Analysis of Inventory Systems
Analysis of Inventory Systems
Vqm
PRENTICE-HALL QUANTITATIVE METHODS SERIES
Dr. W. Allen Spivey, Editor
PRENTICE-HALL, INC.
G. Hadley
University of Chicago
T. M. Whitin
University of California, Berkeley
03295-C
PREFACE
During the last fifteen years there has been a rapid growth of interest in
what is often referred to as scientific inventory control. Scientific inventory
control is generally understood to be the use of mathematical models to
obtain rules for operating inventory systems. The subject has attracted
such wide interest that today every serious student in the management
science or industrial engineering areas is expected to have had some experi¬
ence working with inventory models. Originally, the development of in¬
ventory models had practical application as an immediate objective. To
a large extent this is still true, but as the subject becomes older, better
developed, and more thoroughly explored, an increasing number of indi¬
viduals are working with inventory models because they present interesting
theoretical problems in mathematics. For such individuals, practical appli¬
cation is not a major objective, although there is the possibility that their
theoretical work may be helpful in practice at some future time. Thus,
today work is being done with inventory models at many different levels,
ranging from a concern only for practical problems to a concern only for
the abstract mathematical properties of the model.
The purpose of this text is to introduce the reader to the techniques of
constructing and analyzing mathematical models of inventory systems.
In doing so, it cuts across many of the levels of work being done with
inventory models. Thus, by reading the entire text it should be possible
for the reader to gain an understanding of the sort of work that is being
done and the kinds of problems that are encountered when studying inven¬
tory models at all levels—from the purely practical to the purely theoretical.
It was recognized that in many cases readers would be only interested
in one facet of the subject, such as practical applications. An effort has
been made to write the text in such a way that it can be read by a broad
group of readers with widely varying mathematical backgrounds. To ac¬
complish this, the material presented first in each chapter is of practical
§ interest, and involves the most elementary mathematics. The more de¬
tailed and exact developments, which are also more advanced mathe-
PC V
<
£ THE fillin' USMHs'
tmmi INSTITUTE OF TECHH0L06*
VI PREFACE
one for Chapter 5. Also the authors are indebted to a number of their
students who pointed up errors and misprints in the manuscript. Finally,
the authors express their appreciation to the editor of Operations Research
for permission to reprint a number of the Poisson properties which they
first published there.
G.H.
T.M.W.
CONTENTS
2- 1. INTRODUCTION 29
2-2. THE SIMPLEST LOT SIZE MODEL—NO STOCKOUTS 29
2-3. ADDITIONAL PROPERTIES OF THE MODEL; AN EXAMPLE 34
2-4. ACCOUNTING FOR INTEGRALITY OF DEMAND 40
2-5. CASE WHERE BACKORDERS ARE PERMITTED 42
2-6. THE LOST SALES CASE 47
vin
IX CONTENTS
7- 1. INTRODUCTION 323
7-2. DYNAMIC PROGRAMMING 323
7-3. DYNAMIC PROGRAMMING FORMULATION OF OTHER
PROBLEMS 330
7-4. DYNAMIC-DETERMINISTIC LOT SIZE MODEL 336
xi CONTENTS
8- 1. INTRODUCTION 361
8-2. THE BASIC FUNCTIONAL EQUATION FOR PERIODIC
REVIEW SYSTEMS 361
8-3. OPTIMALITY OF Rr POLICIES FOR PERIODIC REVIEW
SYSTEMS—QUALITATIVE DISCUSSION 364
8-4. PROOF THAT AN Rr POLICY IS OPTIMAL WHEN f(y; T) IS
CONVEX AND THE UNIT COST IS CONSTANT 367
8-5. USE OF DYNAMIC PROGRAMMING TO COMPUTE OPTIMAL
VALUES OF R AND r 374
8-6. THE FUNCTIONAL EQUATION FOR TRANSACTIONS
REPORTING SYSTEMS 376
8-7. OPTIMALITY OF Rr POLICIES FOR TRANSACTIONS
REPORTING SYSTEMS 381
8-8. EXPLICIT SOLUTION OF THE FUNCTIONAL EQUATION
WHEN A POISSON PROCESS GENERATES DEMANDS
AND UNITS ARE DEMANDED ONE AT A TIME 382
8- 9. EXPLICIT SOLUTION OF THE FUNCTIONAL EQUATION IN
THE GENERAL CASE 388
8-10. EXPLICIT SOLUTION OF THE FUNCTIONAL EQUATION FOR
THE STEADY STATE PERIODIC REVIEW MODEL 391
8- 11. LOST SALES CASE 394
9- 1. INTRODUCTION 401
9-2. RELEVANCE OF THE MODEL 401
9-3. DATA PROBLEMS 405
9-4. THE DEMAND DISTRIBUTION 406
9-5. DEMAND PREDICTION 413
9-6. THE LEAD TIME DISTRIBUTION 419
9-7. DETERMINATION OF COSTS 420
9-8. MULTI-ITEM PROBLEMS 423
9-9. PERSONNEL AND PROCEDURAL PROBLEMS 425
9- 10. THE EVALUATION PROBLEM 427
9-11. SUMMARY 429
THE NATURE
OF INVENTORY SYSTEMS
Albert Einstein
1-1 Inventory Problems
other words, the resupply of the inventory of water within the dam depends
on the rainfall, and the organization operating the dam has no control over
this.) We shall not consider this type of problem here. The only problems
with which we shall concern ourselves are those in which the organization
controlling the inventory has some freedom in determining when, and in
what quantity, the inventory should be replaced. On the other hand, we
shall assume that, in general, the inventory system has no control over the
demands which occur for the item, or items, which it stocks. Again, this is
just the opposite of what one encounters in dealing with inventory problems
such as storage of water within dams, since the efflux of water through
the dam is completely within the control of the organization operating the
dam. In short, we are going to consider the type of inventory problem
encountered in business, industry, and the military.
We shall concentrate on showing how mathematical analysis can be used
to help develop operating rules for controlling inventory systems. When
mathematics is applied to the solution of inventory problems, it is necessary
to describe mathematically the system to be studied. Such a description is
often referred to as a mathematical model. The procedure is to construct
a mathematical model of the system of interest and then to study the
properties of the model. Because it is never possible to represent the real
world with complete accuracy, certain approximations and simplifications
must be made when constructing a mathematical model. There are many
reasons for this. One is that it is essentially impossible to find out what the
real world is really like. Another is that a very accurate model of the real
world can become impossibly difficult to work with mathematically. A
final reason is that accurate models often cannot be justified on economic
grounds. Simple approximate ones will yield results which are good enough
so that the additional improvement obtained from a better model is not
sufficient to justify its additional cost.
In this book we shall study a variety of mathematical models of in¬
ventory systems. Many of these are intended for practical application.
Others, however, have no immediate practical application because of the
restrictive nature of the assumptions. They are interesting and relevant,
however, because they exhibit some theoretical properties which are
important in understanding the nature of inventory systems.
Although inventory problems are as old as history itself, it has only been
since the turn of the century that any attempts have been mdc to employ
analytical techniques in studying these problems. The initial impetus for
the use of mathematical methods in inventory analysis seems to have been
SEC 1-2 BRIEF HISTORICAL SKETCH 3
the simplest forms of interaction involves one stocking point which serves
as a warehouse for one or more other stocking points. This leads to what
is referred to as a multiechelon inventory system. One possible type of
multiechelon system is illustrated in Fig. 1-1. The arrows indicate the
normal pattern for the flow of goods through the system. This might be
referred to as a four echelon system since there are four levels. Each level
is called an echelon. In the system shown, customer demands occur only
at the stocking points in level 1. These stocking points have their stocks
TTTTT 1
FIGURE 1-1.
the multiechelon system in its entirety. The reason for this is that different
organizations operate different parts of the system. For example, Fig. 1-1
might refer to a production distribution system in which the source is a
plant where the item is manufactured, level 4 is a factory warehouse, level
3 represents regional warehouses, level 2 represents warehouses in various
cities, and level 1 represents the retail establishments which sell the item to
the public. In such a system the manufacturer might control only the
plant and the factory warehouse, while different organizations operate the
regional warehouses and still different organizations operate the city ware¬
houses and the retail establishments. Note that even at a given level many
different organizations may be involved. For example, each of the ware¬
houses in different cities may be under different ownership. In such a
system, each organization has the freedom to choose the operating doctrine
for controlling the inventories under its jurisdiction. One could not, in
general, attempt to analyze the system as a whole and dictate what oper¬
ating doctrine should be used by each stocking point at each level. Instead,
one might be concerned with the best way for one of the warehouses at
level 2 to control its inventories. In making the analysis, the customers
would be the retailers at level 1 and the source from which replenishments
are obtained would be the appropriate warehouse at level 3.
Frequently, there will be just a single source from which an inventory
system replenishes its stocks when it is desirable to do so. This source may
be the plant where the item is made, a factory warehouse, or simply a ware¬
house at a higher echelon. Sometimes, however, the system has two or
more alternative sources of supply available. For example, one of the
retailers in Fig. 1-1 might be able to order from several different warehouses
at level 2. A special case which can occur is that where the system under
study also controls the source of its supplies, i.e., the plant where the item
or items are made. In this case the problem is not strictly an inventory
control problem but also involves production scheduling. In this book we
shall not consider the general problem of combined production scheduling
and inventory control except in some cases where production is carried out
in lots rather than being continuous.
The basic inventory system that will be studied in this text will be much
simpler than the general sort of multiechelon system shown in Fig. 1-1. It
will consist of just one stocking point with a single source for resupply.
Customer demands arrive at the single stocking point, and at appropriate
times orders are placed with the source for replenishing the inventory. The
operation of this system is illustrated schematically in Fig. 1-2.
There are good reasons for restricting our attention to the structure illus¬
trated in Fig. 1-2. Perhaps the most important reason is that it is very
difficult to study analytically multiechelon systems of the type shown in
Fig. 1-1. In fact, very little work has been done in this area. It will be seen
SEC. 1-5 THE NATURE OF THE ITEMS 7
Customer Orders
demands Single Single
stocking
source
Goods to point Goods
customers
FIGURE 1-2.
that even the relatively simple structure that will be studied can become
very complex to analyze. The other reason is that for practical applications
the simple structure of Fig. 1-2 is often (although by no means always)
adequate. This is true because, as was noted above, even though real
world systems are usually of the multiechelon variety, it is often necessary
to consider the various stocking points individually because different
organizations control them. Even when a single organization controls a
number of stocking points, however, the interactions between them are
frequently sufficiently small to allow each to be studied independently of
the others.
A large military supply system stocks over 500,000 different items, while
a typical department store may carry as many as 150,000 items. Other
inventory systems stock only one or two items. The items stocked can
differ from each other in many ways, They differ in cost, and in their
physical properties, such as weight and volume. Some items are perishable
and cannot be stored for long periods of time; others can be stored in¬
definitely without deterioration; others are subject to rapid obsolescence.
Often, items can be stored only under specially controlled conditions of
temperature, humidity, etc., and require special types of packaging for
storage. When more than a single item is stocked there can be interactions
between the items. For example, the items may be substitutes for each
other so that if the system is out of stock on one item a customer will accept
another. On the other hand, they may be complements so that usually one
will not be sold without the other. Frequently, the interactions will take
the form of having items compete for limited warehouse floor space or for
investment dollars, which are also limited.
Another different form of interaction can exist between the items carried
by an inventory system. This type of interaction is represented in what
might be called a multistage inventory system associated with a production
process. A typical block diagram for such a system is illustrated in Fig. 1-3.
8 THE NATURE OF INVENTORY SYSTEMS CHAP. 1
FIGURE 1-3.
obsolete ultimately. Only one of the models will deal specifically with
perishable items that can be stored only for short periods of time. We shall
never need to concern ourselves directly with any of the special, controlled
conditions required for storage. Only a limited amount can be done in
treating interactions between items. Simple cases, where items are com-
peting for warehouse floor space or investment dollars, can be examined.
In a case where there is competition for floor space we shall also make use
of the unit volume of the items.
Generally speaking, it is almost never true that enough is known about the
process which generates demands for items carried by an inventory system
SIC l-d STOCHASTIC PROCESSES ASSOCIATED WITH THE INVENTORY SYSTEM 9
restrictive assumptions. The backorders case and lost sales case represent
fundamentally different stochastic processes.
For any particular inventory system it can happen that some demands
occurring when the system is out of stock are backordered while others are
lost. If it is possible to develop models of the system for the backorders and
the lost sales cases respectively, it is not difficult to develop a single model
in which some sales are lost and some are backordered.
We shall define the procurement lead time (or simply the lead time) for
an inventory system as the interval between the time when the stocking
point decides that an order for a replenishment should be made and the
time that the order arrives and is on the shelves, available to customers.
Often the procurement lead time will not be constant, since the time to fill
the order at the source, the shipping time, and the time required to carry
out the paper work, etc., can vary from one order to another. It is seldom
possible to predict in advance precisely what the lead time will be. Some¬
times the variations in the lead time will be small enough so that in the
mathematical model the lead tune can be assumed to be absolutely con¬
stant. In other situations, however, it will exhibit sufficient unpredictable
variability that it is necessary to assume that a stochastic process generates
the lead times. We shall consider both models in which the lead times are
constant and models in which lead times are stochastic variables.
Consider next the costs which are independent of the quantity ordered.
These costs include paper, postage, telephone charges etc., as well as the
labor costs incurred in processing the order. They also include those parts
of receiving and inspection costs which are independent of the order size.
If the inventory system controls the plant where the item under consider¬
ation is made, then assuming that the item is made in lots, the set up costs
for a production run will fall into this category. These costs which do not
depend on the quantity ordered are incurred each time an order is placed.
They will be referred to as the “fixed” costs of placing an order. We shall
usually denote the fixed cost of placing an order by A. The total cost of
placing an order for Q units will then be A + C(Q).
FIGURE 1-4.
It is important to note that the fixed ordering costs have the property
that, usually, to a good approximation, the total fixed costs incurred on
placing N orders is simply AN. In other words, the total fixed costs in¬
curred are proportional to the number of orders placed. This proportion¬
ality of total fixed costs to the number of orders placed need not be exactly
true because of phenomena like the following: The number of people in the
office staff, and hence the annual cost of the office staff as a function of the
average number of orders placed per unit time might look something like
that shown in Fig. 1-4. As more and more orders are placed, the size of
the office staff must be increased. Each step in Fig. 1-4 might indicate an
increase in the office staff by one person. However, for a given size of office
staff, there is a range of ordering rates which requires this size of staff and
hence a range over which the annual cost of the office staff does not change.
SEC 1-9 INVENTORY CARRYING COSTS 13
This is indicated by the flat portion of Fig. 1-4. Often, though, in the range
of interest, the curve shown in Fig. 1-4 can be approximated well enough
by a straight line through the origin so that a linear dependence of annual
cost on the average ordering rate is obtained.
We shall assume in the mathematical models developed in this book that
the fixed cost of placing N orders is AN, where A is the fixed cost of placing
a single order, so that the total fixed costs incurred are directly proportional
to the number of orders placed. As seen above, this will not in general be
Precisely true in any real world situation. Usually, it is a quite satisfactory
approximation. However, more importantly, models which make this
assumption can be used to handle situations in which there exist step
functions in the costs such as those shown in Fig. 1-4 which must be
explicitly accounted for. The technique for doing this will be presented in
Chapter 9.
re mn L5^y/ ■ -
IC £ x(t) dt
Note that the integral is the area under the curve between t = 0 and t = 1
year. Since the interval of integration is unity it is also true that the
integral is simply the average inventory (averaged over the period of one
year). Thus it follows that our definition of the way carrying costs are
SEC 1-9 INVENTORY CARRYING COSTS 15
concern the fact that the maximum rate of return will also depend on the
sum that is available for investment and the general state of business. Thus
the opportunity cost represented in I can represent only an average rate of
return.
As has been noted previously, the rate of incurring breakage and pilferage
costs is roughly proportional to the investment in inventory, and hence
unless one has information to suggest otherwise, the method of costing we
are using seems reasonable for representing these costs. On the other hand,
our previous discussion has shown that the instantaneous rate of incurring
insurance costs will not in general fluctuate with movements in the inven¬
tory level. However, we have shown above that making the instantaneous
16 THE NATURE OF INVENTORY SYSTEMS CHAP. 1
earned from the purchase date to the obsolescence date if the funds devoted
to the procurement of the item had been invested elsewhere) and its salvage
value. Obsolescence costs are always incurred at a fixed point in time, and
this obsolescence date often cannot be predicted with certainty in advance.
It is much more unrealistic to attempt to include obsolescence costs in the
carrying charge than it is to try to make a provision for taxes in the carry-
ing charge. The usual argument given is that by making a charge against
each item in proportion to the length of time it has remained in inventory,
one is setting up a fund to insure against obsolescence. However, such a
procedure can seriously distort the situation and lead one to using an
operating doctrine which may be far from being a good one. When obso¬
lescence is an important consideration, one should use a mathematical
model that explicitly takes into account the fact that obsolescence costs
are incurred only at a single point in time. We shall consider such models
in Chapter 7.
