We now impose the boundary conditions in (6.11) and (6.12) and solve for a 1 and a 2 as follows: m 1 e 2m 1T m 2 e (m 1+m 2 )T. (6.

Similar documents
CHAPTER 9 MAINTENANCE AND REPLACEMENT. Chapter9 p. 1/66

CHAPTER 7 APPLICATIONS TO MARKETING. Chapter 7 p. 1/53

Solution by the Maximum Principle

CHAPTER 3 THE MAXIMUM PRINCIPLE: MIXED INEQUALITY CONSTRAINTS. p. 1/73

CHAPTER 2 THE MAXIMUM PRINCIPLE: CONTINUOUS TIME. Chapter2 p. 1/67

Differential Games, Distributed Systems, and Impulse Control

Optimal Control. Macroeconomics II SMU. Ömer Özak (SMU) Economic Growth Macroeconomics II 1 / 112

Graphing Systems of Linear Equations

Minimum Fuel Optimal Control Example For A Scalar System

Contents CONTENTS 1. 1 Straight Lines and Linear Equations 1. 2 Systems of Equations 6. 3 Applications to Business Analysis 11.

Part A: Answer question A1 (required), plus either question A2 or A3.

Optimal control theory with applications to resource and environmental economics

Mathematical Foundations -1- Constrained Optimization. Constrained Optimization. An intuitive approach 2. First Order Conditions (FOC) 7

Mathematics Level D: Lesson 2 Representations of a Line

Problem 1 Cost of an Infinite Horizon LQR

10 Exponential and Logarithmic Functions

Doug Clark The Learning Center 100 Student Success Center learningcenter.missouri.edu Overview

Electronic Companion to Tax-Effective Supply Chain Decisions under China s Export-Oriented Tax Policies

The Method of Substitution. Linear and Nonlinear Systems of Equations. The Method of Substitution. The Method of Substitution. Example 2.

Parametric Equations, Function Composition and the Chain Rule: A Worksheet

Review Assignment II

Technical Note: Capacity Expansion and Cost Efficiency Improvement in the Warehouse Problem. Abstract

problem. max Both k (0) and h (0) are given at time 0. (a) Write down the Hamilton-Jacobi-Bellman (HJB) Equation in the dynamic programming

MS-E2140. Lecture 1. (course book chapters )

PART III. INVENTORY THEORY

17 Exponential and Logarithmic Functions

FACULTY OF ARTS AND SCIENCE University of Toronto FINAL EXAMINATIONS, APRIL 2012 MAT 133Y1Y Calculus and Linear Algebra for Commerce

A Proof of the EOQ Formula Using Quasi-Variational. Inequalities. March 19, Abstract

Study Unit 3 : Linear algebra

Queen s University. Department of Economics. Instructor: Kevin Andrew

DYNAMIC LECTURE 5: DISCRETE TIME INTERTEMPORAL OPTIMIZATION

Optimal Backlogging Over an Infinite Horizon Under Time Varying Convex Production and Inventory Costs

C.3 First-Order Linear Differential Equations

Alberto Bressan. Department of Mathematics, Penn State University

Order book modeling and market making under uncertainty.

1 Functions, Graphs and Limits

Final Exam Review. MATH Intuitive Calculus Fall 2013 Circle lab day: Mon / Fri. Name:. Show all your work.

II. Analysis of Linear Programming Solutions

Section 4.2 Logarithmic Functions & Applications

Lesson 1: Multiplying and Factoring Polynomial Expressions

Function: State whether the following examples are functions. Then state the domain and range. Use interval notation.

Chapter Introduction. A. Bensoussan

MS-E2140. Lecture 1. (course book chapters )

Answers to Problem Set Number 02 for MIT (Spring 2008)

Mathematical Economics. Lecture Notes (in extracts)

Dynamical Systems. August 13, 2013

Active Maths 4 Book 1: Additional Questions and Solutions

1 Numbers, Sets, Algebraic Expressions

END3033 Operations Research I Sensitivity Analysis & Duality. to accompany Operations Research: Applications and Algorithms Fatih Cavdur

CHAPTER FIVE. Solutions for Section 5.1. Skill Refresher. Exercises

Minimal Forecast Horizons: An Integer Programming Approach

Question 1. (8 points) The following diagram shows the graphs of eight equations.

