing their relevant research on the website http://diversifyingecon.org/index.php/Videos_on_economists_and_their_research.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

92


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Two videos relate to the role of intellectual-property protection in spurring innovation, with the potential drawback that a temporary monopoly may be created. In the following link, Petra Moser talks about her research into the effect of copyright protection on the quality of Italian operas. https://www.youtube.com/watch?v=F_W1V673YJg&t=7s. Heidi Williams talks about medical innovations in the following link. https://www.macfound.org/fellows/class-of-2015/heidi-williams.

[return to top]

Solutions to End of Chapter Problems Students have access to solutions for the odd numbered problems as well as video problem walkthroughs for problems 1 and 6.

12.1

a.

P = 53 – Q. For maximum profits, set MR = MC: MR = 53 – 2Q = MC = 5. Q = 24, P = 29.  = TR – TC = 24  29 – 24  5 = 696  120 = 576. Consumer Surplus  0.5  (53  29)  24  288

b.

MC = P = 5, P = 5, Q = 48.

Consumer Surplus  0.5  (48)2  1,152

.

1,152 > Profits + consumer surplus = 576 + 288 = 864. Deadweight loss = 1,152 – 864 = 288. Also 1/2Q – P = 1/2(24)(24).

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

93


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

12.2

Market Demand Q = 70 – P, MR = 70 – 2Q. a.

AC = MC = 6. To maximize profits set MC = MR. = 70 – 2Q

6

2Q = 64

b.

Q

= 32

P

= 38

= TR – TC = (32)(38) – (32)(6) = 1,024 2

TC = .25Q – 5Q + 300, MC = .5Q – 5. Set MC = MR .5Q – 5 = 70 – 2Q 2.5Q

= 75

Q

= 30

P

= 40

= (30)(40) – [.25(30) – 5(30) + 300]

2

= 1,200 – 375 = 825. c.

3

2

TC = 0.01Q – Q + 45Q +100. 2

MC = 0.03Q – 2Q + 45. 2

Set MC = MR, 0.03Q – 2Q + 45 = 70 – 2Q 2

0.03Q = 25, implying Q =√

= 28.9

P = 41.1 (

)

[

]

. d.

The graph shows the solutions to parts a, b, and c. Notice only cost conditions vary among these three solutions.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

94


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

12.3

a.

AC = MC = 10, Q = 60 – P, MR = 60 – 2Q. For profit max., MC = MR. 10 = 60 – 2Q, 2Q = 50, Q = 25, P = 35.  = TR – TC = (25)(35) – (25)(10) = 625.

b.

AC = MC = 10, Q = 45 – .5P. MR = 90 – 4Q. For profit max., MC = MR, 10 = 90 – 4Q, 80 = 4Q, Q = 20, P = 50.  = (20)(50) – (20)(10) = 800.

c.

AC = MC = 10, Q = 100 – 2P, MR = 50 – Q. For profit max., MC = MR, 10 = 50 – Q, Q = 40, P = 30.  = (40)(30) – (40)(10) = 800.

d.

The supply curve for a monopoly is the single point on the demand curve that corresponds to the quantity for which MC = MR. Any attempt to connect equilibrium points (price-quantity points) on a series of market demand curves has little meaning and brings about a strange shape. One reason for this is that as the demand curve shifts, its elasticity (and its MR curve) often changes, bringing about widely varying price and quantity combinations.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

95


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

12.4

a.

Graph shows shifts in demand (and MR) for two types of shift.

b.

There is no supply curve for a monopoly, have to examine MR = MC intersection. In case 1, price rises; in case 2, it falls.

c.

This question can be addressed by using the relationship

P  1 MR  P 1    e  e MR P  

e

P . P  MR

Can use this to study the three cases.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

96


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

• Case 1 MC constant, so MR is constant. If –e falls, P – MR rises, so P rises. If –e constant, P – MR constant, so P is constant. If –e rises, P – MR falls, so P falls. • Case 2 MC falling, so MR falls as Q expands. If –e falls, P – MR rises, so P may rise or fall. If –e constant, P – MR constant, so P falls along with MR. If –e rises, P – MR falls, so P must fall. • Case 3 MC rising so MR must rise with increases in Q. If –e falls, P – MR rises, MR rises, so P rises. If –e constant, P – MR constant, so P must rise. If e rises, P – MR falls, so P may rise or fall. This shows P may change in a variety of ways in response to an increase in demand depending on how elasticity changes. 12.5

A multiplant monopolist will still produce where MR = MC and will equalize MC among factories.

MR  100  2(q1  q2 ) and MC1  MC2 q1  5  .5q2  5 or q1  0.5q2 MR  100  3q2  MC2  .5q2  5 so q2  30 q1  15 So, total Q is 45. 12.6

a.

First prove the hint:

In the graph Qmax is the quantity demanded when P = 0. Since MR = 0 at 1/2Qmax (there e = –1), it is clear that MR bisects the distance from the P axis to the demand curve. So, if Q* represents quantity demanded when P = MC, MR = MC at 1/2Q*.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

97


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Notice also that the profit-maximizing price is given by 0.5 (Pmax + MC) where Pmax is the price for which quantity demanded is zero. Note: The results of this hint are used to solve several problems in later chapters. b.

If Q1 = 55 – P, then since MC = 5, Q1 = 55 – 5 = 50

Q1 * = 25 2 At that output level, P1 = 30  = (P1 – 5)(Q1) = 25  25 = 625. If Q2 = 70 – 2P2, Q2* = 70 – 2(5) = 60 = 70 – 2(5) = 60

Q2*  30 2 Therefore, P2 = 20  = (20 – 5)  30 = 450. Total profits = 1 + 2 = 1,075. c.

This is a hard problem, so let’s work up to the solution in steps. Total profits across the two markets can be written ( ( (

) )( )(

(

) ) )

( ([

)( ]

) )(

[

])

The first two equations are self-explanatory. The next step substitutes in the demand curves in each market and the specific value of average cost. The crucial step is the last one, where has been substituted in square brackets. Here is where we are using the fact that the two prices can’t differ by more than the arbitrage cost. Using tedious algebra, you can expand the last equation and then combine terms and to finally show

This is a hump-shaped parabola. You can find its maximum in several ways. If you know calculus, you can use standard maximization procedures. If you don’t know calculus, you can use the standard formula for the vertex:

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

98


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Substituting into the relevant formulas gives , , , , , . When the arbitrage cost is $0 between the markets, the prices have to be the same. Substituting into the first set of equations gives (

)(

)

(

)(

)

Sparing you the gory details, one obtains , , , , , . Note that total profits fall as market separation decreases. 12.7

QD = 1,000 – 50P

MR = 20 – Q/25

MC = 10 under PC

MC = 12 under monopoly.

a.

Perfect competition: P = MC = 10 QD = 1,000 – 50(10) = 500 = QS Monopoly: MC = MR 12 = 20 – Q/25 300 = 500 – Q Q = 200 200 = 1,000 – 50P 50P = 800 P = 16.

b.

Loss of consumer surplus due to monopolization can easily be obtained from the graph (shaded portion). Area of shaded portion = (16 – 10)(200) + 1/2(16 – 10) (500 – 200) = 1,200 + 900 = 2,100. This area is much larger than loss of consumer surplus if monopolist’s MC = 10.

c.

The graph shows that the loss of consumer surplus is much greater here than in the usual case where monopolization does not affect costs.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

99


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

12.8

Result depends on how tax affects MR = MC solution. a.

i)

Tax affects MR, MC equally. So profit-maximizing output is not changed. Deadweight loss is unchanged.

ii)

Tax affects MR, not MC. Causes profit-maximizing output to fall. Deadweight loss increases.

iii)

Now  = TR – TC – t(P – MC)—the tax is on monopoly power. The profit-maximizing Q is such that

MR  MC  t (

( P  MC ) ) Q

but

( P  MC ) 0 Q for negatively sloped demand and positively sloped MC. Hence, MR < MC and output must have expanded. Hence deadweight loss falls.

12.9

12.10

b.

The graph shows the three post-tax equilibria:

a.

Setting MR = MC yields Q* = 3. Thus P* = 5 and profit is 9. The profit from 100 such consumers is 900.

b.

An individual’s consumer surplus at a price of 2 is 18, the highest admission fee that can be charged. With 100 such consumers, profit is 100 × 18 = 1,800 (all profit comes from the admission fee because there is no profit margin on drinks).

c.

With the pricing scheme from part b, profit is 115 × 18 = 2,070 with 15 new consumers. With a $3 price per drink, each original consumer buys 5 drinks and each new one 13 drinks. A total of (100 × 5) + (15 × 13) = 695 drinks are sold at a profit margin of $1 each. The admission fee has to be lowered to $12.50 not to deter original consumers (this is an original consumer’s surplus at the $3 price). Total profit from admission fees and drinks is 695 + (115 × 12.50) = 2,132.50.

a.

Setting MR = MC yields Qm = 40. Substituting into demand, Pm = 60. Profit is m = 600, which is computed as total revenue (60 × 40) minus total cost TC

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

100


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

= 1,000 + 20Q = 1,000 + (20 × 40) = 1,800. Consumer surplus equals the area of the shaded triangle in the graph below: CSm = 800. Social welfare is Wm = m + CSm = 1,400.

b.

See the figure below for this outcome. Social welfare is maximized by setting P = MC. From the demand curve, P = 100 – Q. So 100 – Q = 20 implies Q* = 80. Then P* = 20, * = –1,000, CS* = 3,200 (the area of the shaded triangle), and W* = 2,200. This policy would not be sustainable in the long run without subsidies because the firm is making negative profit and would exit if it could.

c.

See the figure below for this outcome. Compute the quantity under this form of regulation by finding the intersection between P (from the inverse demand curve) and AC. To compute AC, start from TC = 1,000 + 20Q, implying AC = TC/Q = (1,000/Q) + 20. Setting 100 – Q = (1,000/Q) + 20 leads to the quadratic equation Q2 – 80Q + 1,000 = 0 with roots (15.5, 64.5). The two roots correspond to the two intersections shown on the graph. As the graph shows, the relevant root is the larger one, Qr = 64.5 (the superscript refers to ―regulation‖). We also have Pr = 35.5, r = 0 (this must be true because P = AC implies zero profit), CSr = 2,079.8 (the area of the shaded triangle in the graph), and Wr = 2,079.8. This policy could be sustainable in the long run because the firm is at least breaking even, so has no

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

101


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

incentive to exit. Compared to the regulation in part b, then, social welfare is lower in c, but the policy may be have a more realistic chance of ―working‖ in practice.