The inventory carrying charge I can then be thought of as the sum of
several terms and can be written
I = Ii + J2 + Iz + . . . (1-1)
warehouse, someone must go to the proper bin and obtain the unit or units.
Next, it may be necessary to package the order for shipment. Finally, the
order is shipped to the customer. After the order has been shipped, a
record of this transaction is usually sent from the warehouse to accounting
where appropriate additional records are made. If a customer’s order ar¬
rives when the system is out of stock, it will usually be necessary to go
through special procedures to inform the customer of the existing situation.
The costs of the accounting operations referred to above, the salaries of
those in the warehouse who are concerned with filling orders, the costs of
packing, and the shipping costs, if paid by the inventory system, are all part
of the normal costs of filling customers’ orders. The important thing to note
about all these costs is that, while they will vary with the demand rate, they
will in general not depend on the operating doctrine used to control the
inventory system. Hence they need not be considered when studying costs
that vary as the operating doctrine is changed. On the other hand, the
costs arising from the special action required if a customer’s demand arrives
when the system is out of stock will depend on the operating doctrine, since
the fraction of the time that the system is out of stock will depend on the
operating doctrine. We shall find it convenient to include these latter costs
as a part of the stockout costs. With this convention, we shall not need
to consider in the future the costs of filling customers’ orders.
Let us now turn our attention to the costs incurred by having demands
occur when the system is out of stock. In doing so, we must distinguish
between the backorders and lost sales cases. Consider first the case where
all demands occurring when the system is out of stock are backordered.
In any practical situation, it is very difficult to determine accurately the
nature of the backorder costs. Backorder costs are inherently extremely
difficult to measure since they can include such factors as loss of customers’
goodwill (i.e., in the future, he may take his business elsewhere), or in
military supply systems, the cost of having part of some first line weapon
system inoperative because of lack of parts. Other parts of the backorder
cost can be somewhat easier to measure; however, these are usually a small
part of the total backorder cost. Such costs include the cost of notifying a
customer that an item is not in stock and will be backordered plus the cost
of attempting to find out when the customer’s order can be filled and giving
him this information. If the system itself uses the part, the backorder cost
may simply be the cost of keeping a machine idle for lack of parts. In such
a case one can obtain a relatively good measure of the backorder cost.
When units are demanded one at a time, then there will in general be a
SEC 1-11 STOCKOUT COSTS 19
backorder cost associated with each unit backordered. When more than a
single unit can be demanded when a demand occurs, the backorder cost
could depend in a complicated way on the size of the demand. However,
we shall in this text assume that there is a backorder cost associated with
each unit backordered, even when units need not be demanded one at a
time. This cost will in general depend on the length of time for which the
unit remains backordered. It will be assumed here that this cost depends
only on the length of time for which the backorder exists. In principle,
the cost of each unit backordered as a function of the time t for which
the backorder remained on the books might be described by a nondecreas¬
ing function x(f), and this function might change from system to system.
In the real world it is essentially impossible to determine the precise shape
of 7r(f). About the most general function that can be used in practice is
ir(t) = x + fit. In other words, there is a fixed cost for each unit.back-
ordered plus a variable cost which is linear in the length of time for which
the backorder remains on the books, tin the mathematical models developed
m this text, when it is necessary to specify a backorder cost, the most
general function used will be x + ft. When r(t) involves t in a more compli-
cated way than x + M, it becomes considerably more difficult to work with
the mathematical models. Since these more complicated possibilities are
not useful in practice we shall not treat them in detail. However, some of
the problems ask the reader to formulate models involving more compli¬
cated functions for x(f).
If n units are backordered in a one year period, then the cost incurred in
that year attributable to the fixed cost x per unit backordered will be im.
If the number of backorders b(t) on the books as a function of time in the
year under consideration looks like that shown in Fig. 1-6, then the cost
incurred in that year attributable to the variable cost x£ per unit, back¬
ordered will be
x[fl + 4 + 4 •+•••+ 4]
since the height of each rectangle is unity, Now note that if t is measured
in years, then
will have the dimensions of units times years, i.e., unit years. Then if we
want the cost for the year to come out in dollars, x must have the di¬
mensions of dollars per unit year of shortage. Note that since the interval
of integration is unity,
I b(t) dt
Jo
is simply the average number of backorders which existed for the year, i.e.,
20 THE NATURE OF INVENTORY SYSTEMS CHAP. 1
FIGURE 1-6.
the number of unit years of shortage incurred for the year under consider¬
ation is numerically equal to the average number of backorders which
existed for the year.
Let us next consider the lost sales case. Since demands are lost if they
occur when the system is out of stock, the cost of a lost sale cannot depend
on time, for a lost sale is lost and there is nothing which corresponds to
the length of time for which a unit remains backordered. Thus when units
are demanded one at a time the cost of each lost sale will simply be a con¬
stant. If units need not be demanded one at a time, then the cost of a lost
sale could depend in a complicated way on the size of the demand. In the
text we shall assume that when there are costs incurred from lost sales,
then there is a fixed cost associated with each unit demanded which cannot
be met from inventory, even if units need not be demanded one at a time.
The cost of a lost sale includes a number of different factors. As in the
backorders case, most of these are not direct out of pocket costs that would
ever appear on a balance sheet. Perhaps the most important component
of the cost of a lost sale is the somewhat intangible goodwill loss. This can
include lost profits on sales of other items or on future sales of the given
item due to the fact that the customer temporarily or permanently takes
his business elsewhere or because he discourages other potential customers
by telling them that he received unsatisfactory service. The cost of a lost
sale also includes the costs of any special procedures used to inform the
customer that his demand cannot be supplied, and the profit lost in not
making the sale. We shall denote by 7r0 the cost, exclusive of the lost
SEC M3 SELECTION OF AN OPERATING DOCTRINE 21
profit on the unit, of a unit having been demanded when the system is out
o stock. Then x will be used to denote the cost, including the lost profit
on the unit, of having a unit demanded when the system is out of stock.
Thus x is the sum of x0 and the unit profit on the item.
Sometimes we shall refer to the cost of having a unit backordered as the
cost of a backorder. Similarly, the cost of having a sale lost because a unit
is demanded when the system is out of stock will be referred to as the cost
of a lost sale. When we don’t care to differentiate between the costs of
backorders or of lost sales, we shall simply refer to stockout costs.
a
(1-4)
1 — a
Thus 3 and 3C differ only by the constant factor 1/i, and hence minimizing
either one also minimizes the other.
Usually, however, the actual profit or cost will fluctuate from one year
to the next, sometimes being above or below the average annual values as
defined above. Because of this it is possible, by making the interest rate
sufficiently high, to obtain different results by minimizing the average
24 THE NATURE OF INVENTORY SYSTEMS CHAP. 1
annual cost and minimizing the present worth of all future costs, or maxi¬
mizing the average annual profit and maximizing the present worth of all
future profits. £As has been noted above, however, the differences will be
negligible in almost all cases of practical value when the interest rate re¬
mains within the limits appropriate to the real world. It might be noted,
however, that the effect of the interest rate becomes more important as the
average time between the placement of orders is increased. In any case
where the interest rate was large enough to make a difference, then the
present worth would be the function to be optimized}
There remains the question of how any stochastic elements in the
problem are to be handled. For models in which the mean rate of demand
is imagined to remain constant over all future times, we shall again maxi¬
mize the average annual profit or minimize the average annual cost. When
stochastic elements are present, however, it will be necessary to introduce
probabilities and compute expected values in order to determine the aver¬
age annual profit or cost. These expected values will arise in a natural way
as time averages.
For situations where the mean rate of demand changes in some fashion
with time, we shall use the criterion of maximizing expected profits or
minimizing expected costs over some relevant time interval. It is not com¬
pletely evident in such cases that expected values are what should be used.
It would be quite reasonable to use expected values if it was true that the
system would repeatedly encounter the same sorts of conditions and that
the expected profit or cost could be interpreted as the average profit or
cost averaged over all times that the system faced a particular set of
circumstances. However, it will frequently be true that the system will
encounter a particular set of conditions only once, and they wifi never be
repeated agam. The rationale for using expected values in such cases lies
m the modem theory of utility introduced by von Neumann and Morgen-
stem [7], We shall not attempt to examine this theory in detail but
instead shall merely point out that it roughly implies the following: If it
is possible for an individual to express consistent preferences between
various situations whose outcomes must be described probabilistically
then there exists a function having a numerical value associated with
evepr possible outcome such that if he selects among the alternatives in
such a way that the expected value of this function is maximized, then he
wifi be acting m a way that is truly representative of his preferences The
numerical function referred to above is often caUed a utility function. We
shall assume m this text that the system under consideration does have a
utility function defined over the possible outcomes in any inventory
problem. Thus the system should always behave in a way that maximizes
its expected utility. We shaU further assume that the utility can be
measured in monetary units and is, in fact, with the proper definition of
REFERENCES 25
the relevant components, simply the expected profit over some appropriate
time period. It is in this way that we justify the use of the expected profit
even in situations where the given set of conditions can never be expected
to be encountered again. We shall also see that in a number of cases the
maximization of expected profit is eequivalent to the minimization of ex¬
pected cost. In these cases then, one can equally well minimize the expected
cost over some relevant time interval.
We have noted previously that it is often very difficult in practice to
determine the stockout cost functions. To avoid this problem, an alterna¬
tive procedure might be to maximize the profit or minimize the cost, each
exclusive of the stockout cost, subject to a constraint that the average
fraction of the time for which the system is out of stock is not greater than
a specified value. Here, instead of specifying the nature of the stockout
cost, one instead specifies an upper limit to the average fraction of the t.imp
for which the system is out of stock. We shall also examine criteria of this
sort for the determination of operating doctrines and shall also investigate
their relation to the cases where a stockout cost function is specified.
Before closing this introductory chapter, we might say something about
the physical dimensions that will be employed in formulating the profit or
cost expressions in the remainder of the text. The equations will hold for
any consistent set of physical dimensions. However, we shall always
imagine that the monetary unit is the dollar and the time unit is a year.
Of course, physical quantities of any good will be measured simply in units,
where the unit need not be defined more precisely. A unit may, for example,
be a case of screwdrivers, a dozen screwdrivers, or a single screwdriver'
depending on the application. The only important thing to remember is
that no matter how the unit is defined, the definition must be used con¬
sistently throughout.
As a final remark it is worth observing that while we shall be concerned
with optimizing the mathematical models to be studied, it does not follow
that the real world system represented by the mathematical model will
also be optimized. Since a number of simplifying assumptions and approxi¬
mations must be made to obtain the mathematical model, the most we can
expect is that by using the optimal solution for the model in the real world
we shall have a “good” operating doctrine or one that is an improvement
over an existing operating doctrine. Indeed, in the real world situations,
there are usually such a large number of complications that it is exception¬
ally difficult to define what an optimal operating doctrine mpa.ns
REFERENCES
1. Arrow, K. J., T. Harris, and J. Marschak, “Optimal Inventory Policy,”
Econometrica, XIX (1951), pp. 250-272. ’
26 THE NATURE OF INVENTORY SYSTEMS CHAP. 1
and the on hand inventory at the point in time when an order arrives will
always have the same valued"!
In this section we wish to study a case where the system is never out
of stock when a demand occurs. This limitation can legitimately be im¬
posed since the demand is deterministic and the procurement lead time is
a constant. Initially it will be convenient to imagine that the quantity
demanded in a time £ is a continuous function of t, and to ignore the fact
that an integral number of units must be demanded. Later, account will
be taken of the integrality of demand. jjSince the system is never out of
stock when a demand occurs, and since X is a constant, the annual revenues
received from the sale of the item are a constant, independent of the oper¬
ating doctrine, and hence, the minimization ofcosts will yield the same
operating doctrine as the maximization of profit&J Following the suggestion
of Chapter 1, we shall determine the optimal Operating doctrine by mini¬
mizing the average annual cost. Recall that only those costs that depend
on the operating doctrine need be included. These costs have been dis¬
cussed in Chapter l.^The costs appropriate to this model are the cost of
the units purchased, the fixed cost of placing an order (the ordering cost),
and the inventory carrying costs^ It will be assumed that the costs of
operating the information processing activity are independent of the order
size and the reorder rule, and therefore do not need to be included in the
cost expression.
[To compute the average annu'al cost, we must compute the total cost
for an arbitrary time period of length f, then divide by f to obtain the aver¬
age annual cost for the period of length and finally allow f to approach
infinity, giving the desired average annual cost| The actual cost of operat¬
ing the system can vary from one year to another because the number of
orders actually placed can vary. Only when a year is an integral multiple of
the time between placing orders can the actual system cost be the same
each year. The differences between average yearly cost and the actual
system cost for a given year will be illustrated later by a concrete example.
The procedure we are about to use to determine the average annual cost
may, at first reading, seem unnecessarily complicated for the simple model
being studied. It is introduced now because it provides a rigorous way of
obtaining the average annual cost and because it allows us to introduce, in
the context of a very simple model, a technique which will later prove quite
useful in dealing with more complex models involving uncertainty. Further,
the technique used to derive the average annual cost is of course the
definition of the average annual cost, i.e., it is by definition the limit as f
approaches infinity of the average cost for a time period of length f.
* A rigorous proof that the optimal operating doctrine has this form can be given
using the methods to be introduced later in Chapter 8.
SEC 2-2 SIMPLEST LOT SIZE MODEL—NO STOCKOUTS 31
This follows since at any time t from the beginning of the cycle (which is
taken for convenience to be the time of arrival of an order), the on hand
inventory is Q + $ — A£, and the cost incurred between t and t + dt is
IC(Q + s — At) dt Summation over the length of the cycle gives (2-1)
when the relation T = Q/X is used.
In the time f there are
icr[| -$] ^l + sJ + ,7
where y is the inventory carrying cost for the fractional cycle of length
$ — vT; note that y is less than the inventory carrying cost for a single
cycle. The total variable cost for the time period of length f is then
+ «] QC + r + «] A + ICT [| - *] [| + ,] + „ (2-2)
32 DETERMINISTIC LOT SIZE MODELS CHAP. 2
The average annual cost for the time period of length f, X?, is obtained by
dividing the total cost by This yields
3Cr = XC + ^ ^ A + j + IC [| + s] - + s] + *
3C = XC + ^ A + IC [| + s]; (2-3)
The formula for X can also be derived using the following simple argument:
Since there are X demands per year and sinceall demands are met^ then
on thejjiverage X units per year must be procured at a cost of XC. Similarly,
if the order quantity is Q, then the number of orders placed per year must
average to X/Q, and the fixed procurement costs per year average to \A/Q.
Furthermore, by assumption, the inventory carrying cost per unit is* pro¬
portional to the length of time it remains in inventory. Consequently, the
inventory carrying cost per year must then be IC times the average inven¬
tory. The average inventory is one half the sum of the maximum inventory
Q + s and the minimum inventory s, i.e., (Q/2) + s. Summation of these
three terms gives (2-3).
For the moment (indeed, in all discussions through Sec. 2-10), we shall
restrict our attention to situations for which the unit cost of the item is
independent of the quantity ordered. Then XC is independent of Q and the
reordering rule and need not be included in the variable cost. Hence the
relevant average annual variable cost in this case, which is the sum of
ordering and inventory carrying costs, is
X = qA + IC ^ J (2-4)
where m = Xt is the lead time demand (i.e., the number of units demanded
from the time an order is placed until it arrives), the on hand inventory will
be zero at the time the order arrives. The number rh is called the reorder
point; each time the on handfinventory in the system reaches rh an order
for Q units is placed. This is illustrated graphically in Fig. 2-1.
It remains to determine the optimal value of Q. We have seen that the
optimal value of s is zero for any Q. Thus the average annual cost really
depends only on the single variable Q; it can be written--
K = ±A+IC\ (2-6)
FIGURE 2-1.
^ = o =-Aa , rc
dQ U Q*A+ 2
or
34 DETERMINISTIC LOT SIZE MODELS CHAP. 2
d2K _ 2\A
(2-9)
dQ2 Qz
for all Q > 0 and hence the Q determined from (2-7) yields the absolute
minimum value of K.
The equation (2-8) is referred to in the literature under a variety of
names. It is sometimes called the lot size formula, the economic order
quantity, the square root formula, or the Wilson formula. We shall oc¬
casionally use several of the above names. Equation (2-8) might also be
called the Harris formula since, as indicated in Chapter 1, Harris seems to
have been the first to have derived it. Because the square root expression
in (2-8) appears frequently in later work, we use a special symbol Qw to
identify it; whenever Qw appears, it will always be defined by (2-8).
The problem of determining how to operate the system has now been
solved. The reorder point, given by (2-5) (with Q* replacing Q), tells us
when an order should be placed. The quantity to be ordered is given by
(2-8). Specification of the reorder point and the order quantity determines
all other quantities of interest.
Q*
2 (2-10)
FIGURE 2-2.
of the sales rate and not proportionally with the sales rate. Similarly, the
average inventory varies inversely as the square root of the cost so that the
average inventory for high cost items should be lower than for low cost
items, all other things being equals The above observations merely serve
to point out that in formulating an inventory policy, items with different
characteristics should be treated differently. This implies, for example,
that if a system was stocking a number of items with widely different rates
of demand and costs, it would not be optimal to require that the average
inventory for each item should be k weeks of stock where the same k was
used for all items.