3.4 The Fundamental Theorem of Algebra

Review questions for Math 111 final. Please SHOW your WORK to receive full credit Final Test is based on 150 points

CHAPTER FIVE. g(t) = t, h(n) = n, v(z) = z, w(c) = c, u(k) = ( 0.003)k,

NOTES ON CALCULUS OF VARIATIONS. September 13, 2012

Chapter 5. Pontryagin s Minimum Principle (Constrained OCP)

Optimal Control of Differential Equations with Pure State Constraints

Replacing the a in the definition of the derivative of the function f at a with a variable x, gives the derivative function f (x).

Linear-Quadratic Optimal Control: Full-State Feedback

Deterministic Dynamic Programming

IM3 Unit 1 TEST - Working with Linear Relations SEP 2015

minimize x subject to (x 2)(x 4) u,

Theory and Applications of Optimal Control Problems with Time-Delays

EXTREMAL ANALYTICAL CONTROL AND GUIDANCE SOLUTIONS FOR POWERED DESCENT AND PRECISION LANDING. Dilmurat Azimov

Solvability in Discrete, Nonstationary, Infinite Horizon Optimization

Final Exam January 26th, Dynamic Programming & Optimal Control ( ) Solutions. Number of Problems: 4. No calculators allowed.

3. Replace any row by the sum of that row and a constant multiple of any other row.

On Your Own. Applications. Unit 1. 1 p = 7.5n - 55, where n represents the number of car washes and p represents the profit in dollars.

New Keynesian Model Walsh Chapter 8

MATH 115 SECOND MIDTERM EXAM

1. (7pts) Find the points of intersection, if any, of the following planes. 3x + 9y + 6z = 3 2x 6y 4z = 2 x + 3y + 2z = 1

Lecture 1: Overview, Hamiltonians and Phase Diagrams. ECO 521: Advanced Macroeconomics I. Benjamin Moll. Princeton University, Fall

2. What are the zeros of (x 2)(x 2 9)? (1) { 3, 2, 3} (2) { 3, 3} (3) { 3, 0, 3} (4) {0, 3} 2

Chapter 1 Review Applied Calculus 31

MTH 65-Steiner Exam #1 Review: , , 8.6. Non-Calculator sections: (Solving Systems), Chapter 5 (Operations with Polynomials)

Section 4.1 Solving Systems of Linear Inequalities

Advanced Microeconomic Analysis, Lecture 6

Lecture 6: Sections 2.2 and 2.3 Polynomial Functions, Quadratic Models

Game Theory and Algorithms Lecture 2: Nash Equilibria and Examples

Introduction to Operations Research Economics 172A Winter 2007 Some ground rules for home works and exams:

Basic Techniques. Ping Wang Department of Economics Washington University in St. Louis. January 2018

Chapter 4. Systems of Linear Equations; Matrices. Opening Example. Section 1 Review: Systems of Linear Equations in Two Variables

Page 1 of 10 MATH 120 Final Exam Review

Maintaining Mathematical Proficiency

Worst case analysis for a general class of on-line lot-sizing heuristics

Chapter 1 Linear Equations

OPTIMAL CONTROL OF A PRODUCTION SYSTEM WITH INVENTORY-LEVEL-DEPENDENT DEMAND

MAGYAR NEMZETI BANK MINI-COURSE

Relativistic Mechanics

Checkpoint 1 Simplifying Like Terms and Distributive Property

The Necessity of the Transversality Condition at Infinity: A (Very) Special Case

Consistent and Dependent

PreCalc 11 Chapter 1 Review Pack v1

Lecture 6: Discrete-Time Dynamic Optimization

Given the table of values, determine the equation

Math 2 Variable Manipulation Part 6 System of Equations

Solutions to Math 53 Math 53 Practice Final

Math 112 Group Activity: The Vertical Speed of a Shell

due Thursday, August 25, Friday, September 2, 2016 test Prerequisite Skills for Algebra II Advanced

Transcription:

158 6. Applications to Production And Inventory For ease of expressing a 1 and a 2, let us define two constants b 1 = I 0 Q(0), (6.13) b 2 = ˆP Q(T) S(T). (6.14) We now impose the boundary conditions in (6.11) and (6.12) and solve for a 1 and a 2 as follows: a 1 = b 2e m 1T m 2 b 1 e (m 1+m 2 )T m 1 e 2m 1T m 2 e (m 1+m 2 )T, (6.15) a 2 = b 1 m 1 e 2m 1T b 2 e m 1T m 1 e 2m 1T m 2 e (m 1+m 2 )T. (6.16) If we recall that m 1 is negative and m 2 is positive, then when T is sufficiently large so that e m 1T and e 2m 1T are negligible, we can write a 1 b 1, (6.17) a 2 b 2 m 2 e m 2T. (6.18) Note that for a large T, e m 2T is close to zero and, therefore, a 2 is close to zero. However, the reason for retaining the exponential term in (6.18) is that a 2 is multiplied by e m 2t in (6.12) which, while small when t is small, becomes large and important when t is close to T. With these values of a 1 and a 2 and with (6.4), (6.11), and (6.12), we now write the expressions for I, P, and λ. We will break each expression into three parts: the first part labeled Starting Correction is important only when t is small; the second part labeled Turnpike Expression is significant for all values of t; and the third part labeled Ending Correction is important only when t is close to T. Starting Correction Turnpike Expression Ending Correction I = (b 1 e m 1t )+ (Q)+ ( b 2 m 2 e m 2(t T) ) (6.19) P = (m 1 b 1 e m 1t )+ ( Q + S)+ (b 2 e m 2(t T) ) (6.20) λ = c(m 1 b 1 e m 1t )+ c( Q + S ˆP)+ c(b 2 e m 2(t T) ) (6.21) Note that if b 1 = 0, which means I 0 = Q(0), then there is no starting correction. In other words, I 0 = Q(0) is a starting inventory that causes the solution to be on the turnpike initially. In the same way, if b 2 = 0,

6.1. A Production-Inventory System 159 then the ending correction vanishes in each of these formulas, and the solution stays on the turnpike until the end. Expressions (6.19) and (6.20) represent approximate closed-form solutions for the optimal inventory and production functions I and P as long as S is such that the special particular integral Q can be found explicitly. For such examples of S, see Section 6.1.5. 6.1.3 The Infinite Horizon Solution It is important to show that this solution also makes sense when T. In this case it is usual to assume that ρ > 0, and to show that the limit of this finite horizon solution as T also solves the infinite horizon problem. Note that as T, the ending correction disappears because a 2 defined in (6.16) becomes 0. We now have λ(t) = c[m 1 b 1 e m 1t + Q + S ˆP]. (6.22) If S is bounded, then Q is bounded, and therefore, lim t λ(t) is bounded. Then for ρ > 0, lim t e ρt λ(t) = 0. (6.23) By the sufficiency of the maximum principle conditions (Section 2.4), it can be verified that the limiting solution I(t) = b 1 e m 1t + Q, P(t) = m 1 b 1 e m 1t + Q + S (6.24) is optimal. If I(0) = Q(0), the solution is always on the turnpike. Note that the triple {I, P, λ} = {Q, Q + S, c( Q + S ˆP)} represents a nonstationary turnpike. If I(0) Q(0), then b 1 0 and the expressions (6.24) imply that the path of inventory and production only approach but never attain the turnpike. Note that the solution does not satisfy the sufficiency transversality condition when ρ = 0. In this case, all solutions give an infinite value (which is the worst value since we are minimizing) for the cost objective function. The limiting solution, as ρ 0, is the Golden Path defined in Section 3.5, and it could be deemed to be the optimal infinite horizon solution for the undiscounted case.

6.1. A Production-Inventory System 161 Figure 6.1: Optimal Production and Inventory Levels in Figure 6.1. The time ˆt shown in the figure is the time at which ˆP + λ(ˆt)/c = 0. By (6.28), if I 0 = Q S/m 1, then P(0) = 0. This suggests that I(ˆt) = Q S m 1. (6.29) For t < ˆt, we have I = S I = I 0 St, (6.30) λ = ρλ + h(i Î), λ(ˆt) = c ˆP. (6.31) We can substitute I 0 St for I in equation (6.31) and solve for λ. Note that we can easily obtain ˆt as I 0 Sˆt = Q S m 1 ˆt = I 0 Q S + 1 m 1. (6.32) For t > ˆt, the solution is given by (6.19), (6.20), and (6.21) with t replaced by (t ˆt) and b 1 = I(ˆt) Q = S/m 1. The solution shown in Figure 6.1 represents a situation where there is excess initial inventory, i.e., I 0 > Q S/m 1. For I 0 Q S/m 1, the