[return to top]

Chapter 13: Imperfect Competition Purpose and Organization of the Chapter This chapter studies oligopoly behavior in a rigorous way using the tools of game theory introduced in Chapter 6. The chapter begins by placing imperfect competition on a continuum between monopoly (or a perfect cartel) and perfect competition. It then presents and analyzes some of the workhorse models of oligopoly pricing: Cournot, Bertrand, Bertrand with differentiated products, Bertrand with capacity constraints, collusion in repeated games, and so forth. It goes on to analyze advertising, strategic investment, entry, and entry deterrence. A section on consumer search complements the treatment of product differentiation and advertising. While most of the chapter focuses on game-theoretic models, some space is devoted at the end of the chapter to a discussion of influential models including price leadership and monopolistic competition that are not game-theoretic in the sense that behavioral assumptions are made for certain of the market participants. A rigorous discussion of limit and predatory pricing would require background on signaling games and other games of incomplete information. This background is not provided until Chapter 15. In the present chapter, we are content to provide an intuitive treatment of limit and predatory pricing, leaving a deeper analysis to Chapter 16.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

102


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Lecture and Discussion Suggestions It should not be hard to motivate student interest in imperfect competition. Most of the industries students would think of can roughly be characterized as oligopolies, so the chapter can be thought of as trying to model and analyze a lot of the real-world industries students might know about. Also, the chapter reinforces the value of the game theory students worked hard to master in Chapter 6 as a tool to analyze important economic questions. One point to make to students is how sensitive the results are to small changes in the assumptions (quantity rather than price competition, importance of the timing of moves, importance of firms’ information, and so forth). The implication is that it is unfortunately difficult to draw broad conclusions about imperfectly-competitive markets. Economists are drawn more and more to focus on individual industries for their analyses rather than cramming hundreds of industries into a single model. There is a high premium in this context on knowing how particular industries operate. Students might enjoy thinking about how the Internet has affected imperfect competition, corresponding to the section on search costs in the text and Application 13.4: Searching the Internet. Because of space constraints, the chapter only briefly mentions policy, in particular the implications for antitrust and regulatory policy. These topics would certainly spur student interest. One approach to discussing antitrust would be to cover a recent antitrust case such as the suit against Apple brought by Epic, maker of the computer game Fortnite, with which most students will be familiar. As of this writing, Epic is suing over the 30% cut that Apple takes for apps sold on its App Store, the only source for apps compatible with Apple’s IOS operating system. Apple contends that it only controls a small portion of the smartphone market, so is not a monopolist over the distribution of apps.

Cengage Supplements The following product-level supplements provide additional information that may help you in preparing your course. They are available in the Instructor Resource Center.  

Test Bank PowerPoint Slides

Chapter Objectives The following objectives are addressed in this chapter: 13.01

Determine how firms behave when they choose output (Cournot model) and when they choose prices (Bertrand model).

13.02

Understand how capacity constraints and product differentiation influence the prices chosen by firms.

13.03

Explain what makes it easier or harder for firms to tacitly collude.

13.04

Determine how two firms behave if one firm has a first mover advantage.

13.05

Explain how an incumbent firm may deter entry or force exit.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

103


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

[return to top]

What's New in This Chapter The chapter has two modest changes from the previous edition. 

Followed some work in the history of economic thought, we have sorted out the relative contributions of Cournot and Bertrand, although we stick with the standard convention of labeling the model with quantity competition the ―Cournot model‖ and that with price competition the ―Bertrand model.‖

We added a passage on recent price spikes in deregulated Texas electricity markets to Application 12.2.

[return to top]

Additional Resources The U.S. Census Bureau conducts a census of domestic business every five years. The bureau reports the number of firms, establishments (many firms have multiple establishments in multiple locations—this entry counts those separately), number of employees, and value of shipments in each industry category. Industries are organized by NAICS (North American Industrial Classification System) codes. This is a hierarchical system, where the successive digits reflect a more refined breakdown. For example, the two-digit code 31 represents manufacturing, 311 food manufacturing, and 3113 production of sugar and confectionery products, while 3114 represents fruit and vegetable preserves. Census uses these data to compute the Herfindahl-Hirschman Index (HHI) for the largest 50 firms in each four-digit industry. The HHI equals the sum of the squared market shares of the firms. For example, if the four-digit industry has four equally sized, firms, we obtain HHI = 252 + 252 + 252 + 252 = 2,500. Three equally sized firms yields a higher value, HHI = 33.32 + 33.32 + 33.32 = 3,333. A three-firm industry in which one has half the market and the other two split the residual equally yields a yet higher value of concentration measured this way: HHI = 502 + 252 + 252 = 3,750. The HHI attempts to provide a measure of the importance of just a few firms in the industry, or in other words, where the industry is on the continuum between perfect competition (HHI = 0) and monopoly (HHI = 1002 = 10,000). Antitrust authorities use the HHI as a quick index of whether enforcement is a concern in a particular market. The higher the HHI, the more scrutiny is devoted to mergers and consumer complaints in the industry. Care has to be used in interpreting the figure. Two industries with

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

104


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

equal HHIs may not be equally competitive; the homogeneity of the good, availability of price and product information, history of firm forbearance, and other factors may affect competition as well. In some cases, a four-digit NAICS code is a good boundary for a market. In other cases, the definition may be too broad or too narrow. The HHI is computed for the entire United States. In some industries, this competition is much more localized, so the U.S. measure probably understates the degree of local concentration. The latest release of the business census is 2017. HHI measures are reported in the table linked below, which the user can manipulate by changing the filter at the right side of the table. Using these data, the student can obtain a picture of the industries in which imperfect competition, the subject of this chapter, is a relevant consideration. https://data.census.gov/cedsci/table?q=concentration&n=N0000.00&tid=ECNSIZE 2017.EC1700SIZECONCEN&hidePreview=false

[return to top]

Solutions to End of Chapter Problems Students have access to solutions for the odd numbered problems as well as video problem walkthroughs for problems 4 and 8.

13.1

a.

The Nash equilibrium is for both to price low. b.

13.2

You could relabel ―Low Price‖ as ―High Output‖ and ―High Price‖ as ―Low Output.‖

a.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

105


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

13.3

b.

At point C, P = MC = 6. Q = 10,000 – 1,000 × 6 = 4,000. Industry profit = 0. Consumer surplus = (4,000)(10 – 6)/2 = 8,000. Social welfare = profit + consumer surplus = 0 + 8,000 = 8,000.

c.

At point M, quantity is given by equating MR and MC: 10 – Q/500 = 6, implying Q = 2,000. Substituting Q = 2,000 into the demand curve, 2,000 = 10,000 – 1,000 P, implying P = 8. Industry profit = (8 – 6)(2,000) = 4,000. Consumer surplus = (2,000)(10 – 8)/2 = 2,000. Social welfare = 4,000 + 2,000 = 6,000.

d.

At point A, price is halfway between 6 and 8, that is, P = 7. Q = 10,000 – 1,000 × 7 = 3,000. Industry profit = (7 – 6)(3,000) = 3,000. Consumer surplus = (3,000)(10 – 7)/2 = 4,500. Social welfare = 3,000 + 4,500 = 7,500.

Equation (13.4) states the marginal revenue for Cournot firm A with the given demand curve is 120 – 2qA – qB. Equating this marginal revenue with marginal cost 30, 120 – 2qA – qB = 30 implies 90 – 2qA – qB = 0. Similarly, for firm B, 90 – 2qB – qA = 0. Solving the two preceding equations simultaneously gives q *A  q *B  30.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

106


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Industry output = 30 + 30 = 60. To find P, solve 60 = 120 – P, implying P = 60. Firm profit = (60 – 30)(30) = 900. Industry profit = 2 × 900 = 1,800. 13.4

a.

Solving the two equations

PA 

1  PB 1  2 PA and PB  4 4

simultaneously gives PA*  PB*  1 / 2.

13.5

13.6

b.

See graph above.

c.

See graph above.

a.

The are many Nash equilibria. Firm A charges any price along the one-centincrement grid from $8.02 to $10.01 (inclusive). Firm B undercuts A by one cent. All of these involve weakly dominated actions for firm A except the highest price one, in which it charges $10.01 and B charges $10. Firm B gets all the demand. Assume throughout the remainder of the answer that this is the Nash equilibrium that is played. Leaving the complications associated with the large number of equilibria aside, it is sufficient that students realize that prices will be around $10 and the low-cost firm will make all the sales.

b.

A earns zero profit. B earns 10 – 6 = 4 per unit and sells Q = 500 – 20 × 10 = 300 units for a profit of 4 × 300 = 1,200.

c.

Price equals marginal cost as in the Bertrand Paradox, though the price is equal to the high-cost firm’s marginal cost. One of the firms earns zero profit as in the Bertrand Paradox, but unlike in the Bertrand Paradox one of the firms earns positive profit.

a.

Substituting

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

107


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

qB 

120  q A 2

into Equation 13.3 and simplifying gives

 120  q A  . 2  

 A  qA

Analogous to Equation 13.3, we can write B’s profit as q B (120  q A  q B ).

Substituting B’s best-response

qB 

120  q A 2

into B’s profit function and simplifying gives 2

q    B   60  A  . 2  

Finally, substituting qB = 0 into Equation 13.3 gives

 M  q A (120  q A ). b.

c.

qA

πA

πA

πA

0

0

2,600

0

20

1,000

2,500

2,100

40

1,600

1,600

3,200

60

1,800

900

3,600

80

1,600

400

3,200

100

1,000

200

2,800

120

0

0

0

It confirms the Stackelberg outcome because πA is highest for qA = 60. If B’s fixed cost were 400 or more, A would need to produce about 80 to deter entry. If B’s fixed cost were 100, A would have to produce about 100 to deter entry. A would try to deter entry even if B’s fixed cost were 100 because A’s monopoly profit given it produces 100 is 2,000, which exceeds the Stackelberg profit of 1,800.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

108


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

13.7

Dividing both sides of equation (13.15) by πM, collusion is sustainable for N  1 /(1  g ) . The following is a graph of the upper bound.