36 DETERMINISTIC LOT SIZE MODELS CHAP. 2
borhood of the optimal Q. If the actual Q is off from the optimal Q m either
direction by a factor of two, costs are increased by only 25%. The same
sort of analysis can be used to study how the actual cost will vary from the
optimum cost tf one of the parameters such as X or A is not measured
proWems Exammatlons of sensitivity analyses of this sort are left to the
In the event that the procurement lead time is less than one cycle there
will never be more than a single order outstanding. Furthermore, there
will be no orders outstanding at the time immediately prior to placing an
order (i.e., just as the reorder point is reached). On the other hand, if the
procurement lead time is longer than one cycle, there will always be at least
one order outstandmg. As an aid in visualizing the situation where the
placed received
FIGURE 2-4.
m is the largest integer less than r/T, and (to + 1) Q immediately after
hitting the reorder point (or, in the razor’s edge case where r/T is an
integer to, there are always precisely to orders outstanding).
It is often useful to consider the quantity on hand plus on order. At the
reorder point the on hand inventory is p - mQ, and the quantity on hand
plus on order increases by Q units when the order is placed. Thus the
quantity on hand plus on order fluctuates between p and p + Q. The re¬
lationship between the on hand inventory and the on hand plus on order
inventory is shown in Fig. 2-5. The on hand inventory reaches its minimum
value 0 just prior to receipt of a procurement, and its maximum value
in terms of the on hand plus on order inventory. This reorder point will
be denoted by r, and r = (ji being the lead time demand); thus when the
on hand plus on order inventory reaches a level M an order is placed Later
when uncertainty in the demand is introduced, we shall see that the reorder
point cannot always be legitimately determined in terms of an on hand
inventory. Instead, it must be specified in terms of the on hand plus on
order inventory (or on a more general level defined to be the quantity on
hand plus on order minus backorders). It is only for certain special cases
that the reorder point can be specified in terms of an on hand level.
A = $8.00, ip r = 1 year
C = $0.30 £
To be more specific, it can be imagined that this is a low cost item carried
by a department store. The high fixed cost of ordering arises because the
item is procured in Europe. This also accounts for the long lead time.
The optimal order quantity is
r* - 91 _ 400 2
X 600 3F'
H = Xr = 600 units.
f^ie reor<^er P°in^ based on the on hand plus on order inventory level is then
r* = 600. To compute the reorder point based on the on hand inventory
level, we first note that t/T = f. The largest integer less than r/T is 1
Thus
rt = /x — Q* = 200 units.
Thus the ordering and holding costs are a small part of the total cost
which includes the cost of the units.
The actual yearly procurement and holding costs vary from one year to
another. This can be seen from Fig. 2-6. Assume that we start measuring
tune immediately after the placing of an order. Then in the first year
shown, one order is placed, while in the second year two orders are placed.
After the second year a repetition of the above is obtained. For the first
year shown in Fig. 2-6, the actual system costs are $8 for ordering and $10
for holding, or a total of $18. In the second year, the actual costs are $16
for ordering and $14 for holding or a total of $30. The average of these
two values gives $24, which is that obtained above for K*.
a time and the time between demands ts is known with certainty. Then
X = 1 /t„.
As in the continuous case, the number of units s on hand when an order
arrives should be zero. Now, however, this requirement does not uniquely
specify the time when an order should be placed since there is a time inter¬
val of length t, between demands. It should be noted, however, that there
is no need to have the order arrive until precisely the moment when a
demand occurs for otherwise costs will be increased. Thus the order should
arrive at a time t„ after the demand which reduces the on hand inventory
to zero, i.e., the system will have a zero on hand stock level for a time t,
each cycle. Immediately after the procurement arrives one unit is de¬
manded. Hence, the maximum inventory level is Q-1 and the minimum is
zero.
The inventory carrying costs per cycle are then*
The average number of cycles per year remains X/Q so that the average
mventory carrying costs per year are } 1C (Q - 1). The average yearly
procurement costs remain unchanged at XA/Q. Thus the average annual
cost of holding and procuring inventory is for a given Q > 1
1 +2 +. . n(n + 1)
+n -
2
42 DETERMINISTIC LOT SIZE MODELS CHAP. 2
The model presented in Sec. 2-2 required that all demands could be met
from stock, i.e., the system was never out of stock when a demand occurred.
We shall now study the more general case in which all demands must be
met ultimately, but it is permissible for the system to be out of stock when
a demand occurs. In such a case the demands occurring when the system
is out of stock are backordered until a procurement arrives. When a pro¬
curement does arrive it is assumed that all backorders are met before the
procurement can be used to meet any other demands.!
Clearly, if there were no costs associated with incurring backorders, then
it would be optimal never to have any inventory on hand. On the other
hand, if backorders are sufficiently expensive, then one should never incur
any. However, for an intermediate range of backorder costs, it will be opti¬
mal to incur some backorders towards the end of a cycle. In accordance
with the discussion of Chapter 1, we shall assume that the cost of a back- ]
order has the form w + M where t is the length of time for which the back-!
order exists. The cost of a backorder thus includes a fixed cost t and a
cost M which is proportional to the length of time for which the backorder!
exists. _/'
Let s be the number of backorders on the books when a procurement of
SEC 2-5 CASE WHERE BACKORDERS ARE PERMITTED 43
IC JoTl (Q - s - M) dt = g (Q - ,)*
This result is also immediately evident from Fig- 2-7, since (Q — s)2/2X
FIGURE 2-7.
is the area of triangle (1). As before, there are on the average X/Q
cycles per year, so that the average yearly cost of carrying inventory is
IC{Q - s)2/2Q.
The backorder cost per cycle is
ITS Hr
f
X f
Jo
Xt dt = 7TS "j” X
z
TrXTl = TTS + ~-
zx
* Note that the minimum value of Q — s is zero. Costs could never be reduced
.by having backorders on the books immediately after the arrival of a procurement.
Thus, for a given Q, s must be in the interval 0 < $ < Q.
t Here and in the remainder of the chapter, we revert to treating the demand and
Q and s as continuous variables.
44 DETERMINISTIC LOT SIZE MODELS CHAP. 2
since
rr = s/x
2
§['>*+ **’] 1
The average annual variable cost X, which includes the cost of ordering,
holding inventory, and backorders then becomes
ax = ax
dQ ds (2-18)
As we shall see, the optimal solution need not always satisfy (2-18). In
certain cases, the optimal $ may be on the boundaries.
Consider now the problem of finding the solution to (2-18). From (2-17)
we see that
dx i r i -j
dQ ~ Q2 ^ 2 s)2 2
jn
+ ~q (Q — s) =0 (2-19)
or
or
f («--)+irX + ^.-0 (2-22)
Q = + (* + Ic)s (2-23)
sec 2-5 CASE WHIM BACKORDERS ARE PERMITTED 45
«* [* b IV\ ‘ | nX } [(2X/1/0 (l } *
fr / 0 (2-20)
To determine Q explicitly, (2-2.1) is used to eliminate a in (2-21). This
yields
2 y i vkt( _
tv kA f t I IVQ i n
tv »x
tv ! 1 )l. + tv,Q
* * + IV I
Thus Q*, the optimal value of Q, m
46 DETERMINISTIC LOT SIZE MODELS CHAP. 2
In the event that the s value computed from (2-26) is negative, then the
optimal s lies on the boundary, i.e., s* = 0. We leave it for the reader to
demonstrate in Problem 2-62 that if xX > Kw, s* = 0. When it ^ 0, s*
cannot be infinite (why?). Equation (2-27) holds only when the s computed
from (2-26) is positive; otherwise, Q* = Qw. Furthermore, if the s com¬
puted from (2-26) is positive, it is the optimal s, i.e., s* does not lie on the
boundary s = 0. To prove this recall that when s = 0, Q* = Qw; however
dX/ds < 0 at s = 0, when Q = Qw and Kw > xX, so that s = 0 cannot be
optimal if Kw > xX.
When, x = 0, (2-26), (2-27) become respectively
y 0* _ r 2XAIC 1^/2
[_#(x + /C)J~ + IC)] v* (2-28)
and
* + IC l1'2
^ Q* = : (2-29)
It will be noted from (2-28) that when x = 0, then s* > 0 unless x = oo,
i.e., under optimal operating conditions, some backorders will always be
incurred. Give an intuitive explanation for this result. —
The computation of the reorder point for this model is, in principle, the
same as that presented in Sec. 2-3. Now, however, the inventory levels
used must be redefined. The on hand inventory is no longer appropriate
since there may be no inventory on hand, but instead there may be back¬
orders at the time when an order should be placed. The appropriate level
to replace the on hand inventory is the amount on hand minus the back-
prders, which will often be referred to as the net inventory. If there is
inventory on hand, there will be no backorders and this level will be
positive. If there are backorders, there wifi be no inventory on hand and
this level will be negative. In .terms of the net inventory level, the reorder
pomt is then rt - „ - mQ* - s* where, as before, m is the largest integer
less than or equal to r/T. [It is possible for r*h to be negative. This
that an order is placed when the backorders reach a level jr£[.. The level
to replace the on hand plus on order inventory of Sec. 2-3 is the amount
on hand plus on order minus baekorders. This will often be referred to as
the inventory position of the system. The reorder point in terms of the
inventory position is r* = M - s*; r* can also be negative!
I = 0.20; A = $5.00
r = 9 months
From (2-26)
The lead time demand is \i = 0.75 (200) = 150 units. Thus the reorder
point based on the inventory position of the system is
r* = M - s* = 150 — 5 = 145
To determine the reorder point rh based on the net inventory level, we note
that m, the largest integer less than r/T is 6. Thus
rt = m ~ - s* = 150 - 144 - 5 = 1
In the previous section it was assumed that all demands incurred when
the system was out of stock were backordered. We shall now examine the
lost sales case, i.e., the case where a demand which occurs when the system
is out of stock is lost forever. If demands occurring when the system
is out of stock are lost, it is no longer true that the annual revenues received
will be independent of the operating doctrine! They will depend on the
length of time for which the system is out of stock, and hence on the oper¬
ating doctrine. Thus we cannot immediately conclude that maximization
of the average annual profit will yield the same operating doctrine as the!
minimization of the average annual cost. We shall now show, howeverj
that with the proper definition of the stockout cost, the minimization of
48 DETERMINISTIC LOT SIZE MODELS CHAP. 2
the average annual cost will yield the same result as the maximization of
the average annual profit. Let S be the unit selling price of the item, <P be
the average annual profit, tt0 the cost of.a lost sale exclusive of the lost profit,
and C the unit cost of the item] Then if jo is the fraction of time during
which the system is out of stock, the average annual profit is !
:_A + ~ Xf
XT 2 Q + XT Q + XT
since on the average there are X/(Q + XT) cycles per year, the inventory
carrying cost per cycle is ICQ2/2X, and the cost of lost sales per cycle is
tXT. Geometrically, the behavior of the system can be represented as in
Fig. 2-8.
A necessary condition that f *, Q* be optimal is that they satisfy
P \ T C1 ~
— = 0 = - (Q + Xf)~* \*A + ~ Q2 + rX3f + (Q + \f)~W = 0 I
or
. XA , IC n
Xx = ~Q + Y Q (2'31)
and
provided that
0 < T* < oo, 0 < Q* < oo
2\A11/2
/ (2-33)
IC _
- - w t [(0 - <^5>
However, because of (2-34), it follows that T < 0 for both signs. Hence,
the optimal T again does not lie in the interval 0 < T < 00. In this case
50 DETERMINISTIC LOT SIZE MODELS CHAP. 2
the optimal value is f = 0, since (xX)2 > 2XAIC, i.e., the cost of running
the system with no lost sales is at least as cheap as that of running it with
any positive quantity of lost sales. In the special case where (xX)2 =
2XAIC, any value of f is optimal.
What we have shown above is that if the inventory system should be
operated at all, then it is never optimal to incur any stockouts. Even if
lost sales are allowed, it follows that when (xX)2 > 2XAIC, the optimal
solution is precisely the same as the optimal solution to the model studied
in Sec. 2-2. The results of this section can be seen intuitively as follows.
We rearrange the time sequence of events in Fig. 2-8 so that there are no
lost sales for a long time (i.e., in this region we have a situation like Fig.
2-1) and for a long time there is nothing but lost sales. This does not change
the average yearly cost. However, if (xX)2 > 2XAIC, and the Q in Fig. 2-8
is that given by (2-8), then costs can be reduced by stocking during the
period of lost sales and ordering in lots of size Qw.
In the previous sections it has been assumed that an order for Q units
will arrive in the inventory system as a lot of size Q units, i.e., all Q units
are received at the same time. We shall now consider a situation in which
the inventory system is the factory warehouse. It win be imagined that
t e item is produced in lots at the factory, and goes directly from the
actory to the factory warehouse. Once the factory is set up to produce a
lot it will be imagined that the production rate is * units per year (inde¬
pendent of the size of the lot). Itwill be supposed that demands are
deterministic and are incurred at the factory warehouse at a rate X units
per year. Both the number demanded and produced will be treated as
continuous variables. ^Clearly, the system cannot operate unless \[/ > X/ .
Let us compute the average annual variable costs when Q is the size'of
the lot produced. It will be imagined that there is a fixed setup cost A
for each lot produced. It will also be assumed that there is an inventory
carrying cost of 1C dollars per unit year where C is the unit cost of the item.
Ihe unit cost of the item will be assumed to be independent of the lot size.
irst the case will be studied where the requirement is made that all de¬
mands must be met from inventory, i.e., no backorders or lost sales are
permitted.
Durrng the periods when the item is being produced in the factory, there
will be a net rate of inflow ^ - X of units into the warehouse. During the
periods when the factory is not producing the item there is a net rate of
outflow X of units from the factory warehouse. If s is the quantity on hand
m the factory warehouse when the factory starts to turn out units, it is
SEC 2-7 THE CASE OF A FINITE PRODUCTION RATE 51
clear from the analysis of Sec. 2-2 that the optimal value of 5 is zero. It
remains to determine Q, the lot size, in such a way that the costs are
minimized.
The situation is illustrated geometrically in Fig. 2-9. The length of time
required to produce a lot is Tp = Q/ip. The on hand inventory in the
factory warehouse reaches its maximum value just as production is cut off
at the factory. The maximum on hand inventory is
- X) = Q ^1 - ^ (2-36)
-!(>-;) (2-37)
The length of a cycle is then T = Td + Tp = Q/\. •
The average annual costs of setup and holding inventory are
+ (MS)
10 (l
(2-39)
■reK(1-f)+£(1'
52 DETERMINISTIC LOT SIZE MODELS CHAP. 2
^ = 0 = 2 A -r 0 (2-40)
dQ U Q H)
which has the imique positive solution
n* _ f2 XA $ -|V2 r ^ “]i/2
The Q* of (2-41) does yield the absolute minimum value of K for Q in the
range 0 < Q < co.jlf & time r is required from the time the warehouse
submits an order tolhe factory until the first unit comes off the production
line then to compute the reorder point rh based on the on hand inventory
level, let m be the greatest integer less than or equal to r/T. If r — mT
< Td, the reorder point is rh = ^ — rnQ. However, if T — mT > Td the
reorder point is ’
Th
(2-42)
where i) =jfeT;We leave for Problems 2-30, 2-31, the task of sketching the
behavior of the on hand plus on order inventory and the determination of
tte reorder point in terms of the on hand plus on order inventory level.
^ t might be pointed out that it is sometimes difficult to decide what cost
C to use in computing inventory carrying charges at the factory warehouse.
The cost should be the variable production cost and should not include any
faxed production costs"]
Consider now the case where backorders are allowed. The cost of a
backorder will be assumed to have the form ir + fit where t is the length
of time for which the backorder exists. Let Q* be the optimal lot size and
s_ be the optimal value of the maximum number of backorders incurred
Rer cycle. Then it can be shown that if fi = Q s* =~lT8Foo andffs* = 0
Q is given by (2-41). When it 5* 0 ’
(2-43)
and
detailed discussion of when the optimal s lies on the boundaries of the re¬
gion. We also leave for Problem 2-29 the proof of the fact that if demands
occurring when the system is out of stock are lost instead of backordered,
then provided that the system should be operated at all, it is never optimal
to incur any lost sales.
This lot size is 15.5% greater than the lot size that would have been ob¬
tained assuming the production rate to be infinite, i.e., by use of (2-8).
The reader should answer for himself why the Q* for the finite production
rate is always greater than that for infinite production rates. Note that
this approach has ignored any holding costs on units in production. These
costs are independent of Q and hence do not appear in the variable cost
provided that the amount of time that each unit spends in production is
independent of Q. |
According to the above, the time between orders Q*/X = 0.298 years.