6.2. Continuous Wheat Trading Model 165 Figure 6.4: Solution of Example 6.1 with I 0 = 30 of this model to one having two control variables appears in Ijiri and Thompson (1972). 6.2.1 The Model We introduce the following notation: T = the horizon time, x(t) = the cash balance in dollars at time t, y(t) = the wheat balance in bushels at time t, v(t) = the rate of purchase of wheat in bushels per unit time; a negative purchase means a sale, p(t) = the price of wheat in dollars per bushel at time t, r = the constant positive interest rate earned on the cash balance, h(y) = the cost of holding y bushels per unit time. In this section we permit x and y to go negative, meaning borrowing of money and short-selling of wheat are allowed. In the next section we disallow short-selling of wheat.

6.2. Continuous Wheat Trading Model 167 In Exercise 6.6 you are asked to provide the interpretation of this optimal policy. Equations (6.39), (6.40), (6.47), and (6.48) determine the two-point boundary value problem which usually requires a numerical solution procedure. In the next section we assume a special form for the storage function h(y) to be able to obtain a closed-form solution. 6.2.3 Complete Solution of a Special Case For this special case we assume h(y) = 1 2 y, r = 0, x(0) = 10, y(0) = 0, V 1 = V 2 = 1, T = 6, and 3 for 0 t 3, p(t) = (6.49) 4 for 3 < t 6. We will apply the maximum principle developed in Chapter 2 to this problem even though h(y) is not differentiable at y = 0. The answer we obtain can be obtained rigorously by using the maximum principle for models involving nondifferentiable functions discussed, e.g., in Clarke (1989, Chapter 4) and Feichtinger and Hartl (1985b, 1986, Appendix A.3). For this case with r = 0, we have λ 1 (t) = 1 for all t from (6.46) so that the TPBVP is ẋ = 1 y pv, x(0) = 10, 2 (6.50) ẏ = v, y(0) = 0, (6.51) λ 2 (t) = 1 2 sgn (y), λ 2(6) = 4. (6.52) For this simple problem it is easy to guess a solution. From the fact that λ 1 = 1, the optimal policy (6.48) reduces to v (t) = bang[ 1, 1; λ 2 (t) p(t)]. (6.53) The graph of the price function is shown in Figure 6.5. Since p(t) is increasing, short-selling is never optimal. Since the storage cost is 1/2 unit per unit time and the wheat price jumps by 1 unit at t = 3, it never pays to store wheat more than 2 time units. Because y(0) = 0, we have v (t) = 0 for 0 t 1. This obviously must be singular control. Suppose we start buying wheat at t > 1. From (6.53) the rate of buying

168 6. Applications to Production and Inventory Figure 6.5: The Price Trajectory (6.49) is 1; clearly buying will continue at this rate until t = 3, and no longer. In order not to lose money because of storing wheat, it must be sold within 2 time units of its purchase. Clearly we should start selling at t = 3 + at the maximum rate of 1, and continue until a last sale time t. In order to sell exactly all of the wheat purchased, we must have 3 t = t 3. (6.54) Thus, v (t) = 0 in the interval [t, 6], which is also singular control. With this policy, y(t) > 0 for all t (t, t ). From (6.52), λ 2 = 1/2 in the interval (t, t ). In order to have a singular control in the interval [t, 6], we must have λ 2 (t) = 4 in that interval. Also, in order to have singular control in [0, t ], we must have λ 2 (t) = 3 in that interval. We can now conclude that t t = 2, (6.55) and therefore t = 2 and t = 4. Thus from (6.52) and (6.53), 3, 0 t 2, λ 2 (t) = 2 + t/2, 2 t 4, (6.56) 4, 4 t 6.