As indicated by the dotted line, for g = 0.95, collusion is sustainable with 20 or fewer firms. 13.8

a. B Enter

Don’t

Enter

–10, –10

20, 0

Don’t

0, 20

0, 0

A

13.9

b.

The mixed-strategy Nash equilibrium is for each firm to enter with probability 2/3 and stay out with probability 1/3.

c.

The mixed-strategy Nash equilibrium shares the feature with the Bertrand Paradox that firms earn no expected profit. The similarity ends there. With some probability, only one firm enters and behaves as a monopolist. With some probability both firms enter and earn negative profit.

First suppose FI > 2,000. Then I will not prey. E earns 1,600 – K > 0 if it enters, and so will enter. Next suppose FI < 2,000. Then I would prey if E entered. E would earn –K – FE < 0 if it entered, and so would choose not to. Predation would not be observed in either case. The only case in which I would be inclined to prey (if FI < 2,000), E does not enter and so there is no firm to prey upon.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

109


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

13.10

QD = –2,000P + 70,000. a.

1,000 firms. MC = q + 5. Price taker: set MC = P, implying q + 5 = P, in turn implying q = P – 5. 1000

QS   q  1,000 P  5,000 1

To find equilibrium, set QD = QS . –2,000P + 70,000 = 1,000P – 5,000. 3,000P = 75,000. P = 25, Q = 20,000. b.

Demand for leader = Market demand – Quantity supplied by fringe. QDL = (–2,000P + 70,000) – (1,000P – 5,000). QDL = –3,000P + 75,000.

c.

Have that MRL = –QL/1,500 + 25 and that MCL = 15. For profit maximization, set MRL = MCL. This implies –QL /1,500 + 25 = 15. Solving, 10 = QL /1,500, or QL = 15,000 and P = 20. Total QD = 30,000.

[return to top]

Chapter 14: Pricing in Input Markets Purpose and Organization of the Chapter This chapter provides a brief introduction to supply and demand in input markets. To make the analysis less abstract, most of the focus is on labor markets though it might hold equally well for any other input market. The first half of the chapter develops the marginal productivity theory of input demand. It concludes with a summary of substitution and output effects, both of which suggest that any input demand curve will be downward sloping. The second half of the chapter begins with a brief discussion of labor (and other input) supply indicating why supply curves are likely to be up-ward sloping. Much of the material in this section is pursued at greater length in the Appendix, which examines the labor/leisure model of individual input supply. Two final topics are discussed: (1) Comparative statics of supply and

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

110


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

demand (with a summary of why curves shift, in Table 14.2); and (2) Monopsony (including the case of bilateral monopoly).

Lecture and Discussion Suggestions The presentation of input demand in Chapter 14 includes two analytical concepts that are especially difficult for students and these should be featured in lectures. First, the output and substitution effects from a change in input prices should be carefully described and differentiated from the analogous presentation in consumer theory. The notion that individuals have budget constraints, but firms do not (that is, firms sell the level of output that maximizes profits) is straightforward, but somehow difficult for students to grasp. One approach is to use the result that cost minimization requires MPL MPK  w v

(1)

and note that this common ratio equals 1/MC. Changes in an input’s price, therefore, prompts substitution effects from equation 1 and output effects from changes in 1/MC (which must equal 1/MR for profit maximization). Both of these changes imply input prices and levels of input use move in opposite directions. Marginal expense (sometimes termed ―marginal factor cost‖) is the second concept from Chapter 14 that is difficult for students to grasp. A brief review of monopoly theory and the marginal revenue concept may help to make the point since the arguments are formally identical. In particular, students might be asked to study the similarities between Figures 14.5 and 12.1. A supply-demand lecture on the material in the second half of Chapter 14 should try to help students to get the comparative-statics analyses correct. Reasons for shifts in demand or supply curves in labor markets are complex, both because traditional roles of producers and individuals are re-versed and because the analytics of responses to price changes are difficult. Explaining in detail why each of the entries in Table 14.2 shifts the curves in the directions indicated, though a dull exercise, may significantly aid students’ understanding. There is no end to potentially interesting discussion topics for the material covered in Chapter 14. The Applications in the chapter provide a starting point for some of these (for example, Application 14.2: Controversy over the Minimum Wage or Application 14.3: Why is Wage Inequality In-creasing?). Other topics might include (1) the effects of trade on wages and whether trade is to blame for the widening wage dispersion, (2) compensating wage differentials for dangerous jobs, or (3) changing rates of workforce unionization.

Cengage Supplements The following product-level supplements provide additional information that may help you in preparing your course. They are available in the Instructor Resource Center.  

Test Bank PowerPoint Slides

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

111


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Chapter Objectives The following objectives are addressed in this chapter: 14.01

Determine how much of an input a profit maximizing firm will hire.

14.02

Explain how a firm will respond to a change in the price of one or more of its inputs.

14.03

Understand the factors that cause equilibrium input prices to change.

14.04

Explain how a firm with buying power in input markets can influence the quantity and price of that input.

[return to top]

What's New in This Chapter Most of the changes in this chapter involve additions and modifications to the applications. Some examples include the following. 

In Application 14.1, the relationship between aircraft types and jet-fuel demand has been updated.

Application 14.2 provides a new summary of the literature on minimum wage effects.

Application 14.3 expands the discussion of measuring wage inequality.

Application 14.5 contains a new section on public sector unions.

Application 14A.2 provides a further look at the Earned Income Tax Credit.

Many of the more complex figures in the chapter are improved using multi-colors. That is especially the case for the monopsony and bilateral monopoly figures.

[return to top]

Additional Resources The Econofact website offers a number of even-handed summaries of the literature on the employment effects of various public policies. The articles on the minimum wage and on unionization are especially recommended. See: https://econofact.org/tag/employment. Under the CARES Act unemployment benefits were increased by as much as $600 per week during the early stages of the COVID pandemic. Prior studies of the effect of unemployment insurance on job finding would have predicted large negative employment effects of such an expansion. The research on this topic is just beginning, but Marnescu, Skandalis, and Zhao in

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

112


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

―Job Search, Jobs Postings, and Unemployment Insurance During the COVID-19 Crisis‖ find only minor impacts. See: https://papers.ssrn.com/sol3/papers.cfm?abstract_id=3664265. Gerald Scully’s original study of the effects of the reserve clause on pay in major league baseball remains the most accessible and interesting work on the topic. See ―Pay and Performance in Major League Baseball‖, American Economic Review, December, 1974.

Solutions to End of Chapter Problems Students have access to solutions for the odd numbered problems as well as video problem walkthroughs for problems 1 and 7.

14.1

a.

With five workers, put each successively where its marginal product is greatest. First worker goes to A, second goes to B, third goes to A, fourth goes to C, fifth goes to A. Output = 21 + 8 + 5 = 34. MP of last worker is 4. b.

P  MPL = $1.00  4 = $4.00 = w. With five workers, the wage bill is wL = $20. Profits are  = TR – TC = PQ – wL = $34 – $20 = $14.

c.

Marginal products of labor on the various farms are: Workers

MPA

MPB

MPC

1

10

8

5

2

7

3

2

3

4

2

1

Because output costs $1, these figures also represent marginal value products. With a wage of $5 four workers are hired (two on A and one on B and C). With a wage of $4, five workers are hired (three on A, one on B and C). With a wage of $3 six are hired (three on A. two on B and one on C). 14.2

For this problem, the production function is q = 10,000 L and MPL = 5,000/ L . a.

Since P = 0.01 here and the firm is a price taker, profit maximization requires that w = P · MPL = .01(5,000/ L ) = 50/ L . Since w = 10, this means

L = 5, L = 25.

If w = 5: 5 = 50/ L

L = 100.

and if w = 2: 2 = 50/ L L = 625

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

113


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

As these points suggest, this demand curve has a hyperbolic shape.

b.

Assuming w = $10, the value of the marginal product is P  5,000/ L ; If P = 0.10:

10 = 500/ L :

L = 50, q = 500,000

If P = 0.05;

10 = 250/ L :

L = 250, q = 250,000

If P = 0.02:

10 = 100/ L :

L = 10, q = 100,000

The graph shows the supply curve for licked envelopes.

14.3

a.

w = v = $1, so K and L will be used in a one-to-one ratio. TC  vK  wL  K  L  2 L so 2L 2L 2L AC    2 q KL L2

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

114


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

b.

Since P = 2, quantity demanded is Q = 400,000 – 100,000(2) = 200,000 pipes

q

200,000 pipes  200 pipes / firm 1000 firms

q = 200 =

L K = 4 so 200 workers are hired per firm, 200,000 by the industry.

c.

When w = $2 and v = $1, cost minimization requires K/L = 2. TC = wL + vK so = 2L + K = 2 2.

d.

Now Q = 400,000 – 100,000 (2 2 ) or Q = 117,200 q = 117.2 = L( 2 ) L = 117.2/ 2 = 82.9 So hiring is 82,900

e.

14.4

If output had stayed at q = 200, L = 200/ 2 = 141.4 so total hiring would be 141,400. Reduction from 200,000 to 141,000 is the substitution effect. From 141,400 to 87,900 is the output effect.

Demand: L = –50w + 450. Supply: L = 100w. a.

Equilibrium can be found by setting quantity supplied equal to quantity demanded. 100w = –50w + 450. w = 3, L = 300.

b.

With a subsidy, demand is now L = –50(w – s) + 450. Where s = the subsidy, want w = $4 and so Ls = 400. The new equilibrium is formed by 400 = –50(4 – s) + 450. Hence, s = 3 and this total subsidy is 400  3 = $1,200.

c.

With a declared minimum wage of $4, D = 250, S = 400 so there is unemployment of 150.

d.

The graph shows these various equilbria in the labor market.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

115


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

14.5

a.

Demand: K  1500  25v Supply: K  75v  500 Equilibrium is found by setting quantity supplied equal to quantity demanded. 1500  25v  75v  500; 2000  100v; v  20, K  1,000 .

b.

With g = 2, demand is K  1100  25v . Equilibrium is v  16, K  700 . With g = 3, demand is K  800  25v . Equilibrium is v  13, K  475 .

c.

d.

Need to restore the rental rate to v  16. Let s be the subsidy per car. Then demand is K  800  25(v  s) . Setting this equal to supply yields:

800  25(v  s)  75v  500  1300  100v  25s. With v  16, this implies s  12. 14.6

MEL 

Supply:

L = 100w

Demand:

MRP  10  0.01L

a.