The on hand inventory in the warehouse reaches a maximum value of
m 745
= 0.0745 years = 27.2 days (2-47)
v 10,000
after the first unit comes off the production line. Since the lead time is
54
DETERMINISTIC LOT SIZE MODELS
CHAP. 2
r 2/12 - 0.167 years, it follows that the reorder point based on the on
nana inventory is
2-8 Constraints
Most real world inventory systems stock many items, not merely a single
item. It is permissible to study each item individually only as long as there
are no interactions among the items. There can be many sorts of inter-
fofeTch nfw11 lt6r F°r eXa“Ple> the it6mS “ay be Partial substitutes
netirT?offlh caPacity “ay be hmited and the items are com-
peting for floor space; there may be an upper limit to the number of orders
and the items are competing for these; or there may be an upper limit on
the maximum investment m mventory and the items are competing for in-
on^heffl om-0 arS' TT "T C°nSider CaS6S where there are constraints
nlaced ^ f***’^0* 0n the number of orders per year which may be
placed, and/or on the maximum dollar investment in inventory
of wSSfolff ^ °aSe t1616 there is “ Upper limit /t0 tbe square feet
onJun? J + °0?fCer ®UPP°Se that n items are beinS stocked and that
case whirl ST T UP/5' lqUare f6et °f fl0OT space- We sbaU study the
orU T V“Smet fr0m inventoly 80 no backorders
O lost sales are allowed. If Q,- is the order quantity for item j, then if the
floor space constraint is not to be violated at any time, it must be true that
n
^eltn8
phased tS certamty case
m the t0 so
T*that it** P0SsibiIitythat
will never be necessary0rders canthe
to have be
maximum quantity 0f each item on hand at the same time.
T be,the yearly demand rate (assumed to be deterministic), A, be
LdSl 6rm? C°Su Ci the Unit C0St (assumed to be independent of QA
L*an fcr the
2. A-i + ljCj
ilQs 2j
If the Qj of (2-51) satisfy (2-49), then these Qj are optimal. In such a case
the constraint is not active, i.e., sufficient floor space is available so that
average yearly costs could not be reduced by increasing the amount of
floor space available.
On the other hand, if the Qj of (2-51) do not satisfy (2-49), then the
constraint is active and the Qs of (2-51) are not optimal. To find the opti¬
mal Qj, the Lagrange multiplier technique is used. We form the function
dJ yt
dO = ® ^ ~~ f (2-54)
i=i
Qj* =
vljCj + 2d*fj i j l,...,n< (2-55)
where 6* is the value of 6 such that the Q* of (2-55) satisfy (2-54) The
function
6* dollars per square foot per year, then the average annual variable cost
would be
The set of Q} which minimize this cost expression are precisely the same as
those which minimize (2-50) subject to (2-49). The problem described by
{2-56) which assigns a cost but no upper limit to the amount of floor space
is called a dual of the problem described by (2-50) and (2-49), which does
not charge for floor space but has an upper limit on the amount of floor
space available.
Consider next the case where there is a constraint on the total number
of orders which can be placed.) Assume that no more than h orders can be
placed per year. This requires that
(2-57)
3 = 1 't*3
Here we ignore the fact that in any given year the number of orders placed
must be an integer and can differ from the average value X,/Q, by as much
as unity. It is assumed that this influence is small. We assume now that
there is no fixed cost per order. (In Problem 2-35, the reader is asked to
generalize the theory to be developed to the case where there is also a fixed
cost associated with each order.) The only costs are then the inventory
carrying charges. Thus the average annual variable cost is
(2-58)
3=1 *
n \
= 0-2a
= 1 **
7
_ 2V%-
, j = 1,... ,n
-£ ataT (2-63)
J
In this case it is easy to solve explicitly for the optimal value of the La¬
grange multiplier. The value tj* can be interpreted as the imputed cost of
placing an order.
Finally consider the case where there is an upper limit D to the dollar
investment in inventory at any one time. This constraint requires that
+- -») (2-65)
* It is conceivable that the constraints (2-57), (2-64) are inconsistent so that there
is no solution. Here we assume that they are consistent.
58 DETERMINISTIC LOT SIZE MODELS CHAP. 2
0 = _A t2-r.7)
&{Qi
= 0 = J C/h - D (2-tlS)
j-i
From (2-66)
VCi(Ij + 24,*y J
<2-691
ns :i r optiT1'......!i t
oSTie otr t* cr“'“' ”>th" .....
SLl When Z ““V.. r"rt ..
2SLZTw *r ' ■',IV“ thc ““S'nliiiK M... ,l„„r ,.
but ignoring the investment eonstmiut. It Ihe (J, „,Mv tlie
SEC. 2-9 CONSTRAINTS—AN EXAMPLE 59
investment constraint they are optimal. If they do not, we are then sure
that both constraints are active. Thus we introduce two multipliers and
solve the problem treating both constraints as being active. It is important
to note that it is necessary to examine the cases where one or both Lagrange
multipliers are zero, i.e., where one or both constraints hold as strict
inequalities.
Consider a shop which produces and stocks three items. The manage¬
ment desires never to have an investment in inventory of more than
$14,000. The items are produced in lots. The demand rate for each item is
constant and can be assumed to be deterministic. No backorders are to
be allowed. The pertinent data for the items are given in Table 2-1. The
carrying charge on each item is I = 0.20. Determine the optimal lot size
for each item. '
TABLE 2-1
Data for Example
Item 1 2 3
Demand rate (units per year) Ay 1000 500 2000
Variable cost (dollars per unit) C3 20 100 50
Set up cost per lot (dollars) Aj 50 75 100
= ylcs(i + 2p*j’ j= lj 2’ 3
ou
DETERMINISTIC LOT SIZE MODELS CHAP. 2
tiJrffi- D
\/ + 2p*
- 14,000 - Jj X 10*'
\0.10 + P!
FIGURE 2-10.
SEC. 2-9 CONSTRAINTS—AN EXAMPLE 61
FIGURE 2-11.
62 DETERMINISTIC LOT SIZE MODELS CHAP. 2
<2-ra>
If (2-73) is multiplied by X we obtain the quantity ordered each period, i.e.,
Q* = XT* = = qw
In the real world, it is not always true that the unit cost of an item is inde¬
pendent of the quantity procured. Often, discounts are offered for the pur¬
chase of large quantities. These discounts sometimes take the form of
price breaks of the following type: There are given quantities qo = 0, qi,
<b> ■ ■ ■ ,qm, qj < qj+i, j = 1, . . ., m and qm+i = <®, such that if a quantity
Q is purchased, q; < Q < qj+1, then the unit cost of each of the Q units is
SEC 2-11 QUANTITY DISCOUNTS—"ALL UNITS" DISCOUNTS 63
Cj, i.e., the cost of Q units is CjQ, and Cy+i < Ct-. The total cost of purchas¬
ing Q units then has the form shown in Fig. 2-12. We shall refer to quantity
discounts of this sort as “all units” quantity discounts since the discount
applies to every unit purchased.
The introduction of “all units” quantity discounts increases the difficulty
of determining the optimal order quantity for an item. Let us now examine
how it is done. Let us consider the case where backorders are allowed, and
v = 0,7T ^ 0. For any Q value, the optimal s is given by (2-23). Hence, if
is the unit cost of the item, the average annual variable cost optimized
with respect to s is
For q, < Q < qj+h Kj is the average annual variable cost. Note that it is
now necessary to include the cost of the units themselves in the variable
cost since the unit cost of the item depends on the quantity procured.
We can think of (2-74) as defining a cost curve for all Q, not simply for Q
in the interval q, < Q < qj+1. Jin this way one can obtain m + 1 cost
curves, one for each C/.' The important thing to notice is that these cost
curves do not intersect and that K}+1(Q) < Kj(Q) for all Q (proof?). A
typical set of these curves is shown in Fig. 2-13. The actual cost curve for
the system is represented by the solid portions of the curves shown in Fig.
2-13. The dashed portions are not physically realizable.
The problem is to determine the lowest point on the solid (broken) curve.
This can be done as follows: '.Firstjwe compute the Q value, call it Qim),
64 DETERMINISTIC LOT SIZE MODELS CHAP. 2
which yields the minimum cost on the Km curve. If Q(m) satisfies qm < Q<m),
then Q(m> is optimal since it yields the minimum cost on the Km curve and
no cost on any of the other Kcurves, j < m, can be lower than this cost.
If Q(m) < qm, then Q("> is not physically realizable. In this case compute
Km(qm) which is the cost on the Km curve at the with price break; Km(qm)
is to be used in the next stage.
To begin the second stage of the computation (if it is needed), deter¬
mine which is the point of minimum cost on the Km^ curve. If
Qm-i < Q^~l) < qm, compute and compare it with Km{qm). If
■Km—i^ then Q^m ^ is optimal since no cost on the curves
Kj,j<m — l can be lower and no realizable cost on the Km curve is lower.
When Km(qm) < ») then qm is optimal (if Km(qm) = Hm_1(Qtm_1))
then either qm or is optimal). In the event that lies outside the
interval gw < < qm, compute Hm_1(g„_1) and = min
LKm{qm), K^iq^)]. Let gm_i be the q,- appropriate to Km-U i.e., in this
case either qm or qm~i.
If the optimal Q was not determined during the second stage, the third
stage is begun by computing », the Q value at which Hm_2 takes on its
minimum cost. If Q<"-» lies in the interval gm_2 < Q<—» < qm_, compute ■
2>) and compare it with lL_i. If Km^2(Q(m~2)) < Km^ then
Q 18 °PtimaL If K^i > Km^2(Q(m~2)) then w is optimal (if Km_,
SEC 2-11
QUANTITY DISCOUNTS—“ALL UNITS” DISCOUNTS
65
n(2, = E [2(600)8
V ICt \ 0.20(0.28) ~ 414
GT/cJ
K2(q2) \C2 + ^A+IC2^ $200.80
y2 A
Q(1) = 406
3 , °U 1 *2^ 1 c
- i'-m -400 " a/ *-
and Q(0) is allowable since 0 < Q(0) < 500. Then
JTr v.a
Another type of quantity discount, which we shall call the incremental type
discount, charges C0 per unit for units 1, . . ., qh Ci per unit on units
Qi + 1, • • •, q2, etc. Geometrically, the total cost of Q units can then be
represented graphically as in Fig. 2-14. The total cost of Q units, C(Q),
when qj < Q < qj+1 can be written
Thus when no stockouts are allowed, the average annual variable cost if
Qi< Q < Qi+1 is
Kj = XCj + q (A+ R} - Cjqj) H ^ + ICj 2 — ICj ^ (2-76)
FIGURE 2-14.
SEC 2-12
INCREMENTAL QUANTITY DISCOUNTS
67
(The corresponding cost equation for the case where backorders are allowed
becomes somewhat complicated to write down. We leave the analysis of
this case to Problem 2-66. The general results described here hold, however
w en backorders are allowed.) We can imagine K,- in (2-76) to be defined'
for all positive Q even though K, is physically realizable only if
& < y < Qj+1. If there are m price breaks we then obtain m-f- 1 curves-
as shown in Fig. 2-15. The actual total cost curve is the solid portion of
these curves.
The computation of the optimal Q for incremental discounts is somewhat
FIGURE 2-15. -
For those Q(]) which are physically realizable, i.e., gy < Q® < qj+1, deter¬
mine Kj{QW). The Q(f> corresponding to the smallest of these costs is the
optimal Q. Note that for this case it does not follow that if Q(m) is physi¬
cally realizable, then is optimal (the following example illustrates a
case where Q(m) is allowable, but not optimal).
-M-60 t
Since Qco) is in the allowable range, we determine
~ \ ICiqi = $51,234
I
REFERENCES
PROBLEMS
2-1. The soft goods department of a large department store sells 500 units
per month of a certain large bath towel. The unit cost of a towel to
the store is $0.50 and the cost of placing an order has been estimated
to be $2.00. The store uses an inventory carrying charge oil = 0.17.
Assuming that the demand is deterministic and continuous, and that
70 DETERMINISTIC LOT SIZE MODELS CHAP. 2
The function is called strictly convex if the strict inequality holds for
a < 1. Geometrically, a function is convex if the line joining
two pomts on the curve y = /(*) lies on or above the curve. Illus¬
trate a convex function graphically. Show that the sum of convex
functions is also convex. Prove that a strictly convex function can
only have a single relative minimum in any interval, and that this is
also the absolute minimum in the interval.
2-7. Stow that UQ) = XA/Qisa strictly convex function, and that/,(Q)
- /CQ/2 is convex, for 0 < Q < =o. Thus show that the function K
defined by Eq. (2-6) is strictly convex. By use of the results of Prob¬
lem 2-6 show that dK/dQ = 0 can have only one solution in the
interval 0 < Q < «, and that the solution will yield the absolute
PROBLEMS 71
2-10. Prove in general for the case of deterministic demand and constant
lead times that the average amount on order is equal to the lead time
demand.
2-11. At a large automobile repair shop a certain part has a very low de¬
mand. The demand is 8 units per year. This can be assumed to be
deterministic and constant over time. The cost of placing an order
for this part is $1.00. The unit cost of the part is $30, and the shop
uses an inventory carrying charge of I = 0.20. Use the theory
developed in Sec. 2-4 to take account of the integrality of demand
and determine the optimal order quantity. What is the time between
placing of orders? What Q value would be obtained using Eq. (2-8)?
2-12. For the example solved on p. 37, what would be the optimal Q value
if the approach of Sec. 2-4 was used?
2-13. From the theory of the backorders case discussed in Sec. 2-5, show
that the average on hand inventory can be written iQ — s + B(Q, s)
where B(Q, s) is the average number of backorders (this average is
to be taken over all time not merely for the time over which back¬
orders exist).
2-14. From the theory of the backorders case discussed in Sec. 2-5, show
that the average annual cost of backorders arising from the M term
is simply f times the average unit years of shortage incurred per year.
72 DETERMINISTIC LOT SIZE MODELS CHAP. 2
Furthermore, show that the average unit years of shortage per year
is numerically equal to the average number of backorders.
2-15. Introduce the reorder point r explicitly into the model of Sec. 2-5 and
eliminate the variable s.
2-16. For the material presented in Sec. 2-5 show that the optimal values
of Q and s will be independent of whether one bases the inventory
carrying costs on the quantity on hand or the quantity on hand plus
on order.
2-17. Carry out in detail the derivation of Eqs. (2-26) and (2-27).
2-18. Show that if the inequality sign is replaced by an equality in Eq.
(2-25) then any s > 0 is optimal, i.e., show that the same cost will be
obtained for any s provided that one optimizes over Q.
2-19. Consider the case studied in Sec. 2-5 where backorders are allowed.
Instead of taking the cost of a backorder to be ir -f- fit, assume in¬
stead that the cost of a backorder has the form aebt, a, b > 0, where
t is the length of time for which the backorder exists. Derive the
equations which yield the optimal values of Q, s.
2-20. For the model discussed in Sec. 2-5 determine the fraction of the time
that the system is out of stock when the optimal values of Q and s
are used. Assume that i = 0, r ^ 0.
2-21. A chemical company produces a certain organic chemical in batches.
The annual demand rate for this chemical is 100,000 pounds. The
demand can be considered to be known with certainty and the rate
does not change with time. The fixed cost of producing a batch is
$500. The variable cost of production is $2.00 per pound. There is
a backorder cost of $5.00 per pound per year, and no fixed cost of
a backorder, i.e., t = 0. Determine the optimal batch size and the
optimal number of backorders to incur.
2-22. Show that if Eqs. (2-28), (2-29) for the case of x = 0, * * 0, are
substituted into Eq. (2-17), the minimum cost is
no* _ 77"
r /rni/2
— JS-u 1 + j- - 1C\*(* + IC)]- 1/2
2-25. Consider an item with the characteristics listed below. Solve for the
optimal values of Q and s. Compute the average annual variable cost
using Eq. (2-17) and also the equation of Problem 2-22.
2-26. Solve the preceding problem for the case where w = oo. Compute
the average annual variable cost and compare with the results of the
previous problem.
2-27. The inventory turnover rate is a variable frequently referred to,
especially in retailing. The turnover rate is defined to be the annual
demand divided by the average inventory. Give an intuitive inter¬
pretation of the turnover rate. Compute the optimal turnover rate
for the case where no stockouts are allowed and the case where
backorders are permitted. Hint: What is the average inventory in
the backorders case?
2-28. Derive in detail Eqs. (2-43) and (2-44). When will s* = 0 if # ^ 0?
2-29. Examine the situation in which the production rate is finite and lost
sales are allowed. Show that, as in the case where the production
rate is infinite, it is never optimal to have any lost sales.
2-30. Let r be the production lead time for the finite production rate case.
Obtain the reorder points, both for the no stockout case and the back¬
orders case, in terms of the various inventory levels of interest.
2-31. Sketch on the same graph for the case of a finite production rate and
no stockouts the behavior of the on hand inventory and the quantity
on hand plus on order.
2-32. Re-solve Problem 2-3 assuming that the production rate of detergent
is 400 tons per month.
74 DETERMINISTIC LOT SIZE MODELS CHAP. 2
2-37. Derive in detail the equations which determine the optimal order
quantities for a number of different items when the items are coupled
together through a constraint on the available floor space and a
constraint on the maximum investment in inventory. Assume that
no stockouts are allowed.
2-38. Derive in detail the equations which determine the optimal order
quantities for a number of items when the items are coupled together
through a constraint on the number of orders, a constraint on the
available floor space, and a constraint on the maximum investment
in inventory. Assume that no stockouts are allowed.
2-39. Show what modifications are needed in the theory developed in the
chapter to handle a constraint on the maximum investment in inven¬
tory if the constraint is on the average investment in inventory.