6.2. Continuous Wheat Trading Model 169 We can now sketch graphs for λ 2 (t), v (t), and y (t) as shown in Figure 6.6. In Exercise 6.11 you are asked to show that these trajectories are optimal by verifying that the maximum principle necessary conditions hold and that these conditions are also sufficient. Figure 6.6: Adjoint Variable, Optimal Policy and Inventory in the Wheat Trading Model

6.2. Continuous Wheat Trading Model 171 where µ 1, µ 2, and η satisfy the complementary slackness conditions: µ 1 0, µ 1 (v + 1) = 0, (6.63) µ 2 0, µ 2 (1 v) = 0, (6.64) η 0, ηy = 0, η 0. (6.65) Furthermore, the optimal trajectory must satisfy L v = λ 2 pλ 1 + µ 1 µ 2 + η = 0. (6.66) With r = 0, we get λ 1 = 1 as before, and λ 2 = L y = 1/2, λ 2(3) = 4. (6.67) From this we see that λ 2 is always increasing except possibly at a jump time. Let ˆt be a time such that there is no jump in the interval (ˆt, 3]. Then, λ 2 (t) = t/2 + 5/2 for ˆt t 3, (6.68) and the optimal control from (6.60) or (6.61) is v = 1, i.e., we buy wheat at the maximum rate of 1, so long as λ 2 (t) > p(t). The smallest possible value of ˆt is obtained by equating λ 2 (t) = p(t); thus t/2 + 5/2 = 2t + 7 ˆt=1.8. (6.69) Since p(t) is decreasing at the start of the problem, it appears that selling at the maximum rate of 1, i.e., v = 1, should be optimal at the start. Since the beginning inventory is y(0) = 1, selling at the rate of 1 can continue only until t = 1, at which time the inventory y(1) becomes 0. Suppose that we do nothing, i.e., v (t) = 0 in the interval (1, 1.8]. Then, t = 1 is an entry time (see Sections 4.1.1 and 4.2) and t = 1.8 is not an entry time. Hence, as in Example 4.3, λ 2 (t) is continuous at t = 1.8, and therefore λ 2 (t) is given by (6.68) in the interval [1, 3], i.e., λ 2 (t) = t/2 + 5/2 for 1 t 3. (6.70) Using (6.66) with λ 1 = 1 in the interval (1, 1.8] and v = 0 so that µ 1 = µ 2 = 0, we have λ 2 p + µ 1 µ 2 + η = λ 2 p + η = 0,

172 6. Applications to Production and Inventory and consequently η(t) = p(t) λ 2 (t)= 5t/2 + 9/2 and η = 5/2 0, t (1, 1.8]. (6.71) Since h t = 0, the jump condition in (4.29) for the Hamiltonian at τ = 1 reduces to H[x (1), u (1 ), λ(1 ), 1] = H[x (1), u (1 + ), λ(1 + ), 1]. From the definition of the Hamiltonian H in (6.59), we can rewrite the condition as λ 1 (1 )[ y(1)/2 p(1 )v (1 )] + λ 2 (1 )v (1 ) = λ 1 (1 + )[ y(1)/2 p(1 + )v (1 + )] + λ 2 (1 + )v (1 + ). Since λ 1 (t) = 1 for all t, the above condition reduces to p(1 )v (1 ) + λ 2 (1 )v (1 ) = p(1 + )v (1 + ) + λ 2 (1 + )v (1 + ). Substituting the values of p(1 ) = p(1 + ) = 5 from (6.57), λ 2 (1 + ) = 3 from (6.70), and v (1 + ) = 0 and v (1 ) = 1 from the above discussion, we obtain 5( 1) + λ 2 (1 )( 1) = 5(0) + 3(0) = 0 λ 2 (1 ) = 5. (6.72) We can now use the jump condition in (4.29) on the adjoint variables to obtain λ 2 (1 ) = λ 2 (1 + ) + ζ(1) ζ(1) = λ 2 (1 ) λ 2 (1 + ) = 5 3 = 2. It is important to note that in the interval [1, 1.8], the optimal control condition (6.61) holds, justifying our supposition that v = 0 in this interval. Furthermore, using (6.72) and (6.67), λ 2 (t) = t/2 + 9/2 for t [0, 1), (6.73) and the optimal control condition (6.60) holds, justifying our supposition that v = 1 in this interval. The graphs of λ 2 (t), p(t), and v (t) are displayed in Figure 6.7. To complete the solution of the problem, you are asked to determine the values of µ 1, µ 2, and η in these various intervals.