L 50

Hence, profit maximization requires

L L  10  50 100

or L = 333

Can get w from the supply curve, w = L/100 = 3.33 At L = 333, MRP = 6.67, so workers receive only about half their marginal products. b.

For perfectly competitive labor market MRPL = w in equilibrium. So from supply curve

w  L 100  MRPL  10  0.01L L  500 w5 The graph shows both the monopsonistic (M) and competitive (C) equilibria.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

116


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

14.7

Supply: L  80 w MEL 

L 40

Demand: MVPL  10  L 40 a.

For monopsonist MVPL  MEL MVPL  10 

L L  40 40

L  200

Get w from supply curve:

w b.

L  2.5 80

For Carl, the marginal expense of labor now equals the minimum wage and in equilibrium the marginal expense of labor will equal the marginal revenue product of labor. wm = MEL = MVPL wm = $3.00. Carl’s Demand L = 400 – 40MVPL

Supply L = 80w

L = 400 – 40(3)

L = 80(3)

L = 280.

L = 240.

Since quantity demanded exceeds quantity supplied. Carl will hire 240 workers, with no unemployment. To study effects of minimum, try $3.33 and $4.00 wm = $3.33 Carl’s Demand L = 400 – 40(3.33)

Supply L = 80(3.33)

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

117


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

= 267.

= 267.

Demand = supply, Carl will hire 267 workers, with no unemployment. wm = $4.00 Carl’s Demand Supply L = 400 – 40(4.00) L = 80(4.00) = 240.

= 320.

Supply > demand, Carl will hire 240 workers, unemployment = 80.

14.8

c.

The graph shows these various responses to a minimum wage.

d.

Under perfect competition, a minimum wage means higher wages but fewer workers employed. Under monopsony, a minimum wage may result in higher wages and more workers employed as shown by some of the cases studied in part b.

Here marginal value product is $10 per hour: MEm =

Lm /2 = 10 so Lm = 400, wm = 20/3 = 6.67.

MEf = Lf /50 = 10 so Lf = 500, wf = 5. Profits = TR – TC = 9,000 – 5(500) – 6.6(400) = 3,833. If Ajax must pay the same wage, w = MVPL = 10, L = 1,000 + 900 = 1,900. Profits = 19,000 – 1,900  w = 19,000 – 19,000 = 0. 14.9

a.

Budget constraint: C = w(24 – H) + 10.

b.

Due to Mrs. Smith’s preferences, she insists on spending half of potential earnings (w × 24 + 10) on consumption and half on leisure. This means value of consumption = value of leisure (i.e., w  H) for all wage rates. C = wH Substituting for C: w(24 – H) + 10 = wH 24 – H + 10/w = H

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

118


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

2H = 24 + 10/w H = 12 + 5/w

14.10

For w = $1.25

H = 16 C = 1.25(24 – 16) + 10 = 20.

For w = $2.50

H = 14 C = 2.50(24 – 14) + 10 = 35.

For w = $5.00

H = 13 C = 5.00(24 – 13) + 10 = 65.

For w = $10.00

H = 12.5

C = 10.00(24 – 12.5)+10 = 125.

c.

The graph shows Mrs. Smith’s changing choices as the wage rises. Hours of leisure (H) fall toward 12 as w rises.

d.

Mrs. Smith’s labor supply curve can be constructed directly from the data in part b. It is upward sloping, being asymptotic to 12 hours as w rises.

e.

With a $20 inheritance, algebra in problem 14.9b shows that H = 12 + 10/w. Hence, L = 24 – H = 12 – 10/w. The supply curve shown in part d would shift inward.

a.

Earnings on 8 hour days are 400. Hence utility is 20. On the variable hours job, with a wage of 50 earnings are 200 and 600 (and therefore average 400). But utility on the variable hours job is 0.5 200  0.5 600  19.3185 , so this person will require a higher wage to take the variable hours job. Using an Excel spreadsheet shows that the required wage is about 53.5.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

119


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

b.

A proportional tax will not affect the utility calculation because the tax rate will factor out of all of the expressions for utility.

c.

With the progressive tax utility from the constant hours job is 18.7083 whereas from the variable hours job with a wage of 53.5 utility is 18.1657. The difference arises because taxes paid are 50 under the constant hours job and 171 under the variable hours job. A higher wage (about 57.5) would now have to be offered on the variable hours job to get this person to take it.

d.

To answer this part one must assume something about the distribution of jobs. Assuming constant and variable hour jobs are equally numerous, will need to collect 0.5  50  0.5 171  110.5 from each job. This requires a proportional tax rate of t  110.5 400  0.276.

[return to top]

Chapter 15: Capital and Time Purpose and Organization of the Chapter The general purpose of this chapter is to provide students with some basic tools for looking at economic activity in a dynamic context, with a particular focus on capital markets. The chapter starts with a discussion of the why time is important for capital decisions and then turns to the examination of a simple two period model of consumer behavior. The contrasting income and substitution effects of a change in the real interest rate on current consumption are highlighted. At this point students may need reminding that, because current period income is fixed, any change in current consumption also shows up in current savings decisions. Savings are treated here as providing the ―supply of loans‖ in the interest rate determination process. The demand for loans side of the market is treated here as part of the firm’s decision about capital equipment. Specifically, the chapter shows that the real interest rate is an important component of the rental rate on capital, v, via the formula v = P ( r + d). Because the firm’s demand for equipment is a downward sloping function of the real interest rate, the de-mand for loans will be also. Ultimately then the real interest rate is deter-mined by the supply and demand for loans. The final sections of the chapter illustrate how the real interest rate provides a ―price‖ that ties together present and future periods. The concept of present discounted value is briefly described (the chapter’s appendix goes into more detail) and this is then used to discuss pricing of finite resources.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

120


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

The mathematical appendix to Chapter 15 provides a general introduction to compound interest formulas. It is included here not so much for its utility to the textbook, or to the course as a whole, but simply because we have been shocked by how little many students know about the subject. Here the main intent is to illustrate compounding, to discuss the logic of discounting future payments, and to introduce continuous interest concepts. Other than a few references to the present value notion, the text itself does not make much use of these concepts, though some instructors may wish to introduce more inter-temporal material on their own.

Lecture and Discussion Suggestions A complete coverage of this chapter probably requires two lectures (assuming that students can read all of the material on interest rates on their own). The first would focus on the two-period model of consumption choices. Two features of that model might be stressed: (1) Why the real interest rate is the key price (rather than the nominal rate); and (2) How the model might be generalized to more general cases involving many periods and complex income flows (see Problem 15.1 for the case of income in two periods). This lecture would be a good place to tie this model to the variants of the life cycle model usually discussed in macroeconomics. A second lecture for this chapter might involve the interest rate determination process. Students will probably have some trouble seeing how the simplified model of interest rate determination in Figure 15.3 applies to the real world—especially since their macro courses may provide a different view. Indeed, we believe it is for this very reason that this graph should be stressed as the only theoretically sound model of interest rate determination (good for starting an argument with your macro colleagues). The lecture might then be expanded to include the brief material in the chapter and the next about other financial assets to show how they fit into this general picture. This section of the course is the ideal place in the course to discuss financial markets—a subject that never fails to raise student interest. We have found that using various tables from the Wall Street Journal together with the question ―How does this relate to Figure 15.3?‖ can be an effective way of helping students to gain some conceptual understanding of what financial markets are doing. Covering corporate shares in this way is, of course, very complicated because most students will have no finance background. But this is a good place to make the pitch for taking a finance course.

Cengage Supplements The following product-level supplements provide additional information that may help you in preparing your course. They are available in the Instructor Resource Center.  

Test Bank PowerPoint Slides

Chapter Objectives The following objectives are addressed in this chapter: 15.01

Understand how changes in the interest rate affect household saving.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

121


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

15.02

Explain how changes in the interest rate affect firms’ desire to rent or own capital and the demand for loans.

15.03

Explain how the real interest rate is determined by both individuals who supply loans and firms who demand loans.

15.04

Understand how to compare expenditures and receipts received in different periods.

15.05

Explain how resource scarcity affects the current price of resources.

[return to top]

What's New in This Chapter Most of what is new in this chapter appears in the applications. Here are a few specifics. 

Application 15.1 offers some new material on the study of savings and asks students to think about how the financial support offered to families during the COVID-19 pandemic may have affected savings (see also the ―Additional Resources‖ entry on this topic).

Application 15.5 provides some new material on interpreting the volatility in stock and derivatives trading.

Application 15.6 provides updated data on natural resource pricing. Interestingly, the data from 2020 indicate that the real price of some natural resources have risen a bit after long periods of decline.

All the applications in the interest rate appendix to this chapter have been updated to 2020. This updating makes the long-term effects of compounding even more farcical and illustrates some of the very low interest rates currently prevailing.

[return to top]

Additional Resources The increase in the U.S. savings rate in 2020 was remarkable. This situation provides a good opportunity for instructors to combine the theory of savings offered in the text with theories of savings usually studied in macroeconomics. The most important insight provided by the macro literature is the differentiation between ―permanent‖ and ―transitory‖ income in affecting savings. For a discussion of this issue in the COVID-19 context see: https://www.kansascityfed.org/research/economic-bulletin/why-are-americanssaving-so-much-income-2020/. Negative interest rates are not discussed in the text, in part because the theory of such rates in not well-developed. Nevertheless, negative rates are now an important part of the economic landscape, so students might be asked to learn something about them. In theory they could be asked whether Figure 15.3 rules out negative rates as a theoretical possibility (it doesn’t). For real world insight see:

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

122


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

https://www.youtube.com/watch?v=sLWulVrQvsY. There is currently much debate about ―infrastructure investment.‖ Usually this is treated as a government responsibility, but students might be asked how the need to invest in infrastructure relates to the theory of private investment in the text. Specifically, the following YouTube video looks at possibilities for building a new railroad tunnel between New York and New Jersey (and other infrastructure issues). Why should these investments be financed by government? https://www.youtube.com/watch?v=EdvJSGc14xA.

[return to top]

Solutions to End of Chapter Problems Students have access to solutions for the odd numbered problems as well as video problem walkthroughs for problems 3 and 6.

15.1

a.

15.2

The budget constraint shows that spending must equal income in present value terms, but income and consumption are not constrained to be equal in either period. b.

If this individual saves in period zero, consumption will of necessity exceed income in period one.

c.