2-40. A small shop produces three machined parts 1, 2, 3 in lots. The shop
has only 700 sq. ft. of storage space. The appropriate data for the
three items are presented in the following table.
PROBLEMS 75
item ® © ©
X (units/yr.) 5000 2000 10,000
A (dollars) 100 200 75
C (dollars) 10 15 5
fi (sq. ft./unit) 0.70 0.80 0.40
item ® © ©
space requirement | ) 5 10 8
\ umt J
76 DETERMINISTIC LOT SIZE MODELS CHAP. 2
2-46. The purchasing agent for the G. W. Dog Food Company can buy
horsemeat from one source for $0.06 per pound for the first 1000
pounds and $0,058 per pound for each additional pound. The
company requires 50,000 pounds per year. The cost of placing an
order is $1.00. The inventory carrying charge is I = 0.25. Compute
the optimal purchase quantity.
2-47. Derive in detail Eq. (2-74).
2-48. A supplier of made to order metal fixtures quotes the cost of Q units
as 400 + 25Q dollars. The purchaser of these fixtures requires 2000
units per year. He uses them at a uniform rate. His direct costs for
making out purchase orders, receiving, inspection, and other costs
that are incurred on each shipment amount to $20 per order. The
carrying charge is 7 = 0.20. The time elapsing between the placing
of an order and the delivery of the units to the stockroom is three
months. Calculate the economic purchase quantity and the reorder
point (based on the on hand inventory level).
2-49. The XYZ Corporation has a special petty cash fund it uses for a
known steady state flow of contributions to charity. The corporation
periodically replenishes this fund by withdrawing cash from its
savings bank account (which pays 4 percent interest on deposits).
The fund is never allowed to run out. A bank messenger is paid $5 00
per delivery of cash from the bank. If the optimal withdrawal sched¬
ule involves nine withdrawals per year, what is the dollar volume of
annual contributions to charity?
2-50. The fee which a particular brokerage firm charges its clients is 20
+ 0.01 V dollars, where v is the market value of the stocks purchased
or sold on any given day. A client of the firm spends $10,000 annually
over and above the 8 percent interest he receives from his invest¬
ments. Assuming that the client accumulates the funds for invest¬
ment at a uniform rate, how much should he accumulate before
investing, i.e., what is the optimal value of v?
PROBLEMS
77
2-57. An item in inventory is sold at a unit price of p, the rate of sales being
X umts per year. If a quantity Q is ordered the unit cost is 5 + (a/Q)
The cost of placing an order is A and the inventory carrying charge
78 DETERMINISTIC LOT SIZE MODELS CHAP. 2
for Eqs. (2-8), (2-11) respectively. What does it say about the per¬
cent change in Q* for small percentage changes in A. For the ex¬
ample on p. 37 determine the approximate change in Q* if A in¬
creases to $8.50. What is the approximate change in X*? What are
the exact changes in Q* and A*? Derive formulas similar to the
above for changes in X, C, and /.
2-64. An inventory system stocks n items. It has been found necessary to
review and place orders for all items at the same time. Let T be the
time between reviews, A the cost of review and of placing orders for
all n items, I the inventory carrying charge, Xy the annual rate of
demand for item j and Cy its unit cost. Determine a formula for the
optimal value of T by minimizing the average annual variable costs
of review and holding inventory for all n items under the assumption
that no stockouts are allowed. Find the optimal T for a system which
carries items for which the appropriate parameters are: A = $75,
PROBLEMS 79
2-68. Consider a firm which has several different warehouses, and let the
demand rate at warehouse i be A,-. It is possible to operate this
collection of warehouses on a decentralized basis, by having each
warehouse order for its own needs using its own order quantity and
reorder point, or the system can operate on a centralized basis, with
a system reorder point and an order quantity for the entire system
which is allocated in the appropriate amounts. Prove that when
the unit cost of the item is constant, the carrying charge is the
same at each warehouse, the ordering cost and lead time are inde¬
pendent of whether the system is operated on centralized or de¬
centralized basis, and when transportation costs from the source to
the warehouses do not depend on the mode of operation, then the
centralized mode of operation never leads to higher costs than the
decentralized mode of operation. Would the conclusion be altered if
quantity discounts were offered? Assume in the analysis that stock-
outs are not permitted.
d = H e~inT
n—0
Determine the value of Q which minimizes 3. Expand the equation
which determines Q* in powers of i and show that to a first approxi-
mation
n* = / ~~ __
W \(J0 + i)C~Qv>
How does the frequency with which orders are placed affect the good¬
ness of the above approximation? As a specific example consider the
case where A = $15.00, C = $20.00, J. = 0.10, i = 0.10, X = 1000
units per year. Determine the exact value of Q which minimizes 3
and the value Qw. Pick another point other than the point when an
problems
81
r* r:!.. ^-ssesesk
, 1 , an ord< r- U hieh 'vas used in the above? Hint- Tt ran
difficult, to wilve explicitly the equation db/dQ » 0 for ^'lI oZit
mg the limiting form, determine the coefficients of A C hC and
r ('.«ml “ »
r rcr r? “s •***£ &
'■ "my '* w"r,“ "““'y *» ‘h»t
PROBABILITY THEORY
John Gay
3-1 Introduction
We have noted previously that the demands for units stocked by an inven-
ory system can seldom be predicted with certainty. Instead, they must be
described in probabilistic terms. Realistic inventory models must account
for this uncertainty in demand. Often, it is also true that procurement
lead times must be described probabilistically. It is the purpose of this
chapter to present the background material from the theory of probability
and stochastic processes that will be needed in the future chapters. It will
e assumed, however, that the reader has previously had at least a brief
introduction to the theory of probability.
P(A,.) = *
Since n > 0, n{ > 0 and »< < n, (3-1) follows immediately from this
definition. If m, outcomes have the properties At and Ah then the number
of outcomes which have the property At or A, is m + % - nij and hence
Inj = p(Ai\Aj)
This is (3-3). Thus the three basic laws have been derived in this case.
Often it is convenient to imagine that the n outcomes can be represented
symbolically by points in a space. The resulting space is called the sample
space. If we think of the outcomes as being represented by points in a
plane, then it is possible to draw diagrams such as Fig. 3-1. The sample
space will be assumed to consist of the points (represented by crosses)
inside the rectangle. The set of outcomes with property At is the set of all
points inside the crosshatched area marked A{. Similarly, the set of out¬
comes with property A} is the set of all points inside the crosshatched re¬
gion marked As. The set of points with the properties A> and As are those
in the doubly ruled area, i.e., these points represent the set of points with
the property At D Aj. The set of points with the property At (j Aj is
the set of all points in the shaded areas. Fig. 3-1 makes it especially easy
to derive the formula (3-2). If Ai} As are mutually exclusive, then in Fig.
3-1, the areas representing A{ and As do not intersect.
86 PROBABILITY THEORY AND STOCHASTIC PROCESSES CHAP. 3
FIGURE 3-1.
n
= T ?(*) (3-7)
x—0
If the experiment is performed many times, then the fraction of the time
that the outcome can be described by the integer x will approach p(x).
The outcome of the experiment can then be thought of as being described
by the variable x, &n.dp(x) can be looked upon as a function of x which gives
the probability of x. The variable x which can only take on the non¬
negative integral values 0, 1, . . ., n is called a random variable since it
describes the outcome of an experiment, the outcome of which is deter-
SEC 3-3
DISCRETE RANDOM VARIABLES
87
Note that r cao take on the values 0,1.», the same values which a
n assume.. Thus we can define a new function P(x), which will be called
e—^ Probafiility of x or the cumulative function for x- it is the
i’mftoJ r'T™ experimerit -™lds a value less than or
£ r °n W*“Cl‘ ^ gently be of use to us is some-
times called the complementary cumulative function. It is defined by
P(x) is the probability that the outcome of the experiment will yield a
value greater than or equal to *. We shall find it convenient to use the
for each event, i.e., p*(l — p)n~x. This same probability is obtained if we
ask for the probability of x successes and n — x failures in any specified
order.
Next, let us ask what is the probability of having x successes in n trials
without regard to the order in which the successes are obtained. Note that,
events corresponding to obtaining x successes in different orders are mutu¬
ally exclusive. Hence, the probability of having x successes without regard
to the order in which they are obtained is simply the sum of the proba¬
bilities for each possible order. The probability for x successes in any
given order is px( 1 — p)n~x. Thus, all that remains is to compute the
number of possible orders in which x successes can occur. This is simply the
number of ways in which n objects (the trials) can be separated into t wo
groups (without regard to order within a group) such that x objects are in
one group and n - x in the other. Of course, this is the number of combi¬
nations of n things taken x at a time, i.e., n!/x!(n - x)!. Sometimes we
shall use the simplified notation (n) to represent, a! x!(«. ■ x)!. We con¬
p(x) = b(x;n,p)
50- [p + (i - p) I" i
by the binomial expansion. Note that, b(xis the xth term in tin*
binomial expansion of [p + (1 - p)]» We shall always use the notation
b(x; n, p) for the binomial density function. If the random variable x has
the density function b(x; n, p), then x is said to have a binomia l dist ribut ion.
Note: We shall use frequently the binomial expansion and hence it might
be helpful to review briefly its derivation. The binomial expansion of
(x + b)* for any real number n is a special case of the Maelaurin expansion
/(*) = E h/o,(0)xt
and Hence
m = 6- + £ _V * • • V/* .7 T I) _
Jf-2 £3?>n“>
y=i
(x
+ ~ (j) xihn n a positive integer.
= ( ^/(M+j-l)!
^ ; (Ini-11!
(M - 1)!
and so
(x
3!
= 3
211!
Although for the case of finite n we could prove (3-7) from the fundamental
laws, we cannot prove (3-12) directly. Instead we take (3-12) as a postu¬
late.
A particularly interesting density function which is defined for all non¬
negative x is called the Poisson density. It is defined by
p(x‘,n) = e~»
e*
n\
o-?1
’-set-T1)*1-*
Multiplication by P» shows that the sum of 6v(x; n, p) from * = 0 to « is
unity as desired.
A random variable having a negative binomial density can arise in a
variety of ways. One simple probability problem which gives rise to such
a distribution is as follows: Suppose that we toss a coin which has a
probability P of getting a head until we obtain n heads. We then ask:
„tb? Probablb1f that exactly * + n tosses are required to obtain
n heads. This means that n — 1 heads are obtained in the first x + n - 1
tosses and a head is obtained on the (x + n)th toss. The probabUity of
bilitv nf l mA + * t0SS6S is6(n“1^ + n-l.P) and the proba-
bihty o a head on the last toss is p. Hence, the desired probability is
pb(n - 1; x + n - 1, p), which is (3-15), and the random variable x (not n)
nas a negative binomial distribution.
When n = 1 in (3-15) we obtain
FIGURE 3-2.
SEC 3-4 CONTINUOUS RANDOM VARIABLES 93
= (3.19)
dF(x) dF(x)
dx = /(*), dx
= -f(x)
for v from the known density function for x. We shall assume that v = 8(x)
is invertible, and that the inverse has a continuous derivative so that we
can write x = f(v), (possibly with different f functions needed for different
ranges of x), and dx = ^'(w) dv where = df/dv. Then
f(x) dx = dv
where the limits on the second integral are those corresponding to x = —oo
and x = oo respectively. The range of values for the variable v may not
be from -« to ». Suppose that the range is from wmin to iw Then after
we have converted the second integral in (3-20) to the form
Cv max
Hv) > 0, h(v) will be the density function for the random variable v.
In the event that the transformation from a: tow is one to one, and
t 7u°’,S0 that Wmitt = 0(-co) and W = «(«), then h{v) = f[f(v)W(y).
If the transformation is one to one and < 0 so that vmin = fl(oo)
and w - «(-»), then in going from (3-20) to (3-21) the limits of inte¬
ntion must be interchanged so that h(v) = In general, if
me transformation from v to x is one to one, h(v) = fbP(v)]\^'(v)\
When *(») is not unique so that different * functions are needed for
. ferent 1’anges of x values, then the right-hand side of (3-20) must be
m erpreted as the sum of two or more integrals, one for each integrated
over the appropriate range of v values. An example involving a non-unique
v function will be presented below.
We shall make use of several continuous distributions in our later work
Ihe one to be used most frequently will be the well-known normal distri¬
bution for which the density function is
V2^X„ dx = 1 (3-23)
“2/2 dw J_ e~“2/2 du
and the product of two single integrals has been converted into a double
integral. Then write w2 + u2 = r2, w = r cos 6, u = r sin 0, so that on
changing to polar coordinates
P= Jo r6~T,/2 dr d6 = / Te~rV2 dr = jj #= 1
This proves that (3-23) is correct.
Another continuous distribution which we shall use occasionally is called
the gamma distribution. The gamma density function is defined by
«! ’ -0 (3-25)
[o , x < 0
where a is a non-negative integer and p is any positive number. Note that
^Fis case the random variable is restricted to being a non-negative
number it can never take on negative values. The function y(x; a, p) is
called a gamma density of order a + 1.
To show (3-18) holds we observe that
1
(a - 1)! JQ yCt-le-U
converges to a finite positive value for any real value of oc > —1 The
value of the integral depends only on a and is denoted by r(« + 1) The
SS "r11^ ^ fmma function of « + L Thus it is possible to
generahze the gamma density (3-25) to any real a > -1 if a! is replaced
ylt“ + 1-) When “ 1S a non-negative integer, then r(<* + 1) = a\.
{fterfe, x > 0
[0 , x < 0
$(x) = P [ = 1 — e~&x
Jo ' /
The complementary cumulative is
E(x) = 1 - = e-Bx
The tmctiom «(«;». Eiz), £(*) Me plotted in Pig, 3-3 to the ca* of
SEC 3-4
CONTINUOUS RANDOM VARIABLES 9
FIGURE 3-3.
1
V 0-/2)Q—0 fly V (l/2)e-V fly y-( 1/2) e-v fly
2vV i:
V- = X) ^0) (3-27)
x=Q
The number n defined by (3-27) will be called the mean or expected value
of x. It is also called the mean of the probability distribution of x.
Let us now consider some function of the random variable *, say 6{x)
( or any inventory model x might be the demand in a given year and 6(x)
tiie cost of operating the system). Note that 6{x) is also a random variable.
Then if we find the average value of 6(x) by repeating the experiment a
large number of times, as the number of times which the experiment is
repeated approaches infinity, we expect the average value of 6(x) to tend to
the number
Y^e(x)p(x) (3-28)
which will be called the expected value of 6(x). Note that in general, the
expected value of 0(x) will not be 8(n), ju being the expected value of x.
. In event that the random variable x can take on any non-negative
integral value, it is only necessary to replace the upper limit n on the sum¬
mation signs m (3-27) and (3-28) by » to obtain the appropriate expected
values for this case. If the expected value is to have any meaning when x
SEC 3-5 EXPECTED VALUES 99
is allowed to range over all non-negative integers, the sums (3-27), (3-28)
must converge. In all cases of interest to us they will.
When 6(x) — (x — p)2, p being the mean of x, then (3-28) is called the
variance of x and is denoted by <r2. The variance of # is a rough measure of
the sort of spread in x values about the mean that can be expected. If <r2 is
large we can expect a greater spread in the x values than if a2 is small. The
number <r = V<r2 is often called the standard deviation of x. The number
<r2 is also called the variance of the distribution of x.
By its definition
a2 = (x - p)2p(x)
2=0
= Z
2=0
z2p(a:) - 2m2 + m2 Z x2pOO - p2
2=0
(3-29)
- t - <-)*-
In the first step we cancelled the x in xl to yield (x — 1)1 and factored out
np. In the second step we made the substitution u — x — 1. The resulting
summation is simply the binomial expansion of [p + (1 — p)]n_1. Hence
the mean of the binomial distribution is np. This is to be expected intui¬
tively, since if p is the probability of a success on a single trial, one would
expect the average number of successes in n trials to be np.
100 PROBABILITY THEORY AND STOCHASTIC PROCESSES CHAP. 3
Note that the mean and variance of the binomial distribution depend on
born n and p. *
Consider next the Poisson distribution. We first note that the mean is
(3-32)
12 (3_33)
(3-34)
-nJLfAt(uV‘)^-^
r «ls=n A
SEC 3-5 EXPECTED VALUES 101
Furthermore
oo
<r2 = n, p) - p2
x—0
n(n + 1)(1 - pY = P?
~~ M
M P2
The geometric distribution is a special case of the negative binomial for
which n = 1. Thus for the geometric distribution, as defined by (3-16)
1 — P
(3-36)
P '
/: 6(x)f(x) dx
(3-37) is called the expected value of 6(x). The mean p (when 0(x) = x)
and variance <r2 [when 0(x) = (x — jt)s] of x and of the distribution of x
are then defined to be
Let us now compute the mean and variance for the normal and gamma
distributions defined in Sec. 3-4. For the normal distribution given by
(3-22) the mean is
i r
=— / xe~ (1/2*’) Or-*)’ dx _1 f (X-p[ e-(l/2*>)(x-
dx + p
2ir <r J—<*
V/2jt *v 2x J — ® O'
+ + (M8)
102 PROBABILITY THEORY AND STOCHASTIC PROCESSES CHAP. 3
where
1 (x - ixV
V 2\ <r )
Thus the parameter m in the definition (3-22) of the normal distribution is
its mean.