176 6. Applications to Production and Inventory straint y 1 or 1 y 0, (6.74) changing the terminal time to T = 4, and defining the price trajectory to be 2t + 7 for t [0, 2), p(t) = (6.75) t + 1 for t [2, 4]. The Hamiltonian of the new problem is unchanged and is given in (6.59). Furthermore, λ 1 = 1. The optimal control is defined in three parts as: v (t) = bang[ 1, 1; λ 2 (t) p(t)] when 0 < y < 1, (6.76) v (t) = bang [0, 1; λ 2 (t) p(t)] when y = 0, (6.77) v (t) = bang[ 1, 0; λ 2 (t) p(t)] when y = 1. (6.78) Defining a Lagrange multiplier η 1 for the derivative of (6.74), i.e., for ẏ = v 0, we form the Lagrangian L = H + µ 1 (v + 1) + µ 2 (1 v) + ηv + η 1 ( v), (6.79) where µ 1, µ 2, and η satisfy (6.63)-(6.65) and η 1 satisfies η 1 0, η 1 (1 y) = 0, η 0. (6.80) Furthermore, the optimal trajectory must satisfy As before, λ 1 = 1 and λ 2 satisfies L v = λ 2 p + µ 1 µ 2 + η η 1 = 0. (6.81) λ 2 = 1/2, λ 2 (4) = p(4) = 5. (6.82) Let ˆt be the time of the last jump of the adjoint function λ 2 (t) before the terminal time T = 4. Then, λ 2 (t) = t/2 + 3 for ˆt t 4. (6.83) The graph of (6.83) intersects the price trajectory at t = 8/5 as shown in Figure 6.9. It also stays above the price trajectory in the interval [8/5, 4] so that, if there were no warehousing constraint (6.74), the optimal decision in this interval would be to buy at the maximum

6.3. Decision Horizons and Forecast Horizons 177 Figure 6.9: Optimal Policy and Horizons for the Wheat Trading Model with Warehouse Constraint rate. However, with the constraint (6.74), this is not possible. Thus ˆt > 8/5, since λ 2 will have a jump in the interval [8/5, 4]. To find the actual value of ˆt we must insert a line of slope 1/2 above the minimum price at t = 2 in such a way that its two intersection points with the price trajectory are exactly one time unit (the time required to fill up the warehouse) apart. Thus using (6.75), ˆt must satisfy 2(ˆt 1) + 7 + (1/2)(1) = ˆt + 1, which yields ˆt = 17/6. The rest of the analysis for determining λ 2 including the jump conditions is similar to that given in Section 6.2.4. Thus,

178 6. Applications to Production and Inventory λ 2 (t) = t/2 + 9/2 for t [0, 1), t/2 + 29/12 for t [1, 17/6), t/2 + 3 for t [17/6, 4]. (6.84) Given (6.84), the optimal policy is given by (6.76)-(6.78) and is shown in Figure 6.9. To complete the maximum principle we must derive expressions for the Lagrange multipliers in the four intervals shown in Figure 6.9. Interval [0, 1) : µ 2 = η = η 1 = 0, µ 1 = p λ 2 > 0, v = 1, 0 < y < 1. Interval [1, 11/6) : µ 1 = µ 2 = η 1 = 0, η = p λ 2 > 0, η 0, v = 0, y = 0. Interval [11/16, 17/6) : µ 1 = η = η 1 = 0, µ 2 = λ 2 p > 0, v = 1, 0 < y < 1. Interval [17/6, 4] : µ 1 = µ 2 = η = 0, η 1 = λ 2 p > 0, η 0, v = 0, y = 1. In Exercise 6.16 you are asked to solve another variant of this problem. For the example in Figure 6.9 we have labeled t = 1 as a decision horizon and ˆt = 17/6 as a strong forecast horizon. By that we mean that the optimal decision in [0, 1] continues to be to sell at the maximum rate regardless of the price trajectory p(t) for t > 17/6. Because ˆt = 17/6 is a strong forecast horizon, we can terminate the price shield at that time as shown in the figure. In order to illustrate the statements in the previous paragraph, we consider two examples of price changes after ˆt = 17/6.