Because period zero savings ( Y0  C0 ) earn interest, more can be spent in terms is dissaving ( C1  Y1 ) in period one.

a.

Felix’s indifference curves are straight lines with slope

C1  (1   ) C0

b.

Since the budget constraint is

I  C0 

C1 1 r

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

123


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

its slope is –(1 + r). So, if r >  utility maximum will be where C0  0 .

15.3

c.

Similarly, if r < , the budget constraint is flatter than the indifference curve and utility maximization requires C1  0 .

d.

The larger  is (that is, the more Felix discounts the future), the more likely he is to be in case c, where savings = 0.

a.

Present value of income is 50,000 + 55,000/(1 + r) = 50,000 + 55,000/1.1 = 100,000.

C1  1.1 3C0

b.

Prudence has MRS = 1 + r or

c.

Budget constraint in present value terms is

100,000  C0 

C1 1.1

Using the utility maximizing condition from part b gives 100,000  C0  3C0

Hence C0 = 25,000. Savings in period 0 are 25,000. d.

3C1 . Substitution into budget constraint (Prudence and Glitter C0 have the same budget constraint) yields

For Glitter MRS 

100,000  C0 

C0 4C0  3 3

Hence, for her, C0 = 75,000. Savings in period zero are –25,000. That is, she borrows with the intention of repaying later. It seems that their names are warranted. 15.4

a.

v = P (r + d) = 2,000 (0.05 + 0.10) = 300.

b.

TC = vR = 300R = 300T /100 = 3T .

c.

MC = 6T = P = 60, hence T = 10.

d.

If T = 10, R = 1.

15.5

a.

Assuming revenues are received at the end of each year gives a present value of $486.841 when r = 0.1. This falls short of the current purchase price of $500,000 for the ten trucks. When r = 0.08, the present value if future revenue is $520,637, which means that the investment would be profitable.

15.6

V  100t  6t 2

2

proportional growth is

2

100  12t V

Value greatest at t = 8.33

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

124


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

If r = 0.05, set r = growth rate

0.05 

100  12t or 100  12t  5t  3t 2 2 . 100t  6t

2

0.3t - 17t + 100 = 0 which can be factored as (0.3t – 2)(t – 50) = 0. Hence, t = 6.67 or t = 50. As the graph shows, however, only the first root provides an optimal solution since the second root is extraneous.

15.7

15.8

a.

Price should be 4,000/(1.05)25 = 4,000/3.3864 = 1,181.

b. c.

Scarcity costs = 1,181 – 100 = 1,081. Assuming real production costs stay at $100, scarcity costs in 25 years are 3,900.00

d.

In 50 years price is 1,181(1.05) = 4,000(1.05) = 13,542.

50

25

Salesman’s pitch ignores the opportunity cost of interest. 4

2,000 (1  r )i 1

PDV ( wholelife)   36

400 i 1 (1  r )

PDV (term)   if r = 0.10,

PDV (whole life) = $6,340 PDV (term) = $3,858 Term is much cheaper. 15.9

The fallacy here is that the calculation assumes that you have borrowed $10,000 for all three years. Since the repayment plan includes some repayment of the $10,000 too, the effective amount borrowed in only about half that amount. The actual effective interest rate on the loan, if the $315 payments are made at the start of each month, is about 8.7 percent, well above the 5 percent opportunity cost.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

125


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

15.10

10 10   200 i 0.05

a.

PDV 

b.

Nominal payments are now Pi  10(1.03)i but real payments are

RPi  10(1.03)i /(1.03)i  10

RPi  10(1.03)i (1.03)i  10 and calculation of PDV

is as before. c.

10  125 or 0.08 10 / (1.03)i 10 10 10 PDV       125 i i i (1.05) [(1.03)(1.05)] (1.08) 0.08 PDV 

[return to top]

Chapter 16: Asymmetric Information Purpose and Organization of the Chapter Up to this chapter in the text, economic agents were all assumed to have the same information about the market. This chapter studies how the situation changes when one or another agent has asymmetric (or private) in-formation that the other does not. Topics include the moralhazard and adverse-selection problems, auctions, the lemons problem, and signaling models. These topics represent some of the most active areas of microeconomic research in the past several decades. The chapter ties in with Chapter 5 on uncertainty and Chapter 6 on game theory. In effect, the chapter studies how game theory can be applied to games where there is uncertainty about other players’ payoffs. The broad theme of the chapter is incentives. Much of the book is spent on the idea of the incentive effect of prices. If a good’s price goes up, consumers may substitute toward another good; a firm may substitute toward another input. The idea carries over in this chapter, except that in an effort to get around the inefficiencies associated with asymmetric information, parties resort to more sophisticated contracts and strategies than just simple prices. Still, the purpose of these contracts and strategies is to provide incentives. In the moral-hazard problem, contracts provide effort incentives. In the adverse-selection problem, contracts try to get around the agent’s incentive to pay a lower price by hiding his or her type. In a signaling

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

126


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

model, a high-ability worker may try to reveal his or her ability by choosing an education level that the lower-ability worker would never have an incentive to choose. Another broad theme of the chapter is that asymmetric information leads the market to be inefficient, just as does monopoly or externalities. (Indeed, this is the logic for placing the chapter in Part 8 of the text on market failures.) Unlike some of these other market failures, there may be no good government solution (such as subsidies or regulation) to the problem of asymmetric information because the government is likely to be as uninformed as the uninformed parties in the market.

Lecture and Discussion Suggestions Unfortunately, most intermediate microeconomics classes have already run short on time before coming to the topic of asymmetric information. It may be possible to cover a simple application of one or both of the two variants of the principal-agent model (hidden actions, hidden types) in one lecture. A second possibility is to focus on signaling games as perhaps a supplement to another unit. It could supplement a unit on game theory after the material in Chapter 6 is covered. Or it could supplement a unit on labor markets after the material in Chapter 14 is covered. This would be an ap-propriate supplement since the chief application of the signaling model dis-cussed in the text is to signaling worker productivity through education. There would be an interesting contrast between a model in which wages are set when productivity levels are known and one in which productivity is private information for workers. A third possibility is to focus on auctions. Auction markets are of growing importance in the economy, and students might be familiar with auctions through participation themselves in online auctions. Classroom exercises could include running auction experiments in the classroom or covering an academic paper (there are a number of readable ones) about online auctions, or simply have the students do a bit of data collection from online auctions they have some experience with (eBay, say). If there is time for fuller coverage of the material in the chapter, there are several approaches to the material that might be considered. The text starts with the principal-agent model and introduces the two variants—hidden actions and hidden types—through that lens. The application to insurance comes a bit later in the chapter. An alternative approach that might also work well would be to introduce the insurance application of moral hazard and adverse selection first, and then move to other applications such as the manager-worker relationship or the monopoly-consumer relationship (if at all).

Note on Terminology Economists generally use the term ―moral-hazard problem‖ for any situation in which the agent takes a hidden action after contracting, whether in an insurance setting or not. We also adopt this usage in the text. Some economists use the term ―adverse-selection problem‖ for any situation in which the agent has a hidden type. This usage is fine; indeed, this usage shows up in some of our previous editions.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

127


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

We try to be a bit more careful in recent editions, distinguishing the adverse-selection problem as a special case of more general hidden-type problems. ―Adverse selection‖ refers to cases in which the worst types tend to show up in the market. One example is health insurance, in which the least healthy people have the most incentive to buy insurance and are also the most expensive to insure. Another example is used-car sales, in which owners with the worst-quality cars tend to be most interested in selling them. Not all hidden-type problems are adverse-selection problems. Take, for example, the application to nonlinear pricing by a monopoly coffee shop that is studied in detail in the chapter. This is certainly a hidden-type problem, but it is a stretch to call it an adverseselection problem. Holding quantity constant, the cost of serving coffee consumers is independent of their type. This contrasts the health-insurance application, in which providing a given quantity of insurance, say full insurance, to a consumer is more costly the less healthy he or she is.

Cengage Supplements The following product-level supplements provide additional information that may help you in preparing your course. They are available in the Instructor Resource Center.  

Test Bank PowerPoint Slides

Chapter Objectives The following objectives are addressed in this chapter: 16.01

Understand the difference between principal-agent problems with hidden actions and hidden types.

16.02

Explain how a manager can design a contract to provide the worker with the right incentive to work hard.

16.03

Know how sellers faced with different consumer types can design contract options to make the most money.

16.04

Explain how asymmetric information can lead to an adverse-selection problem in which only costly consumers participate in the market.

16.05

Understand how a seller can use an auction to reduce some of the problems that occur when its customers have hidden types.

16.06

Know how to solve for the equilibria in an education signaling game.

[return to top]

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

128


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

What's New in This Chapter The chapter is similar to that in the previous edition. The exposition has been tweaked in various places to streamline it. Some of the names have been changed in end of chapter problems, but the content of the problems is the same.

[return to top]

Additional Resources The American Economic Association posts videos of a diverse set of academic economists discussing their relevant research on the website http://diversifyingecon.org/index.php/Videos_on_economists_and_their_research. Two videos touch on performance incentives of agents within firms. Antoinette Schoar discusses whether a manager’s ―style‖ can be uncovered, determined in part by the economic conditions when they got their first managerial position. https://www.youtube.com/watch?v=FdUJqsEx-TU. Richard Freeman discusses the impact of outsourcing on the performance of agents in a firm. https://www.core-econ.org/the-economy/book/text/06.html#2Zm5ZLMKhgQ.

[return to top]

Solutions to End of Chapter Problems Students have access to solutions for the odd numbered problems as well as a video problem walkthrough for problem 5.

16.1

a.

Using the information that iSpys sell for $100 each, the equations for the graphs are , , and , which look as follows:

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

129


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

b.

Naum’s marginal cost of effort is $1. His marginal benefit is the marginal product of effort, , times , the revenue that goes to him from sales of the iSpys he assembles. Equating marginal costs and benefits, √ ⁄ . Keeping the equation in this form helps when we Rearranging, √ ⁄ . That is, Naum’s output is half substitute into the output function: √ of the term in the incentive schemes. So Naum’s output is 0 with the first scheme, 20 with the second, and 30 with the third.

c.

Blair should offer the scheme yielding the most profit. The first scheme yields no output, just a cost of 750 for the wage payment. The second scheme yields profit equal to revenue minus the wage payment: (

)

The third scheme turns out to yield the greatest profit: ( 16.2

a.