The variance of the normal distribution is
(x - dx = ~= / w2g-(”!/2) dw
a v 2t ./-»
Now integrate by parts writing u = w,dv = we~wi/2 dw so that v =
—e~w2/2. This yields
7= / w2e~~w2/2 dw
T(a+ 1) = / dx wae~w dw
We shall now compute the mean and variance for the gamma distribu¬
tion in the case where a can be any real number greater than — 1 and 0 is
any real positive number. From the generalization of (3-25) with T(a+ 1)
replacing a!, we see that the mean is
M = rJ*(Ma e-?xdx
Jo T(a +1) “
(3-42)
_ T(q + 2) 1 f00 f3((3x)a+1 a 4 1
~
e~Px dx =
T(a + 1) 0 Jo T(a + 2)
SEC 3-5 EXPECTED VALUES 103
by (3-41). Similarly, the variance is given by (we leave for Problem 3-22
the proof that the equivalent of (3-29) holds for continuous random
variables)
(a + 2)(a+ 1) a -f- 1
R2
m= ;S2
To express a, /3 in terms of n, a note that on using (3-42) in (3-43)
P=£ (3-44)
SO
a = (;)' - 1 (3-45)
limy
t—+ oo t i: x(t)dt
dently, i.e., the stochastic processes for the various systems operate inde¬
pendently of each other. With this picture in mind, the fraction of the
systems which have demands of precisely x will approach p(x) as iV —> oo.
Thus if one computes the expected demand in accordance with the rules
given in the previous section, this expected demand will be the ensemble
average demand, i.e., the average over the ensemble of systems of the
week’s demand.
Let us now imagine that the stochastic processes associated with a sys¬
tem do not change with time. Again we can imagine that we have an
ensemble of N of these (identical) systems operating. After the ensemble
of systems has been operating for a long time, it will reach a condition of
statistical equilibrium. This means that the ensemble average of any
relevant variable will become independent of time. When statistical equi¬
librium is reached, then, for any random variable, an ensemble average
over the ensemble of systems will be precisely the same as a time average
over all time for a single system,* so that p\x) can be interpreted as the
fraction of the time that a single system will have the value x, or the frac¬
tion of the systems in the ensemble that will have the value x at any given
point in time.
The notion of an ensemble average is sufficiently general so that every
expected value used in this book can be thought of as an ensemble average,
and every probability as the limit as the number of systems in the ensemble
approaches infinity, of the fraction of the systems in the ensemble which
yield the given value of the random variable under the conditions specified.
When the stochastic process under consideration does not change with
time, an expected value can be interpreted both as a time average and an
ensemble average. Furthermore, in this case, the expected value will be
the same regardless of which way it is computed. Similarly, when the
stochastic process does not change with time, a probability statement can
be interpreted as referring to the long run fraction of the time that the
random variables will have a specific value, or the fraction of the systems in
the ensemble that will have the value.
The equivalence of time averages and ensemble averages in this case is, roughly
speaking, a statement of the celebrated ergodic theorem of statistical mechanics. No
mathematically satisfactory proofs of the ergodic theorem were available until 1931
when proofs were provided by G. D. Birkhoff and John von Neumann.
SEC 3-7 PROBABILISTIC DESCRIPTION OF DEMANDS 107
mantis does not change with time. Let us begin by noting that the number
of units demanded in any time period will depend on the time between
demands and on the number of units demanded when a demand occurs.
In the real world, both the time elapsed between demands and the quantity
demanded can be random variables. The time t between demands will be a
continuous random variable and the number of units demanded when a
demand occurs will be a discrete random variable.
Consider first the random variable t representing the time between the
occurrence of demands. Let G(t) be the probability that a time greater
than or equal to t elapses from the time of a given demand until the time of
the next demand. In general G(t) could be a function of the tirmw of occur¬
rence of all previous demands, of the quantities demanded at each of the
previous demands, of calendar time, and perhaps many other things. We
shall here restrict ourselves to the case where G(t) depends only on the tima
since the last demand, and not on the times of any other demands, quan¬
tities demanded, or calendar time.
If we were to plot G(t) as a function of t, it might have one of the shapes
shown in Fig. 3-6. Then G(t) must be unity at t = 0, since the probability
is unity that the demand under consideration will occur after the previous
demand. Furthermore, we assume that it is certain that another Hamanrl
will occur at some future time, so that G(t) must approach 0 as t approaches
oo. In the event that G(t) can be represented by curve 1 in Fig. 3-6, then
the time between demands is deterministic, i.e., the time between demands
108 PROBABILITY THEORY AND STOCHASTIC PROCESSES CHAP. 3 S
is always t0. Curves such as those shown in Fig. 3-6 can often be represented
quite well by the complementary cumulative function of one of the gamma
densities for an appropriate choice of a, 0. The distribution of the time
between successive demands is often referred to as the interarrival dis¬
tribution. h
A form of G(t) which will be of greatest interest to us will be that where
G(t) is the complementary cumulative function for the exponential dis¬ V
tribution (see example on p. 94), i.e.,
T
Git) = <rXf (3-46)
When G(t) has the form (3-46), we say that the times between demands
are exponentially distributed. The density function for t is S.
dC
= Xe~x< *e(Jt’x) (3-47)
T
and the probability that the next demand arrives between t and t + dt
after the present demand is e(t) X) dt. The average time t between demands
is
t = Jq \te~u dt = i (3-48) T
th
so that the average number of demands per unit time is X. tr
If G(t) is given by (3-46), it is not difficult to compute the probability (3
Vn(t) that precisely n demands occur in a time period of length t imme¬
diately following the occurrence of a demand. This is most easily done by
obtaining a recursive relation for the Vn(t). Note first of all that
This is true ami can easily lie proved by induction. Wo have already shown
that (d-o2) is true for n 0 and « 1. It remains to show that if it is
true for « m 1, it is also true for « m. Tills follows, since if w<i use
(2-f>2) evaluated fur n m 1 in Cl-ol.), we obtain
A"c K‘ (A t)m
{t — r)w 1 dr hi
1)! «; ml *'
p{B\A) = = \dt
U0(t) = (3-54)
Recall that we began our discussion by assuming that the probability that
the time from one demand to the next is greater than or equal to t is
SEC 3-7 PROBABILISTIC DESCRIPTION OF DEMANDS 111
Note that Tn-i rather than Un-1 appears, since the time interval beginning
at time r and extending to time t begins immediately after the occurrence
of a demand. Thus on averaging over r
and Un(t) also has a Poisson distribution with mean \t. We have proved
then that the probability of having n units demanded in a time interval of
length t is p(n; Xi), and this is independent of whether we begin observation
immediately after the occurrence of a demand or at a random point in time.
The equality of the Un and Vn is a peculiar feature of the exponential inter-
arrival distribution and does not hold for an arbitrary interarrival distri-
bution. When the interarrival distribution is exponential, we sometimes
say that the demands are generated by a Poisson process. An important
characteristic of the Poisson process is that it has no memory, i.e., the
probability that a demand occurs in a time period of length dt is X dt inde¬
pendently of when any previous demands occurred.
Let us suppose that a Poisson process is generating demands and that
units are always demanded one at a time, so that the quantity demanded is
always unity. Then Un(t) is also the probability that n units are demanded
in a time period of length t. Thus, in this case, the probability of having
112 PROBABILITY THEORY AND STOCHASTIC PROCESSES CHAP. 3
Hence the density function for the time until the Qth demand is a gamma
density with a = Q - 1, i.e., a gamma density of order Q. We have just
shown that if for some inventory system a Poisson process generates de¬
mands, and if units are demanded one at a time, then if an order is placed
each time Q units are demanded, the density function for the time between
the placing of orders is a gamma density of order Q.
In the event that the number of units requested when a demand occurs,
i.e., the quantity demanded, can vary randomly from demand to demand,
the task of determining the probability that any given number of units will
be demanded in a time period of length t becomes much more complicated.
We shall illustrate this by studying one of the simplest possible cases in
SEC. 3-7 PROBABILISTIC DESCRIPTION OF DEMANDS 113
where p(n\j) is the probability that precisely n units are demanded when j
demands occur. Equation (3-58) does not hold for n = 0. The probability
that no units are demanded is the probability that no demands occur. Thus
* Stated differently, if y is the demand in excess of one unit, then y has the geometric
distribution (3-16).
Thus
The Un{t) given by (3-59), (3-61) are referred to as the stuttering Poisson
distribution. It is rather difficult to work with this distribution, which is
one of the simplest representing situations in which more than a single unit
can be demanded at a time. Fortunately, as we shall see shortly, there is a
simple approximation to all these distributions in terms of a continuous
random variable which lends itself quite easily to analytical work and which
often provides a sufficiently good approximation in practical applications.
12 12 p(*» y) 1
x=0 y=0
= (3-62)
OO
The functions p(x), p(y) are called the marginal densities for the random
variables re and y respectively. Note that they are legitimate densities
smee p(x) > 0, p{y) > 0 and from (3-62)
Z p(?) = h Z p(y) = 1
x=0 y—0
If either or both of the above random variables are defined only over a
finite number of integral values, the summations are taken only over the
allowed values of the variable. When p(x) ^ 0 and p(y) ^ 0, the con¬
ditional probabilities p{x\y) and p{y\x) are defined by
Consider next the case where 2 is a discrete random variable defined over
the non-negative integers and y is a continuous random variable defined for
all real y. We then take f(x, y) to be the joint density function for x and y;
f(%, y) dy is the probability that the discrete random variable has the value
x and that the continuous random variable lies between y and y + dy.
Then, since some pair of values must occur, we have
Then p(x) is the probability that the discrete random variable has the value
x} and v(y) dy is the probability that the continuous random variable lies
between y and y + dy.
Next when v(y) ^ 0, p(x) 0, we define conditional density functions
g(y\x) and h(x\y) such that
Then h(x\y) is the conditional probability that the discrete variable has the
value x given that the continuous variable has the value y, and g(y\x) dy is
the conditional probability that the continuous random variable lies be¬
tween y and y + dy given that the discrete random variable has the value x.
Then (3-68) can be written
/ CO
h(x\y)»(y) dy,
go
of units demanded in a time t is p(x; \t) where X is the mean rate of demand.
Thus
y(t; a, 0)
my e -fit.? p(x; \t) (3-72)
T(a + 1)
my
v(x\ — [ PO'O-—iMil—e-os+x)(^ = —ft--—— [ ix+ae-w+\)t
P{X) ~ Jo xl F(a + 1)
L)°6 ' M x\T(a + 1) Jo 1 6
(3-73>
by (3-26). Thus if a is a positive integer, the marginal distribu¬
tion of lead time demand p(x) has the negative binomial distribution
bxlx; a + 1, 0/(0 + X)]. There is no reason why a must be an integer in
the definition of the negative binomial distribution. We can consider (3-73)
to be a generalized definition of the negative binomial distribution which
includes cases where a is not an integer. We have shown, therefore, that if
the lead time is gamma distributed, and demands are generated by a
Poisson process, units being demanded one at a time, then the marginal
distribution of lead time demand is a negative binomial distribution.
It remains to consider the case where the two random variables x, y are
continuous. Now we define a density function/(x, y) such that f(x, y) dx dy
is the probability that x lies in the interval x to x + dx, and y lies in the
interval y to y + dy. Then it must be true that
f(x, y) dx dy = 1 (3-74)
As a special case, imagine that 6(xi, x2) depends only on x%, i.e.,
6(xi, x2) = iP(Xl). Then from (3-78) the expected value of x//(xi) is
“ » »
£ £ Kxdpixi, X2) = £ *60 £ p{xh x2) = £ t(xi)p(xi) (3-79)
zi-u x% u XI = 0 XI = 0 X1 = 0
where p(xi) is the marginal density for xx. As another special case imagine
that 6(xi, x2) = Xi ± x2. Then the expected value of xx ± x2 is
00 oo ^
where mi, M2 are the expected values of xi, x2 respectively. More generally,
the expected value of a finite sum of random variables xx + . . . -|- xn is
the sum of the expected values, i.e., mi + . . . + m». This is'proved by
letting the event x% = x2 + . . . + Xn and applying (3-80) to x± + x%. Then
the argument is repeated on xt Relations like (3-78), (3-79), (3-80) also
hold if one random variable is discrete and the other continuous, or if both
are continuous, provided that the appropriate summation signs are replaced
by integrals. Similarly, the expected value of
Zl±X2d= . . .=tx„ is Ml ± M2 ± . . . ± nn
3-9 Convolutions
Consider the discrete random variable x, defined over the set of non¬
negative integers, whose density function is p(x). Then the generating
function or e-transform of x, written <P(s) or z{p{x)} is defined as
SEC. 3-9 CONVOLUTIONS 119
for those $ for which the series converges. Because the p(x) must sum to
unity, we have <P(1) = 1. Consequently, the series (3-81) will always
converge for s in the interval — 1 < s < 1. Note that a knowledge of (P(s)
is as good as a knowledge of p(x) since if <P(s) is expanded in a power series
as in (3-81) the p(x) values can be found, i.e.,
p) = = E xl
M! e-p = y x!
rc = 0 *=0
(3-83)
n, p) = y ^ ^ j ^ pn(l — p)^1
1 - (1 - P)«
The same result is obtained for any positive n, even if it is not an integer,
if (3-73) is used. The proof is left for Problem 3-23.
3. The generating function for the geometric distribution will be denoted
by p). It is found by setting n = 1 in (3-84). Thus
P
@0(8} p) (3-85)
1 - (1 - p)s
The density function p(y) is called the convolution of the density functions
Pi(xi) and p2(x2).
Now denote by (Pi(s) and (P2(s) the generating functions for p1(x1) and
Pz(xz) respectively. Then by definition
” eo
+ Pi(2)p2(0)s2 + pi(2)p2(l)s3
Pi(S)p2(0)s3 + . . .
+ Pi(2)p2(l) + Pi(3)p2(0)]s3 + . . .
T P^k)v2(j - k)
L&=o
s3
density function, and if <P(s) is the generating function for this distribution,
then the generating function for y = xi ■+■ x2 is (P2(s).
Consider the random variable y = xx + x2 + . . . + x„ which is the sum
of n independent random variables xi} the generating function for the
probability density of x< being (Pt<s). To compute the density function for
y, we note that the density function for yi = xi + x2 is the convolution of
the density functions for xi and x2. The generating function for the density
function of yi is then (Pi(s)(P2(s). The density function for y2 = yi + x3
= Xi + x2 + x3 is the convolution of the density functions for y1 and x3.
The generating function of the density function for y2 is <Pi (s) <P2 (s) CP3 (s).
Continuing this process, we see that the density function for y = yn_2 + Xn
is the convolution of the density function for yn_2 and xn. The generating
function for the density function of y is then (Pi(s)(P2(s) . . . (P„($), i.e., the
product of the generating functions corresponding to each of the random
variables Xi. In the event that each x» has the same density function p(x),
then the generating function for the probability density of y is <P»(s) where
<P{s) is the generating function for p{x), and the probability density of y
is called the n-fold convolution of p(x). The density function which is the
n-fold convolution of p(x) we shall denote by p«(x). Note that
pm(x) = p{x)
and
X
Thus 2/ has a negative binomial distribution bN(y; n, P). We have used this
result previously without proof when developing the stuttering Poisson
distribution. We have now provided the proof.
In the above paragraphs, we have shown how to determine the gener¬
ating function of the probability for a finite sum of independent random
variables xi + being given and specified. Let us now study a
somewhat different case where n can also be a random variable. We shall
imagine that each random variable Xi has the same probability density
Pi(x) whose generating function will be denoted by <*(«). The * are as-
sumed to be independent random variables. The generating function of the
p obabihty density function for n, p2(n), will be written (P2(s). The proba-
-Jl ^isV ~ h V ~ Xl + ‘ ’ ‘ + Xn’ where ■?' is a specified non¬
ruLIrV integer
negative
^(s) = E P2(n)W(s)
n=0
(3-94)
However, by definition
00
<*(«) = E pz(n)s"
n=0
Thus
<?*(«) = <Pt[a»i(*)] (3-95)
t Writt6n aS a function of a function, i.e., (P*(s) is <p2 of
L r' Justobtamed Provides a way for deriving the generating
obtlmthf f! ,StUttepn? Poisson distribution, and hence another way to
obtain the stuttering Poisson itself. For the stuttering Poisson, p2{n) is a
SEC 3-9 CONVOLUTIONS 123
Thus tlu* generating function for the stuttering Poisson distribution, which
will be written d)» is
fm * p r« i s.y
««*) / L ( D1 J-‘f(x)dx (3-98)
J »» j« o LJ JJ*
The numhe*r p, is e*alle*el theg/th mome*uf of fix) about, the origin. The func¬
tion ‘J(.v), if if e*xists, is <*alle*el the* momtmf g(*ne*rat ing function of f(x) or x,
and it taims the* plaea*, as we* shall se*e*, e>f the* generating functie>n for discrete
variables.'1’ Whe*n x is de*line*d only for non-negative x1 then fF(.v) is also
calle*el the* Laplace transform oi fix). If might be* noted, although we shall
not. prove* it, that a, density function is imi((U<*ly de*tc*rmincd by its moment
generating func*tiom
Le*t Xi, x-i he* two e'ontinuous independent, ranelom variable's with densities
fi(A),fnU'j) and moment generating functions 9a(s) and fhi(,s) respectively.