182 6. Applications to Production and Inventory Note that if we had solved the problem with T = 1, then y (1) = 0; and if we had solved the problem with T = 17/6, then y (1) = 0 and y (17/6) = 1. The latter problem has the smallest T such that both y = 0 and y = 1 occur for t > 0, given the price trajectory. This is one of the ways that time 17/6 can be found to be a forecast horizon which follows the decision horizon at time 1. There are other ways to find strong forecast horizons; see Pekelman (1974, 1975, 1979), Lundin and Morton (1975), Morton (1978), Kleindorfer and Lieber (1979), Sethi and Chand (1979, 1981), Chand and Sethi (1982, 1983, 1990), Bensoussan, Crouhy, and Proth (1983), Teng, Thompson, and Sethi (1984), Bean and Smith (1984), Bes and Sethi (1988), Bhaskaran and Sethi (1987, 1988), Chand, Sethi, and Proth (1990), Bylka and Sethi (1992), Bylka, Sethi, and Sorger (1992), and Chand, Sethi, and Sorger (1992). EXERCISES FOR CHAPTER 6 6.1 Verify the expressions for a 1 and a 2 given in (6.15) and (6.16). 6.2 For the model of Section 6.1.4, derive the turnpike triple by using the conditions in (6.26). 6.3 Verify (6.34). Note that ρ = 0 is assumed in Section 6.1.5. 6.4 Verify (6.36). Again assume ρ = 0. 6.5 Given the demand function S = t(t 4)(t 8)(t 12)(t 16) + 30, ρ = 0, Î = 15, T = 16, and α = 1, obtain Q(t) from (6.34). 6.6 Give an intuitive interpretation of (6.48). 6.7 Assume that there is a transaction cost cv 2 when v units of wheat are bought or sold in the model of Section 6.2.1. Derive the form of the optimal policy. 6.8 Set up the two-point boundary value problem for Exercise 6.7 with c =.05, h(y) = (1/2)y 2, and the remaining values of parameters as in the model of Section 6.2.3. 6.9 Use a computer to solve the TPBVP of Exercise 6.8 by using the flow chart given in Figure 6.12. (This method is called the shooting method.) You may use EXCEL as illustrated in Section 2.5.

Exercises for Chapter 6 183 Figure 6.12: The Flow Chart for Exercise 6.9 6.10 In Exercise 6.7, assume T = 10, x(0) = 10, y(0) = 0, c = 1/18, h(y) = (1/2)y 2, V 1 = V 2 =, r = 0, and p(t) = 10 + t. Solve the resulting TPBVP to obtain the optimal control in closed form. 6.11 Show that the solution obtained for the problem in Section 6.2.3 satisfies the necessary conditions of the maximum principle. Conclude the optimality of the solution by showing that the maximum principle is sufficient. 6.12 Re-solve the problem of Section 6.2.3 with V 1 = 2 and V 2 = 1. 6.13 Compute the optimal trajectories for µ 1, µ 2, and η for the model in Section 6.2.4. 6.14 Solve the model in Section 6.2.4 with each of the following conditions: (a) y(0) = 2. (b) T = 10 and p(t) = 2t 2 for 3 t 10. 6.15 Re-solve the model of Section 6.3.2 with y(0) = 1/2 and the warehousing constraint y 1/2. 6.16 Verify that the solution shown in Figure 6.10 satisfies the maximum principle.

184 6. Applications to Production and Inventory 6.17 Solve and interpret the following production planning problem with linear inventory holding costs: { T max J = [hi + c } 2 P 2 ]dt subject to 0 I = P, I(0) = 0, I(T) = B; 0 < B < ht 2 /2c, (6.85) P 0 and I 0. 6.18 Re-solve Exercise 6.17 with the state equation I(t) = P(t) S(t), I(0) = I 0 0, I(T) not fixed. Assume the demand S(t) to be continuous in t and non-negative. Keep the state constraint I 0, but drop the production constraint P 0 for simplicity. For specificity, you may assume S = sinπt + C with the constant C 1 and T = 4. (Note that negative production can and will occur when initial inventory I 0 is too large. Of course, how large is too large depends on the other parameters of the problem.) 6.19 Re-solve Exercise 6.17 with the state equation I(t) = P(t) S, where S > 0 and h > 0 are constants, I(0) = I 0 > cs 2 /2h, and I(T) not fixed. Assume that T is sufficiently large. Also, graph the optimal P (t) and I (t), t [0, T].