)

With a half share, EU = (0.5)(1,000/2) + (0.5)(400/2) – 100 = 250. With a quarter share, EU = (0.5)(1,000/4) + (0.5)(400/4) – 100 = 75. She would accept either contract because either provides her with positive expected utility. The lowest share s she would accept solves (0.5)(1,000s) + (0.5)(400s) – 100 = 0, implying s ≈ 14%.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

130


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

16.3

b.

The most she would pay equals (0.5)(1,000) + (0.5)(400) – 100 = 600.

c.

Orsina would need to be offered a fixed salary solving (0.5)(100) + f – 10 = 0, or f = 50.

a.

From part a of Problem 16.2, if she receives half of a firm’s return, Orsina’s expected utility from exerting effort is 250. If she does not exert effort, her utility is 400/2 = 200 < 250. So she will exert effort. We saw in Problem 16.2 that she would accept the contract. With a quarter share of gross profit, by part a of Problem 16.2 her expected utility from working is 75. Her expected utility from not working is 400/4 = 100 > 75. Orsina would accept the contract and not exert effort. For Orsina to exert effort, her gross-profit share must solve (0.5)(1,000 s) + (0.5)(400 s) – 100  400 s, or s  1/3.

b.

If she works hard, her expected utility with the bonus is (0.5)(100) – 100 = –50. If she does not work hard, her utility is 0. So she would not work hard. (Adding a fixed part to the wage would not change the answer.) The bonus b that would induce her to work hard solves (0.5) b – 100  0, or b  200. She would not need an additional fixed wage since the bonus also would give her at least as much expected utility as her outside option.

16.4

Adult bundle: 2 ounces sold at 36 cents. Children’s bundle: 4 ounces sold at 112 cents.

16.5

a.

Small cup: 8 ounces sold at 80 cents. Large cup: 10 ounces sold at $1.50. Consumers obtain no net surplus. Ahab earns (50)[0.80-(8)(0.05)] = $20 profit from small consumers and (100)[1.50-(10)(0.05)] = $100 profit from large consumers for a total of $120.

b.

Big consumers would obtain (8)(0.15) – 0.80 = 0.40 > 0 net surplus.

c.

The 8-ounce cup sells for 0.80. The price for the 10-ounce cup satisfies (0.15)(10) – p  0.40, where the right-hand side, 0.40, is the large consumer’s net surplus from buying the 8-ounce cup (see part b). The highest such price is p = 1.10.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

131


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Ahab’s profit is $20 from sales of the 8-ounce cup (see part a) and (100)(1.10 – 0.50) = $60 from large consumers for a total profit of $80. d.

The 6-ounce cup is sold for 60 cents to small consumers for a profit of (50)(0.60 – 0.30) = $15. Large consumers would obtain a net surplus of (6)(0.15) – 0.60 = 0.30 from consuming the 6-ounce cup. The large cup must be sold at a price satisfying (10)(0.15) – p  0.30. The highest such price is p = 1.20. Profit from the large consumers is (100)(1.20 – 0.50) = $70. Total profit is $85, greater than the profit in part c.

16.6

a.

The expected cost of a replacement pair is (0.5)[(0.2)(25) + (0.8)(0)] + (0.5)[(0.6)(25) + (0.4)(0)] = $10. The first term is the product of the probability of a desk worker times the expected replacement cost for a desk worker; the second is the product of the probability of an active user times the expected replacement cost for an active user. Added to the $25 cost of the original pair, the expected cost is 10 + 25 = $35.

b.

Desk workers would drop out of the market. All consumers would be active users. Total expected cost would rise to 25 (for the original pair) + (0.6)(25) (expected replacement cost) = $40. The inefficiency is that desk workers may value shoes at more than the 25 + (0.2)(25) = $30 cost of serving them, but are not served in equilibrium.

16.7

c.

Shoes sell for $25. Replacement guarantees sell for (0.6)(25) = $15.

a.

The equilibrium is for each to bid their valuation. The price paid will be $1 million unless both have high values, in which case the price will be $2 million. Expected revenue thus is (3/4)(1 million) + (1/4)(2 million) = $1.25 million.

b.

With three bidders, the price paid will be $2 million if at least two have high valuations and $1 million otherwise. The probability of at least two having high valuations is ½. You can see this by listing the 23 = 8 equally likely permutations of valuations (LHL, HHL, and so forth) and noting that half of them involve two or ore high valuations H. Expected revenue equals (1/2)(2 million) + (1/2) (1 million) = $1.5 million.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

132


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

With N bidders, expected revenue increases in N. Computing the probability of at least two high valuations is a difficult mathematical exercise that students are not expected to be able to solve. For the record, expected revenue can be shown to be N N  1  1 1  ( N  1)   (2 million)  ( N  1)  (1 million).  2   2 

16.8

c.

Expected revenue is the same from a first-price auction as from a second-price auction by the revenue-equivalence theorem.

a.

(1/2)(10,000) + (1/2)(2,000) = 6,000.

b.

If sellers value good cars at $8,000, they would not offer them for sale at the price from part a of $6,000. They would drop out of the market. Only bad cars would be sold, and the market price for them would be $2,000. If sellers value good cars at $6,000, they would be willing to offer their cars for sale at the price from part a of $6,000 (they would be indifferent between selling and not, so we can assume they sell their cars if we want). There is an equilibrium in which all cars are sold at $6,000.

16.9

a.

(1/4)(100) + (3/4)(200) = 175.

b.

200 – cL  100 and 200 – cH < 100 or together, cL ≤ 100 < cH.

c.

There is a pooling equilibrium in which both get an education. This is an equilibrium as long as the firm’s beliefs are that an uneducated worker is unproductive. By obtaining an education in this equilibrium, low-productivity workers obtain surplus 175 – cL = 175 – 50 = 125. If a low-productivity worker does not get an education, their surplus is 100 < 125. So the low-productivity worker would indeed prefer to get an education. Of course, a high-productivity worker would as well since they have a lower cost of obtaining an education. There is also a pooling equilibrium in which neither type gets an education. This is an equilibrium if the firm believes an educated worker is equally likely to be highor low-ability. There would be no return to education, and so both types would not get an education in equilibrium.

16.10

a.

This part of the problem is similar to Problem 13.5. Bertrand competition between a low- and a high-cost firm results in the low-cost firm meeting all demand at a price slightly less than the high-cost firm’s cost.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

133


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

If the incumbent firm is high-cost, in equilibrium the entrant sells at slightly less than 20 and earning a positive margin 20 – 10 on each unit sold. If the incumbent firm is low-cost, it will make all the sales at a price slightly less than 15. The entrant would earn nothing. b.

If the incumbent firm is certainly low cost, the entrant would earn nothing in the second period and would not pay even a small entry cost. If the incumbent firm is certainly high cost, the entrant would earn positive profit and would be willing to pay a small entry cost.

c.

Suppose the low-cost incumbent charged more than 20 + a in a separating equilibrium. then, if the high-cost type deviated to 20 + a, it would earn 20 + a – 20 = a > 0 from each sale in the first period, and it would earn the monopoly profit in the second. In equilibrium, the high-cost type earns the monopoly profit in the first period and nothing in the second because the entrant learns the incumbent’s type and enters when it learns this. With no discounting, the high-cost type would deviate to the low-cost price of 20 + a. Such a price cannot be part of a separating equilibrium.

[return to top]

Chapter 17: Externalities and Public Goods Purpose and Organization of the Chapter Externalities were first introduced explicitly in Chapter 11. This chapter provides a more detailed analysis of them. In defining externalities, actual physical interactions are stressed. Following standard practice, external effects that operate through the market are not termed ―externalities.‖ Principal attention in the chapter is directed toward ways of coping with externalities. The classical taxation and merger solutions are first presented. This is followed by an extended discussion of the possibilities for bargaining, ending with the Coase Theorem. When bargaining costs are high, externalities may need to be addressed through regulation, a topic that is also very briefly explored here. The second half of Chapter 17 discusses public goods. The analysis is divided into two distinct sections. The first analyzes traditional theories of public goods. Nonexclusivity and nonrivalry are stressed as the identifying features of such goods and it is shown how these features can lead to un-der-production using a Nash approach. Lindahl’s solution to the public goods issue is then illustrated and criticized. The second section of Chapter 17 concerns voting. Condorcet’s paradox is illustrated first. The discussion of voting then turns to a presentation of the ―median voter theorem‖ and an analysis of its applicability. Finally, the chapter briefly touches on issues of representative government and rent-seeking activities.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

134


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Lecture and Discussion Suggestions Two lecture suggestions might be offered for this chapter. In discussing the Coase Theorem, Meade’s bee-apple orchard example is very instructive. Several articles have re-analyzed Meade’s fable and demonstrated that, in fact, well developed markets in bee rental exist. Students seem to enjoy the bucolic triviality of these bee examples. Alternatively, one might focus on a law and economics example. The Coase Theorem can be applied to product safety (as in Application 17.3) or to the distinction between tort and contract law (the classic reference is Calabresi and Melamed (Harvard Law Review, 1972). A lecture on public goods should, I believe, focus on the theoretical material related to the allocational problems they pose. Most important for students is to understand why both nonexclusivity and nonrivalry may lead to inefficient allocations and why societies may develop financing mechanisms that may make everyone better off. Discussions of externalities might focus on either issues in environmental regulation or on the law and economics literature. For the former, students might be asked to comment on the efficiency properties of various types of taxes and to speculate about the political forces that affect their adoption. The law and economics literature is an extremely rich source of discussion material—we especially recommend the theories of optimal precaution in tort law. Instructors who have the time to discuss public goods can choose from a wealth of material from the public choice literature. Tax limitation provisions (Application 17.7) seem a particularly thoughtprovoking topic.

Cengage Supplements The following product-level supplements provide additional information that may help you in preparing your course. They are available in the Instructor Resource Center.  

Test Bank PowerPoint Slides

Chapter Objectives The following objectives are addressed in this chapter: 17.01

Provide some examples of positive and negative externalities.

17.02

Explain how a negative externality leads to an inefficient allocation of resources.

17.03

Explain how property rights and bargaining play a role in solving the externality problem.

17.04

Understand ways by which the government addresses externality problems.

17.05

Explain the characteristics of public goods.

17.06

Explain why resources are under allocated to the production of public goods.