We*, shall uenv compute the demsity function g(y) and moment generating
function <;(s) for the random variable* y xi L x«. The probability that
* If is also possible* to define moment, generating functions for discrete random var¬
iables, We* leave an analysis of t heses momemt generating functions for discrete random
variables to the* problems.
124
PROBABILITY THEORY AND STOCHASTIC PROCESSES CHAP. 3
Let us next compute the moment generating function of g(y) in the case
where x1; x2 and hence y must be non-negative. By (3-98)
f“ [v
S(«) h(y ~ Xi)e~ys dxi dy (3-101)
= Jo Jo hiXl)h
FIGURE 3-7.
SEC 3-9 CONVOLUTIONS 125
over Xi, infinitesimal strips of area parallel to the y axis such as the one
shown in Fig. 3-7. Thus (3-101) can be written
This proves that the moment generating function for g(y) is the product of
the moment generating functions for fi(xi) and fzixf). Precisely the same
result holds if xi, X2 can take on any real values. The proof is left for
Problem 3-24.
The above discussion shows that, given n independent random variables
Xi with moment generating functions ^-(s), the moment generating function
g($) of the random variable y = xi + ... + «» is
S(«) = SitoftC*) . . . SF«(«) (3-104)
In the special case where each Xi has the same density function f(x) and
moment generating function ^(s), then g (s) = 3rn(s). In this case g(y) is
called the n-fold convolution of f(x), and when the random variables must
be non-negative
(3-106)
Q~ (1/2) (mAO2 / g” (l/2tr2) [x2+2(or2a —ja)x]
v/2?r a
If Xi, x2 are two independent random variables which are normally dis¬
tributed with means /u, M2 and standard deviations 01, <r2 respectively, then
the moment generating function of y = xi -f- is
ea/2)(<n!j!-2ws)e(l/2)(«*s!-2^M) _ e(l/2) [(»i«+«‘)»»-2(w+/o)«]
(1 + l)u' e~ B+<*/£)]« du
r(a + 1)
sV (3-108)
1 +
0/
The above result can be used to provide another way to obtain the
probability density for the time required to incur Q demands when a
Jroisson process generates the demands. Recall that the time between
demands is exponentially distributed. The moment generating function
for^the exponential distribution is 1/[1 + (s/d)] as is seen by setting
“ ., J m • If foe Qth demand occurs at some time t after observation
on the system is begun, then t can be written t = 4 + 4 + . . . 4. tQ where
V if dfmaQ,d °CCUrs at the time ll>the sec°od demand occurs at a time t%
after the first demand, etc. Then each «,■ is exponentially distributed with
th ^ ,meaa- Furtlierinore, the U are independent random variables.
Thus the density function for t is the Q-fold convolution of the exponential
distribution, and its moment generating function is
= T(«; Q-1,0)
1 + (3-109)
X On = 1 (3-110)
l
The are called transition probabilities, and they are the conditional
probabilities of finding the system in state j at time 4+i if it was in state i at
time 4.
If the system is in state i at time 4, the probability that it is in state j
at time 4+2 is
n
a\f = aikakj (3-111)
k=i
since from 4 to 4+1 it can undergo a transition from i to any state h with
probability da (which may be 0) and from state Jc at time 4+1 to state j at
time 4+2 with probability a*y. The probability of transitions from i to h to j
is diicCLkj, and therefore the probability of moving from i to j from time 4 to
4+2 is found by summing over all h.
Let pr(i) be the probability that the system is in state i at time 4. Then
Pr+i(j) is given by
n
Pr+l(j) == ^ y Pr(i)dij (3-112)
i=l
It follows thus that if we are given the probabilities p0(i) that the system
is in state i at time to, and the transition probabilities a#, then it is possible
to compute the probabilities that the system is in state j at time
4, t = 1, 2,.... The type of process we have been describing is called a
128 PROBABILITY THEORY AND STOCHASTIC PROCESSES CHAP. 3
Markov process which is discrete in space and time (it is also referred to as
a Markov chain). It is said to be discrete in space and time because the
states of the system (referred to as the space variable) are discrete, i.e. can
be represented by integers, and because we are interested in the states of
the system only at discrete points in time. We cannot say anything about
the system at times other than to, h, k,. . .. The characteristic feature of a
Markov process is that the probability that the system is in a given state
at time tr+l depends only on the state that it was in at time tr, and the
transition probability, and is independent of the previous history of the
system (i.e., how it got to the state it is in at tr).
It might be expected that as time goes on, the probabilities pJj) will
become less and less influenced by the initial probabilities p0(j). We shah
say that the system is in a steady state if the probabilities pr(j) do not
change with r, i.e.,
Pr+l{JJ Pr\jj ior ail j
J2 p(f) 1
y=i
= (3-115)
With this additional constraint, the p(j) will be uniquely determined for
cases of interest to us (this is proved below).
it is a well known result from the theory of Markov processes that the p(j)
defined by (3-114), (3-115) exist, are unique, and each p(j) > 0. We shall
prove this without making use of the theory of Markov processes. How¬
ever, the proof does require some understanding of linear algebra, which
may not be familiar to the reader. An understanding of the proof is not
needed in later work. It is only necessary for the reader to recall that the
proof demonstrates the uniqueness of the p(j). To proceed with the proof,
let the matrix A = | |at-j| | be an nth order matrix containing the transition
probabilities, and p = [p(l),. . . , p(n)] be a row vector containing the
p(j). Then (3-114) can be written p(/„ - A) = 0 where /„ is the identity
matrix of order n. By (3-110), each row of A sums to unity and each row
of ln — A sums to zero. Thus the columns of — A are linearly dependent,
and \ln A[, the determinant of ln - A, vanishes. Thus there exist p(J),
not all zero, which satisfy p(Jn — A) = 0.
Let us now drop the nth equation in (3-114). The resulting set of n — 1
equations in n variables can be written p„_i(/»_! - An_a) = p(n)an, where
Pn-1 is a vector containing the first n — 1 components of p, /„_j is the iden¬
tity matrix of order n — 1, An_i is the submatrix formed from A by crossing
off the last row and column, and an = [a^, . . ., an,n-i]. Note that at
least one component of a, is positive, for otherwise it would be impossible
to ever leave state n. Thus at least one row of A„_i sums to a non-negative
number less than unity. From the theory of Leontief matrices, we know
that (In—i — An-i)-1 exists and can be written as a power series. Thus
For a given p(n), the vector p„_i is uniquely determined. If p(n) = 0 all
p(j) vanish and hence cannot sum to unity. Thus p(n) > 0. When
p(n) > 0, each component of p„_x will be positive, since for each i and j
there exists an N such that element aiP in AN is positive. Hence, on apply-
ing (3-115), we see that the p(j') exist, are uniquely determined, and each
p(j) is positive. This proves what we wanted to show.
In the previous section, we assumed that the state of the system was ob¬
served only at discrete points in time. We were interested in determining
the steady state probabilities that, at the time of observation, the system
was in one of a finite number of different discrete states. Now we wish to
study situations where the system is continually under observation and
the state of the system is known at each instant of time. Again we shall
assume that the states are discrete and can be represented by integers,-
130 PROBABILITY THEORY AND STOCHASTIC PROCESSES CHAP. 3
The aij(t) dt are called the infinitesimal transition probabilities, and the
aa(t) are called the transition densities. Furthermore, from (3-118)
t + dt) — p(i; Q
dt ~p(i; t) X a,ij(t) + X p(i; f)ay<(0
J =0 y=o
y
However, in the limit as approaches zero, the quantity on the left in the
above equation becomes dp(i; t) /dt, so that
The probability that the system leaves state i in time dt is p(i) a^dt,
and the probability that it changes from some other state to i is
dt. On equating these two terms and cancelling dt, we obtain
(3-124) which may be called the balance equations. The left-hand side of
(3-124) can be interpreted as the rate of change of p(i) arising from transi¬
tions out of state i, and the right-hand side the rate of change of p(i) from
transitions into p(i). As noted above, these two rates must be equal.
Markov processes continuous in time and discrete in space which are
concerned with systems that service in some way randomly generated
demands are often referred to as queuing problems. Hence, when the
theory is applied to inventory systems, we often say that queuing theory
is being applied to study the inventory system.
It is important to note that the theory developed above can easily be
generalized to systems where the description of each state requires the
specification of two or more non-negative integers rather than merely a
single integer. For example, suppose that two non-negative integers are
required to describe each state. Then let p(i, j; t) be the probability that
the system is in state (£, j) at time t, and (kj;mn{t) dt be the probability that
if the system is in state (i, j) at time t, it will be in state (m, n) at time
t + dt. With these definitions, (3-122) becomes
dp(i, i; i0
dt ~"P(h 3) 0 EE <'Hj;mn(t)
m=0n=0
ij 7*mn
= Xp[0|l, i, yU, 0]
Tf probability that no units are serviced in a time interval
l Sf 1Sn1’Slnce eyen/ a customer is in service, the probability that
' *° C°mplete Se"i0e “ «-"■ Consequently, we
3 =i+ 1
all other j > i
Similarly, if j < i
StioLTo stltef V t 01 ^ P1 that the aii are Zer° eXCept for tran-
i e ^ arrtfr ^ invo ve the <*™*ce of just a single Poisson event,
i ;a nT1Val °J tbe completion of service, so that transitions in the time
d can occur only to adjacent states. The states of the system and the
allowable transitions can be represented in a diagram such as Fig. 3-8.
FIGURE 3-8.
SEC 3-12 OTHER TYPES OF MARKOV PROCESSES 135
We can immediately write down the balance equations (3-124) for the
steady state probabilities p(j) in this special case; they are
There are two other types of Markov processes which are of interest. One
is that which is discrete in time and continuous in the space (state) variable,
and the other is that which is continuous in both space and time. We shall
discuss very briefly Markov processes which are discrete in time and con¬
tinuous in space, but shall do no more than merely mention something
about the nature of those continuous in space and time.
To introduce Markov processes which are discrete in time and continuous
in space, we imagine a system which is observed only at times U,
r = 0, 1, 2,..., where tr = 4-i + At, At > 0 being a finite constant, the
state of which can be described by a continuous non-negative variable x.
Let gT(x) dx be the probability that the state of the system lies between x
and x + dx at time tr, andf(x\y) dx be the probability that the state of the
system lies between x and x + dx at time tr+1 > tr given that the state of
the system was y at time tr. Then
If a steady state density g(x) exists such that g(x) dx is the steady state
probability that the system lies between x and x + dx, then g(x) is given by
rather than the cumulative function. This is done because the more com-
piete tables of the Poisson distribution [2,7] tabulate the complementary
cumulative function rather than the cumulative, and also because, if compu¬
tations are being made with a digital computer, it is often easier to evaluate
rather than the cumulative function. The purpose of the deriva¬
tions to follow is to express all relations involving the Poisson distribution
directly in terms of p(x; ft) and P(x; ft) so that numerical computations
can be made easily using the tables in [2,7], Even when the computations
are to be made on a digital computer, it will usually be desirable to make
hese transformations, since the resulting expressions will be easier to
handle on the computer than the expressions in their original form [which
may involve sums of the P(j; ft) or integrals over time of the p(j; n) or
P(j; wJ- The relations derived in the following paragraphs will hold for all
SEC. 3-13 PROPERTIES OF THE POISSON DISTRIBUTION 137
xp(x;*) = ^7 c—
(3-129)
= - l;*t), X = 1,2,
An especially easy way to reduce such sums to a simple form is through the
use of a recurrence relation. Note that
00 00 m—1 / i \
= T (* +
k = r— 1
/*) = /* E E (m r ) knp(k> /*)
k — r— 1 n=0 \ n /
171— 1 / -j \
which is A3-3.
In determining the expected number of backorders for some of the future
models to be discussed, it is necessary to evaluate numerically the ex¬
pression
E O' ~ f)p(i; aO
CO CO
Qm(r) = YJjmP(i)ti
y=r
These are also most easily reduced to computable form through the de¬
velopment of a recurrence relation. To obtain such a relation note that
00
Z [0- i)m+l-jm+1]P(j;»)
j-r
= Zz
y=ri = 0
(-l)m+1“{m t X) j4P(i; /i)
\ 1 /
(3-135)
(-i)—‘("+ 1) aw
which is A3-7.
In order to use (3-136) we must evaluate
O0(r)=ZPO» (3-137)
i—r
(r ~ l)2
P(r;/*)
This is A3-8.
Several integrals over time involving the Poisson distribution will also
be needed. The simplest of these is
Jo Jo rl A Jo rl
= 1i r" ^r+i
I XT PXT
(3-140)
A _ (r + 1)! 6 o +Jo (r + 1)
= -P(r+ 1;XT)
e~z dx
(n + r)!
Finally
f”P(r; \t) dt
fn+l ” f* fn+1
——- P(r; Xf) — X / —— 'p(r — 1; \t) dt (3-142)
n+ 1 v ’ o yow+1
The normal density function n(x; p, <r) is defined by (3-22). The proba¬
bility that the random variable x lies in (he interval ,r, < .r < j-» is given
by the integral of n(x; p, or) from xx to x«. In (3-22), p is t he expected value
of x and or2 is the variance of x. A normal distribution with p 0 and
v = 1 is called the standardized normal distribution, and its density func¬
tion given by (3-24) will always he denoted by <f>{w). The complementary
cumulative of <t>(w) is
n(x-, p, a) =
(3-144)
s fm /,. \
n(x; p, a) dx = J cj>(w) dw = <I> f ^^ J
* (^7^) (3-145)
where w = (z - p)/a.
It is often easier to work analytically with models of inventorv systems
if all variables can be treated as continuous. This allows one to eliminate
he problems caused by discreteness and to take derivatives instead of
dealing with differences, etc. In many cases the demand rate for an inven-
SEC. 3-14 THE NORMAL DISTRIBUTION 141
tory system will be sufficiently high that the problems caused by discrete¬
ness can be ignored and all variables can be treated as continuous. When
demand is treated as continuous, the most frequently used distribution to
describe the quantity demanded in a given time is the normal distribution.
There are several reasons why the normal distribution holds a central
place in working with continuous random variables representing the de¬
mands on the system. One is that the normal distribution is especially
easy to work with and is well tabulated. Another and more important
reason is that empirical studies have shown that, quite often, the normal
distribution seems to approximate very well the demand distributions
over the relevant time intervals which are encountered in practice. A final
theoretical reason for the importance of the normal distribution is that all
the distributions studied in this chapter, the binomial, negative binomial,
Poisson, gamma, and stuttering Poisson approach, as the mean of each of
the distributions approaches infinity, a normal distribution whose mean
and variance are the mean and variance of the appropriate distribution
under consideration. We shall not prove this theorem since in itself it is
not of great interest to us. The more important questions are how fast
each of the distributions approaches the normal distribution and how
much error can be involved if the normal distribution is used to approxi¬
mate one of these distributions with a given mean and variance. These
questions are in general quite difficult to answer (in fact they have not
been answered completely), and we shall not attempt to present any gen¬
eral discussion of them.
It is important, however, to recognize that for large means, the Poisson
distribution can be approximated by the normal, since tables of the Poisson
distribution are available only up to ft = 100. Por larger ft values, the
normal approximation must be used. It may be desirable to say a little
more about the normal approximation to the Poisson, because at first
glance it may not be clear what is meant, inasmuch as the Poisson distribu¬
tion is discrete and the normal distribution is continuous.
Recall that p(x; ft) is the Poisson probability of exactly x when ft is the
mean. For a given ft, a plot of p(x; ft) could be represented by a series of
vertical fines as shown in Fig. 3-9. Now it is possible to represent proba¬
bility as areas in Fig. 3-9 also if we draw in rectangles as shown. The length
of the base of each rectangle is unity and it extends from x — § to x +
The probability of x is then the area of the rectangle which includes x since
the length of the base of the rectangle is unity and its height is p(x; ft).
Thus we can write
00
FIGURE 3-9.
Note that p(x; p) is the area under the curve from x — § to x + and
P(x; jj) is the area under the curve from x — § to <*>.
The theorem stated above says that as fi —> co for the Poisson distribu-
tion, the broken curve outlining the area of Fig. 3*-9 approaches a normal
curve and
Note that the mean and variance of the normal distribution are the
as the mean and variance of p(x; p).
There is one other point related to approximating the Poisson distribu-
tion with the normal which deserves attention. Recall that P(0; p) = 1.
Hence if we are approximating P(x; p) by $(a: — p/^p), then we are saying
that to the accuracy required in the problem, we can assume that
i.e., the area under the normal curve to the left of x = 0 is negligible. This
will of course become closer and closer to being correct as p increases. This
point will arise a number of times in later work.
When approximating p(x; p) with the normal distribution for finite p,
increased accuracy can be obtained over using (3-146) if one notes that in
Fig. 3-9, p(x; p) is the area from » - } to * + | and P(x; p) is the area
from x — % to infinity. Thus for finite p, more suitable approximations are
1 /•*+( 1/2)
p(x; p)
Jx-(l/2)
(3-147)
= $
)
SEC. 3-15 PROPERTIES OF THE NORMAL DISTRIBUTION
x i M
(3-148)
^ .