17.07

Show how an optimal tax may be used to achieve the efficient quantity of public goods, but that it may not be practical.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

135


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

17.08

Understand how majority voting may lead to paradoxical results and policies favored by the median voter.

[return to top]

What's New in This Chapter Other than some modest changes to the discussion of environmental taxes and regulation, most of the changes in this chapter are in the applications. 

Application 17.1 is a new application drawn from the law and economics literature. It first discusses the picturesque case of Spur Industries v Del E. Webb to make the point that the causation of externalities can run both ways. This insight is then applied to the concept of ―environmental justice‖. The graphical discussion of product safety in Application 17.3 is greatly improved through the use of multi-colors.

Application 17.4 provides an extended section on climate change economics and provides a brief discussion of the ―Green New Deal‖.

The discussion of tax limitations in Application 17.7 has been expanded to discuss the special case of funding schools.

[return to top]

Additional Resources Robert Bullard is often credited with being the ―father‖ of the environmental-justice movement. A provocative interview with him is provided by https://www.pbs.org/wnet/amanpour-and-company/video/robert-bullard-howenvironmental-racism-shapes-the-us/. Most economists favor some form of carbon tax. Again, the Econofact website provides an even-handed discussion: https://econofact.org/the-carbon-tax-video. ―Modern Monetary Theory‖ argues that the issue of paying for public goods is unnecessary because the government can print any amount of money needed. Stephanie Kelton presents the basic ideas in: https://www.cnbc.com/video/2019/03/01/stephanie-kelton-explains-modernmonetary-theory.html. Larry Summers thinks this is all nonsense: https://www.youtube.com/watch?v=g6iB6QzuQ70.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

136


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

[return to top]

Solutions to End of Chapter Problems Students have access to solutions for the odd numbered problems as well as video problem walkthroughs for problems 1 and 3.

17.1

a.

MC = 0.4q. P = $20. Set P = MC. 20 = 0.4q, q = 50. b.

MCS = 0.4q + 0.1q = 0.5q. Set P = MCS. 20 = 0.5q. q = 40. At optimal production level of q = 40, the marginal cost of production is MC = 0.4q = 0.4(40) = 16, so the excise tax t = 20 – 16 = $4.

17.2

c.

The graph shows the optimal tax in this widget market.

a.

Fishers will arrange themselves so that the average catch on each lake is the same. Since the average on lake Y is always 5, it must be 5 on lake X also. So

FX  10  0.5LX  5 so LX  10, LY  10 LX 2

Total catch is FX = 10(10) – 1/2 (10) = 50 FY = 5(10) = 50 b.

Total = FX + FY = 100.

To maximize the catch, should set marginal productivities equal MPX = 10 – LX = MPY = 5 or LX = 5 so LY = 15. Total output can be calculated as 2

FX = 10(5) – 1/2(5) = 37.5 FY = 5(15) = 75 Total output = 112.5

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

137


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

c.

The license fee should be set so that the average catch on lake X minus the fee is equal to 5 (the catch on Y) when 5 fishers use the lake (the optimal number). In this way, fishers themselves will opt for the correct allocation.

FX  fee  10  0.5LX  fee  5 when LX  5 LX Hence, 7.5 – fee = 5 fee = 2.5, though how to collect a half fish fee is unpleasant to contemplate. 17.3

AC = MC = 10,000/well. a.

Produce where revenue/well = 10,000 = 100q = 50,000 – 100N. N = 400. There is an externality here because drilling another well reduces output in all wells.

b.

Produce where MVP = MC of well. Total value = 50, 000 N  100 N 2 . MVP = 50,000 – 200N = 10,000. N = 200.

c.

Let Tax = T. Want revenue/well – T = 10,000 when N = 200. At N = 200, average revenue per well = 30,000. Charge T = 20,000.

17.4

a.

Suppose that equipment causes expected damage of d. With full information, the equilibrium would be independent of legal rules. Suppose demanders incurred all costs – the equilibrium is shown by P*,Q* in the figure. Now if suppliers are required to pay the damage costs, the supply curve would shift up by d as would the demand curve (because buyers now have their costs reimbursed). Equilibrium would stay at Q*.

b.

There would be no change because all parties fully expect Mr. Coyote to be careless and use that information in assessing d.

c.

In this case, the size of d would depend on the choice of legal regime. Now retaining the same equilibrium is unlikely. One would have to look into the reasons why Mr. Coyote behaves differently and how he gains utility from doing so in order to analyze the situation thoroughly.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

138


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

17.5

17.6

d.

If Acme has a monopoly we must look at how legal liability affects the marginal revenue-marginal cost intersection. As before, d directly affects marginal cost. But d affects the demand curve for equipment directly, the marginal revenue curve only indirectly. Only in special cases will the marginal revenue and marginal cost curves both shift by d as they did in part a. Which legal regime has the larger output will depend on the relative sizes of the shifts.

a.

For profit maximization, set P = MC, 50 = 30 + 0.5Q. Hence, Q = 40 hives. There will be enough bees only to pollinate 10 acres.

b.

Orchard owner would pay up to $25 per hive. A $20 subsidy would result in total receipts per hive of 70 and profit maximization would dictate 70 = 30 + 5Q or Q = 80—enough hives to pollinate the entire 20 acres.

a.

Setting MB = MC yields 100 – R = 20 + R or R = 40.

b.

The fee should be set so that farmers choose R = 40. So, Fee = MC = 20 + R = 20 + 40 = 60. A fee of 60 for each percent not reduced would prompt farmers to choose R = 40. It is cheaper to pay the fee than reduce methane by 41 percent, however, since that would cost 61.

c.

With a mandate of a 40 percent reduction, the average reduction will be 40 percent. 2

2

MC1 = 20 + /3(40) = 46 /3 MC2 = 20 + 2(40) = 100. Total cost is given area under MC up to the required level. For farm 1 this amounts to 2

1

1

20  40 + (46 /3 – 20)(40) = 800 + 533 /3 = 1,333 /3. For farm 2, this is 1

20  40 + /2(100 – 20)(40) = 800 + 1,600 = 2,400. 1

So total costs of achieving the 40 percent reduction are 3,733 /3. d.

2

With a fee of 60, farm 1 sets 60 = 20 + /3R, and calculates R1 = 60. Farm 2 calculates 60 = 20 + 2R2 or R2 = 20. Again, the average reduction is 40. Total costs now for farm 1 are 20  60 + 1/2(60 – 20)(60) = 1,200 + 1,200 = 2,400 and for farm 2 20  20 + 1/2(60 – 20)(20) = 400 + 400 = 800 so total costs are 3,200. The fee achieves the same average reduction at a much lower total cost.

e.

The fee procedure recognizes the differential costs of methane reduction. By having the low cost farm (farm 1) make relatively greater reductions, overall costs are reduced.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

139


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

17.7

a.

Marginal valuation for person A = P = 100 – qA; for B, Marginal valuation = P = 200 – qB. Because of the public good nature of mosquito control these should be added "vertically." Marginal value = 300 – 2q (since qA = qB) Set this equal to marginal cost of 50, gives q = 125.

b.

Free rider problem could result in having no production. Each person would let the other do it.

c.

Total Cost = 50  125 = $6,250. Area under demand curve for A = $4,688—for B = $17,188. One solution would be to share costs in proportion to these values.

17.8

a.

Total Net Benefits = $340 > $300. Under equal sharing A and B would vote for the project, C against it. Net benefits for person A = 50, for person B = 40, and for person C = –50.

b.

Now net benefits fall short of costs, but A and B would still vote for the project.

c.

In case a, Person C is willing to pay up to $50 as a bribe. Since B’s net gain from project is 40, C could bribe B to vote with a bribe of more than 40. But A could counteract such a bribe with a bribe of 10 or more to B. In case b, the situation is different. Now C will pay up to $75 in bribes. If all went to B, net gains from the project ($15) plus A’s maximum bribe ($25) would be insufficient to deter B from accepting the bribe.

17.9

17.10

a.

The pool is nonrival (by assumption), but exclusion is possible.

b.

Building the pool would generate $6,000 per day in economic value at a cost of $5,000 per day. It would be efficient to build it.

c.

A price of $3 would generate $3,000 in revenue, a price of $2 would generate $4,000, and a price of $1 would generate $3,000. Obviously a price of $0 would generate no revenue (though it would be equal to marginal cost). No single price policy is viable.

d.

If it were possible to differentiate between families’ willingness to pay, a policy of perfect price discrimination would be efficient – each family could be charged their maximum willingness to pay.

e.

Now pool attendance must be limited to 2,000 in order to maximize economic value. If families can be differentiated, a two-price policy charging $3 and $2 would generate just enough revenue to cover the cost of the pool. Group 3 families would be barred from the pool even though they would obtain some value from it. Admitting them would cause a negative externality on groups 1 and 2 that would more than compensate for their gain.

a.

Since Q = 200 – 100P, profit maximizing price is 0.5(2 + .50) = 1.25. At this price, Q = 75, p = (1.25 – 0.5)(75) = 56.25. Firm will be willing to pay up to this amount per period as a bribe.

b.

The bribe is only a transfer from monopoly profits to the government official. This is not a welfare cost.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

140


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

c.

The true welfare cost is the loss of consumer surplus from the monopoly itself. The value of this loss is 0.5(1.25 – 0.5)(75) = 28.125. Notice here that the output restriction from the monopoly is 75 since the competitive output would be Q = 150.

[return to top]

Chapter 18: Behavioral Economics Purpose and Organization of the Chapter This chapter surveys recent developments in behavioral economics. In contrast to the neoclassical perspective occupying the rest of the book, behavioral economics studied in this chapter does not take as its point of departure the assumption that economic agents make perfectly rational decisions. Rather, behavioral economics seeks to under-stand how decisions are made by real-world (and thus possibly imperfect) economic agents. It seeks to understand the extent of possible mistakes, if the mistakes show any predictable patterns, or if what seem like mistakes are perhaps maximization of different social goals than simply selfish payoff maximization. In brief, behavioral economics seeks to integrate psychology and economics. This is exciting material because it has been an active area of research over the past several decades. Being relatively young, the subject is less developed than neoclassical economics and the ultimate value of the approach is still the subject of considerable debate. There is too much material to offer a comprehensive survey in this chapter, so it only covers a few of the highlights. The chapter is organized around three broad limits to selfish maximizing behavior studied in three separate sections: limits to cognitive ability, limits to willpower, and limits to self-interest. The section on limits to cognitive ability studies possible mistakes made in complicated environments in which uncertainty, strategy, or overwhelming numbers of choices play a role. The section on limits to willpower presents the widely used model of hyperbolic discounting. According to this model, the marginal rate of substitution between payoffs in two periods is one thing in the planning stage but another when one is actually living in the periods, leading to time inconsistency. The section on limits to self-interest starts with the idea of altruism, which is easy to model in the standard neoclassical framework, but then goes on to more complicated social preferences such as fair-ness, reciprocity, and envy. These alternative social preferences have a big impact on how games will be played, so the section is closely connected with Chapter 6. Behavioral economics introduces a new role for government intervention in the market, a paternalistic role leading the government to try to correct mistakes made by market participants. Needless to say, this perspective has been strongly challenged by some neoclassical economists. We mention the debate here and some of the points raised without taking sides. The continued debate as well as the continued revision of our understanding of

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

141


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

behavioral economics may be a bit frustrating to students who want ―the‖ answer, but on the other hand gives them a window into economics as it really is, an evolving science.

Lecture and Discussion Suggestions There are a number of valid approaches to the material in this chapter. One would be to omit it entirely. Instructors were already pressed for time be-fore the addition of this chapter, so there may well be no time at all for it. The chapter is self-contained, so there is no loss to the rest of the material if it is eliminated. Another reason for omitting the chapter is if the instructor prefers to stick to a single, elegant model of economic behavior, the neoclassical model, and not ―muddy the waters‖ with alternative models, especially if the best alternative model is still far from settled. Again, the book fits perfectly with this approach since it is only this last chapter that departs from the neoclassical approach so can be omitted seamlessly. For instructors who want to include at least some coverage of behavioral economics, there are a number of approaches. The chapter can be covered as a self-contained unit, perhaps toward the end of the term. The students might be assigned a popular book on behavioral economics, for ex-ample, Thaler and Sunstein’s recent popular book, Nudge, as outside reading, perhaps as the subject of a term paper to be worked on fairly independently. Another approach would be to sprinkled the material throughout the term to enrich various other topics. The material on limits to cognitive ability could be mentioned when the topic of uncertainty (Chapter 5) is covered, noting that decisions under uncertainty are complex and may be the source of biases. The material on limits to self-interest could be added to a module on game theory (Chapter 6), discussing for example how the predictions of Nash equilibrium would change if players care about fairness in addition to their own monetary payoff. The material on hyperbolic dis-counting could enrich the discussion of standard discounting in Chapter 15.

Cengage Supplements The following product-level supplements provide additional information that may help you in preparing your course. They are available in the Instructor Resource Center.  

Test Bank PowerPoint Slides

Chapter Objectives The following objectives are addressed in this chapter: 18.01

Understand four factors that limit our cognitive power.

18.02

Explain how limits to our willpower can lead to bad decisions.

18.03

Provide examples of how we can use commitment strategies to minimize the negative consequences of limited willpower.

18.04

Understand how limits on our self-interest change the decisions we make.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

142


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

18.05

Explain the pros and cons of government intervention to correct failures in decision making.

[return to top]

What's New in This Chapter The chapter is similar to that in the previous edition. The following improvements were incorporated. 

The prose has been streamlined in places.

Application 18.1: Household Finance has a new Policy Challenge, asking students to consider tradeoffs in the policies enacted by the U.S. Consumer Financial Protection Bureau (CFPB), created in 2010, to protect consumers possibly plagued by behavioral biases from exploitation in financial markets.

Application 18.2: Cold Movie Openings adds a passage on how disruption of movie production due to the COVID-19 pandemic might distort release dates, changing the amount of time for critics to review movies before release, changing the inference consumers draw from the lack of critical reviews.

Application 18.3: Going for It on Fourth Down notes that economists’ recommendations for altering the typical NFL playbook has penetrated the coaching ranks.

[return to top]

Additional Resources The American Economic Association posts videos of a diverse set of academic economists discussing their relevant research on the website http://diversifyingecon.org/index.php/Videos_on_economists_and_their_research. Two videos relate to behavioral economics. Antonio Alonso Arechar discusses the role of honesty in promoting cooperation. https://vimeo.com/250480329. Keith Chen speculates on whether awkward future-tense constructions in the grammar of certain languages reduces savings and other future-regarding behaviors by people speaking those languages. https://www.ted.com/talks/keith_chen_could_your_language_affect_your_ability _to_save_money?language=en.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

143


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

[return to top]

Solutions to End of Chapter Problems Students have access to solutions for the odd numbered problems as well as a video problem walkthrough for problem 8.

18.1

a.

The first prize is 100,000d, and the second is 2d–1/100 (in dollars). b. million $

second prize

first prize

d

18.2

c.

The curves cross between day 29 and 30. The first prize is better for shorter time spans and the second prize for longer time spans.

a.

E(U ( A))  (0.89  1,000)  (0.1 5,000)  (.01 0)  35.2. E (U ( B))  1,000  31.6.

E(U (C ))  (0.11 1,000)  (0.89  0)  3.5.

E(U ( D))  (0.1 5,000)  (0.9  0)  7.1.

18.3

b.

Prefer A.

c.

Prefer D.

d.

Consistently prefer the gamble that involves a lower probability of winning anything but some change of winning the big prize of $5,000. Experimental evidence is not this consistent, for some reason being attracted to the sure thing in gamble B.

a. Both play Rat. b.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

144


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Now there are two equilibria: both play Rat and both play Silent. c.

Both play Rat, as in part a again. 18.4

a.

In one Nash equilibrium, both choose Astro, and in the other both choose Bio.

b.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

145


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

Now players prefer ―discoordinating.‖ In the two pure-strategy Nash equilibria, one player choose Astro and the other Bio. c.

The game has one Nash equilibrium: A chooses Bio, and B chooses Astro. 17.5

a.

Player 1 makes a low offer; player 2 accepts either offer.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

146


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

b.

1 Low

Even

2 A 9.5, 5.5

2 R

A

0, 0

7.5, 7.5

R

0, 0

The equilibrium is the same as in part a. c.

1 Low

Even

2 A 10, 10

2 R 0, 0

A 10, 10

R

0, 0

Now, besides the equilibrium in part a, there is another one in which player 1 makes an even offer and 2 accepts either offer.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

147


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

d.

1 Low

Even

2 A 1, 9

2 R 0, 0

A 5, 5

R

0, 0

In equilibrium, 1 offers an even split and 2 accepts any offer. Paradoxically, 2 gains a higher monetary payoff but lower utility than if he or she received the low offer. 18.6

18.7

18.8

a.

The number of combinations of 24 jars taken two at a time is 24 × 23 ÷ 2 = 276. To see this calculation, there are 24 ways to choose the first jar and 23 ways left to choose the second, but order within the pair doesn’t matter, so we have to divide by 2.

b.

In the first group, there are four to choose from, resulting in 4 × 3 ÷2 = 6 comparisons. In the second group there are two to choose from, resulting in 2 × 1 ÷ 2 = 1 comparison. Proceeding in this way through all the groups, there are 6 + 1 + 1 + 1 + 10 + 0 + 1 + 1 + 6 = 27 comparisons, leaving 9 to compare across groups. There are 9 × 8 ÷2 = 36 comparisons to be made among the 9 that are best from each group. In all, 27 + 36 = 63 comparisons need to be made. This is a significant reduction from the 276 from part a.

a.

Darrick plans to study, and also carries out his plan, if s < b.

b.

Jamila plans to study if s < b, but she only follows through if s < wb.

a.

Present discounted value at planning stage (period 1) of Mr. Consistent’s utility flow from exercise = (0.5)(–100) + (0.25)(250) = 12.5. Since this value is positive, he would plan to exercise. At the stage when the exercise needs to be undertaken (period 2), the present discounted value = (1)( –100) + (0.5)(250) = 25. Since this value is positive, he would carry out the plan.

b.

Present discounted value at planning stage (period 1) of Mr. Hyperbolic’s utility flow from exercise = (0.35)( –100) + (0.25)(175) = 8.75. Since this value is positive, he would plan to exercise. At the stage when the exercise needs to be undertaken (period 2), the present discounted value = (1)( –100) + (0.35)(250) = –12.5. Since this value is negative, he would not exercise, contrary to his plan.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

148


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

18.9

18.10

c.

As seen in b, he obtains a present discounted value of –12.5 if he exercises, so x ≥ 12.5 would induce him to exercise.

a.

Pete’s expected utility from gamble A is 10,000 + (1/2)(250) – (1/2)(2)(100) = 10,025 and from gamble B is 10,030, so he chooses B.

b.

Pete’s expected utility from gamble C is 10,100 + (1/2)(150) – (1/2)(2)(200) = 9,975 and from gamble D is 10,100 – (2)(70) = 9,960, so he chooses C.

c.

A yields the same wealth levels as C. B yields the same as D.

a.

Setting QS = QD yields P/2 = 100 - 2P, implying P* = 40, Q* = 20, PS* = 400, CS* = 100, W* = 500. (The figure in below part b shows the triangles whose areas equal give PS* and CS*.)

b.

~ ~ Setting Q S  Q D (where Q D is mistaken rather than true demand) yields P/2 = 200 - 2P, implying P** = 80 and Q** = 40. The deadweight loss triangle is shown in the figure below. Relative to the efficient outcome with Q* = 20, an excess of 20 units are produced (viewed from the perspective of the true demand curve). The cost of these units is given by the area under S and the consumer surplus they generate is given by the area under D. The difference is deadweight loss, the area of the shaded triangle. This area is (1/2)(20)(50) = 500 = DWL.

c.

A tax of 50 will shift the mistaken demand curve D′ down to now overlap with the true one, D.

d.

Imposing a tax of 50 in a market in which true demand is D′ leads to the shaded deadweight loss triangle in the figure below. This area is (1/2)(20)(50) = 500 = DWL. So if the government is mistaken, it can generate deadweight loss of the same magnitude as from consumer mistakes.

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

149


Instructor Manual: Nicholson/Snyder, Intermediate Microeconomics, 13e

[return to top]

© 2022 Cengage. All Rights Reserved. May not be scanned, copied or duplicated, or posted to a publicly accessible website, in whole or in part.

150


Turn static files into dynamic content formats.

Create a flipbook
Issuu converts static files into: digital portfolios, online yearbooks, online catalogs, digital photo albums and more. Sign up and create your flipbook.