-(l/2)(i-Xf)VXf
(3-149)
V2xX tl'2
y(t) =
II
1 /ct* *r*
r-
=jnr X (x - A01
X,M "L(D01/2JL (DO1/2 22)1/2^3/2J
i f* (x ~ AO (3-158)
'a; - XH
r, 2D1/2^2 0 .(DO1/2J ®
1 fvW , , 1 (T.
= -- / <j>(w) dw - — <t>[y(t)] dt
A Jv(.Ti) XK J jri <
We now observe that
Jo
««»)
i &-»•»■*] - (3.161)
we obtain
e-(2\x/D)
= ^(p)dp[|— __ -±:
D e-(2\*T/D)e- (2\Tll2p/Dlli)
2X __
so
I"
D r(2\x/D)$ |"__ J2 e- (2\2T/D) e-(2X!Ti/2p/-D1/2)^(p)
2X L(DT)“!] 2X /'
7(*- \T)/(DTy 2
(3-164)
FIGURE 3-10.
SEC. 3-15 PROPERTIES OF THE NORMAL DISTRIBUTION 147
/(i-XT)/(Z)T)Ui
where
Jo(x, Th TO =
After substitution of (3-168) into (3-167) and solving for Jt(x, Tx T2) we
obtain ’
T (T m rp \ 2(Dt)112 rx — xni-
Ux,Ti,T0-—
= —- * _ 2 fx + XTl
dx \(DT)u*vl(DTy/*] D liDT)1'*]
2(DT)U> fx - XT'
Wx(x, T)
X2
_ e2\x/D^ x + XT
,{DT) v* ]}
, pu*x , Fx - XT 1 . 2z fa; + X7H
+ x2rw2^ [(Dr)1'2] + x2® $Lcdt)i/2J
x3ri/2 e *L(-D7y/2J
or
+ 17!(^-25,i'’)*[^] (3-m)
+ e,,»(-L) (. _ 5) . [|d_-] - gs ^ [^]
which on using A4-18, is A4-10.
By use of the integration by parts trick, one can express Jn+1(2, Ti, T2)
in terms of Jn(x, Th T2) and JTly T2), thus obtaining a recurrence
relation for the J’s. Integration of Jn(x, Th T2) by parts with dv = tn~w dt
yields
which is A4-11.
REFERENCES 149
which is A4-14. Hence, once the J}s are known, the R’s can be evaluated
using (3-174). For example, from A4-9, A4-10 (which were derived above)
which is A4-16.
REFERENCES
PROBLEMS
3-1. Let a; be a discrete random variable which can only take on the n
values j + 1, j + n. If p(j + i) = l/n, i = 1,. . ., n, then x is
said to have a uniform or rectangular distribution. Similarly, if a: is a
PROBLEMS 151
0 , x < a
/(*) = * a<x<b,b>a
-0 , x > b
3-7. Compute the generating function for the binomial distribution, and
use the results of Problem 3-6 to compute the mean and variance of
the binomial distribution. Let xi,. . ., xm be independent random
variables having binomial distributions with the same value of p,
but perhaps different n values. What is the distribution of
y = xi + ... +XJ!
3-8. Consider the random variable y which is the sum of a random number
of independent, identically distributed, discrete random variables
Xi, i.e., y = xi + . . . + xn where n is a random variable and the xf
are independent and all have the same distribution. Let py, px, pn be
the expected values of y, xi} n respectively and of, <rl, al the corre¬
sponding variances. By use of the results of Problem 3-6 and the
generating function for y show that
3-9. Derive the results of Problem 3-8 directly without the use of gener¬
ating functions. Hint: If p(x) is the density function for the Xi and
r(n) the density function for n, then the probability of any set
(zi,. . ., xn, n) is
P(zi)pfe) . . . p(xn)r(ri)
ay = 23 X) (*i + • • . +
n XI, ,Xn
Xn)2p(xi). . . p(xn)r(n) — pi
given n
3-10. Note that the results of Problem 3-8 allow us to compute the mean
and variance of the distribution of demand in a time t when the time
between demands and the number of units demanded per demand
are random variables, provided that we know the means and var¬
iances of the distributions of the times between demands and the
number of units demanded per demand. By use of the results of
Problem 3-8 compute the mean and variance of the stuttering Poisson
distribution. Note also that the results of Problem 3-8 allow us to
PROBLEMS 153
3-13. Suppose that demands are generated by a Poisson process and that
the number of units demanded per demand is described by a binomial
distribution. Determine the density function for the number of
units demanded in a time t. Find the generating function for this
density, and the mean and variance of the density function.
3-14. Suppose that the procurement lead time is a continuous random
variable with mean \it and variance of. Suppose also that the process
generating demands is such that the mean time between demands is
and the variance of the number of demands occurring in time t is
cfdt. Finally imagine that the number of units demanded per demand
is a random variable with mean \xx and variance By use of the
results of Problems 3-8 and 3-12 determine the mean and variance
154 PROBABILITY THEORY AND STOCHASTIC PROCESSES CHAP. 3
of the lead time demand in terms of the other means and variances.
3-15. Derive the mean and variance of the normal and gamma distributions
using their moment generating functions.
3-16. Let x be a discrete random variable with density function p (x). Then
the moment generating function for x is defined to be
3K(«) = X) «""?(*)
x=0
Given 9TC(s), how can one determine the mean and variance of x?
Compute 971(5) for the binomial, Poisson, and negative binomial
distributions.
3-17. By use of the results of Problem 3-16 compute the mean and variance
of the binomial, Poisson, and negative binomial distributions.
3-18. Let xly x2 be two independent discrete random variables with moment
generating functions m(s) and 9Jl2(s) respectively. Determine the
moment generating function for y = xx + x2.
3-19. Consider a process generating demands for which the density func¬
tion for the time between arrivals is g(t), i.e., the probability that a
demand occurs between t and t + dt after the last demand is g(t) dt,
and this depends only on t. Suppose that we start observing a system
at a time r after the last demand and ask what is the probability
g(t\T) dt that the next demand will occur between t and t + dt after
we begin observation. Show that
g= = Jr
Evaluate for the case where a Poisson process is generating
demands. Now suppose that we begin observation of the system at a
random point in time. This means that when observation is begun
we have no knowledge of the time when the last demand occurred.
The probability that the last demand occurred between r and r + dr
in the past is equal to the joint probability that the next demand
after a given demand does not occur in a time r, i.e., (?(r), and that
we started observation between r and r + dr. This latter proba¬
bility is independent of r if we begin at random and is proportional
to dr, i.e., adr. Thus the desired probability is aG{r) dr where a is
determined so that the integral over r is equal to unity. By use of
these results, show that if u(t) dt is the probability that a demand
PROBLEMS 155
where X is the mean rate at which demands occur. Compute u(t) for
the Poisson distribution and show that g(t) = u(t).
3-20. By use of the results of Problem 3-19, show how to compute the
probabilities Un(t) that n demands occur in a time t when observa¬
tions are begun at random, given the probabilities Vn(t) that n units
are demanded in a time t when t is measured from a particular
demand.
3-21. Compute u(t) when an Erlang process of order m is generating de¬
mands. Attempt also to compute the Un(t) and Vn(f) for this process.
Note that u(t) is defined in Problem 3-19 and Un(t), Vn(t) in Problem
3-20.
3-22. Show that the equivalent of Eq. (3-29) holds for continuous random
variables.
3-23. Show that Eq. (3-84) is the generating function for the negative
binomial distribution even if n is not an integer.
3-24. Let xi, x2 be continuous, independent random variables which can
take on any real values. Show that the moment generating function
for y = xi + x2 is the product of the moment generating functions
for x\, x2.
3-25. Consider a Markov process discrete in space and continuous in time.
Prove that if Poisson processes generate transitions between states,
then the only an, i ^ j, which are different from zero are those which
correspond to the occurrence of a single Poisson event.
3-26. Provide a probabilistic interpretation of Eq. (3-140).
3-27. Compute the Poisson probability p(30; 25) = p(x; M) and the normal
approximation to it. Also compute P(30; 25) and the normal approx¬
imation to it. Next compute P(30; 25) — P(40; 25) and the
normal approximation to it. Compute p(95; 75), P(95; 75), and
P(95; 75) — P(105; 75) and the corresponding normal approxima¬
tions. In making the above computations, it is suggested that
Molina’s tables be used.
3-28. Differentiate Eq. (3-166) with respect to T and use this result to
demonstrate that Wa(x, T) is correct.
3-29. Determine the limiting forms of the Poisson properties derived in
Sec. 3-12 obtained by replacing the Poisson terms with their normal
156 PROBABILITY THEORY AND STOCHASTIC PROCESSES CHAP. 3
Order more.”
Anonymous
4-1 Introduction
ihat the optimal safety stock should be positive. The reason for this is that
if the safety stock was zero, then because of the random nature of the lead
time demand, the system would very frequently run out of stock before
the arrival of the order, thus incurring stockout costs. When it is expensive
to incur backorders or lose sales, then on the average it is cheaper to carry
some additional stock to avoid these stockout costs.
In this chapter, as in Chapter 2, we shall study lot size-reorder point
models for a single installation. Recall that a lot size-reorder point model
is one such that a quantity Q is ordered each time the appropriate inventory
level (the on hand inventory, the net inventory, the on hand plus on order
inventory, or the inventory position) reaches the reorder level. The purpose
jpf the analysis is to determine the optimal value of the order quantity Q
and the reorder point.
It is important to observe that by its definition, a lot size-reorder point
modeb assumes that an order is placed when the inventory level reaches the
reorder point, i.e., there is no overshoot of the reorder point.; In order for
this to be true the state of the system must be examined after every
demand. Thus the use of a lot size-reorder point model (a (Q, r) model)
implicitly requires that a transactions reporting system be usedl
The ability to place an order precisely as the reorder point is reached also
implies that, when the integrality of demand is taken into account, the
number of units demanded per demand cannot be a random variable.
When the number of units demanded per demand is a random variable,
then it is possible to overshoot (unavoidably) the reorder point. For such
situations, it may no longer be appropriate to order a fixed quantity each
time an order is placed. An alternative procedure is to set two levels
r, R(R > r) such that if the inventory level falls to x1 x < r on some
demand, we order up to the level R, i.e., a quantity R — x is ordered. We
shall refer to such an operating doctrine as an Rr doctrine, and a transac¬
tions reporting model which uses such a doctrine as an {R, r) model. The
task of working with (R, r) models when the number of units demanded per
demand is a random variable is much more difficult than for the models to
be discussed in this chapter. We shall discuss such models in Chapter 8.
It should be noted that a (Q, r) model is a special case of an (R, r) model,
with R = r + Q, which is appropriate when units are demanded one at a
time and there is never any overshoot of the reorder point.
In the real world, even if the number of units demanded per demand is
a random variable, then provided that the probability of a large overshoot
is very small, one will often operate the system using a (Q, r) model. To
d° this one might compute a Q* and r* using the model described in the
next section, and then either order Q* each time an order is placed, or
order up to Q* + r* each time the reorder level is crossed!
162 LOT SIZE—REORDER POINT MODELS CHAP. 4
Assumption (3) implies that at the time the reorder point is reached there
are no orders outstanding, so that the inventory position (the amount on
hand plus on order minus backorders) is equal to the net inventory (on
hand minus backorders). Thus the order point will be the same regardless
of whether it is based on the inventory position or net inventory. The
additional assumption will now be made that:
v (5) The reorder point r (based on the inventory position or net inven¬
tory) is positive.
This is almost always true in practice, since one will not normally wait
until there are backorders on the books before placing an order. ’ Because
of (5), there will be no backorders outstanding at the reorder point. In
placed received
FIGURE 4-1.
fact, at the reorder point, the inventory position is equal to the on hand
inventory. The behavior of the on hand inventory and the inventory
position for a system of the type being described is illustrated in Fig. 4-1.
For the model being examined, any one of the inventory levels, on hand,
net, or inventory position can be used to define the reorder point, and the
reorder point has the same value for any one of them. jNote that to use the
on hand level we must assume that after an order arrives, it is sufficient to
meet all backorders and raise the on hand inventory level above the reorder
point. If this ever failed to happen, the reorder point would never be
reached again and the system would proceed to accumulate backorders.
When the reorder point is thought of in terms of the inventory position of
the system, then assumption (3) guarantees that the on hand inventory
will always be raised above the reorder point when an order arrives, for
164 LOT SIZE—REORDER POINT MODELS CHAP. 4
n s c apter and the next, we shall employ the convention that the expected value
o any ran om variable computed on an annual basis will be referred to as the average
annua va ue. of the variable, while expected values for all other time intervals, or at
given pomts in time, will be referred to as expected values.
SEC 4-2 HEURISTIC TREATMENT OF THE BACKORDERS CASE 1 65
cost is high and will therefore make only a slight contribution"to the inte¬
gral. Thus it will be assumed that the expected number of backorders is
negligible, and for computing the inventory carrying charges, the expected
on hand inventory can be taken to be equal to the net inventory!
By definition, the expected net inventory at the time of arrival of a
procurement is the safety stock s. The expected net inventory immediately
after the arrival of a procurement is then Q + s. Thus, if the arrival of an
order initiates a cycle, the expected net inventory at the beginning of a cycle
is Q + s and is s at the end of a cycle. These will also be the expected
values of the on hand inventory at the corresponding times, when the ex¬
pected number of backorders can be neglected. In this case, since the mean
rate of demand is constant, the expected on hand inventory will decrease
linearly from Q + s at the beginning of the cycle to $ at the end of the cycle
and will average to
- (Q + s) + | s = ^ + s (4-f)
which is the average number of unit years of stock held per year (illustrate
this graphically).
In Chapter 2 we introduced the equivalent of s directly as one of the
variables. When dealing with probabilistic models, it is convenient to
eliminate the variable $ and replace it by r (Problem 2-15 showed that this
was easy to do in the deterministic case). Since the reorder point in terms
of the on hand or net inventory is r\ and since nothing arrives between the
time an order is placed on reaching r and the time this order arrives (i.e.,
assumption (3) holds), then if the order requires a time r to arrive and if x
units are demanded in this time,, the net inventory £(rr, r) will be £(x} r) =
r — x at the time of arrival of the order, and the expected value of the net
inventory averaged over all x for a given r is
xh(x) dx (4-6)
and where h(x) f(x, t) if the lead time is a constant r or h(x) is given by
(4-4) when the procurement lead time is a random variable with density
Trtr'rr\ averaSe annual inventory carrying costs are therefore
f k|_(Q/2) + r — ji].
It remains to evaluate the average annual cost of backorders. The
argument used to obtain this cost usually runs as follows: The average
number of backorders incurred per year is simply the expected number of
backorders incurred per cycle times the average number of cycles per
year, i.e., \/Q times the expected number of backorders incurred per cycle.
(This argument is correct, as we shall see later.) Now the number of back-
orders i?(z, r) incurred in a cycle will simply be the number of backorders
on the books when a procurement arrives. If the lead time demand is
X> ^he number of backorders will be
, v fo if x - r < 0
v(x, r) = (4-7)
[x - r if £ — r > 0
Thus the expected number of backorders per period 7j(r) is
where, as above, h(x) is the marginal distribution of lead time demand and
(4-8)
All the terms in the average annual variable cost X have now been found;
X = qA+IC
[1 + r ~ M] + ^ **(*) dx ~ rff(r) J (4-9)
SEC 4-2 HEURISTIC TREATMENT OF THE BACKORDERS CASE 167
since by assumptions (1) and (4) it is unnecessary to include the cost of the
units or the cost of operating the information processing system. We wish
to determine the values of Q and r which minimize the 3C of (4-9). If the
optimal values Q*, r* satisfy 0<Q*<oo?0<r*<oo? then Q* r* must
satisfy the equations
XA IC 7r\ , x
(4-10)
q _ + 7ri?(r)]
(4-12)
and
QIC
H(r) = (4-13)
7r\ \/
r00 /*«
= cr / v<j>(v) dv + j* <l>(v) dv (4-15)
J (r—fi)/cr J(r—pi)/<r
= ^(Lvif) + ^(Lr^)
Thus
The simple approximate treatment of the lost sales case differs very little
from the backorders case studied in the previous section. From our study
of the lost sales case in Sec. 2-6,| it is clear that the minimization of the
average annual cost is equivalent to the maximization of the average annual
profit if in the cost expression the cost of a lost sale includes the lost profit.)
It is also clear that, in general, the average number of cycles per year is no
longer X/Q but is instead X/(Q + XT), where f is the average length of timo
per cycle for which the system is out of stock. In the real world T is usually
a very small fraction of the total length of the cycle. Since it is very incon¬
venient to include f in the analysis, the following assumption is usually
made in the simple treatments.
(a) The value of T is small enough to be neglected, so that the average
number of cycles per year is X/Q.
We shall suppose that assumptions (1), (3) and (4) of Sec. 4-2 also apply
here. Assumption (2) of Sec. 4-2 is not needed here, since for the lost sales
case the cost of a lost sale will always be x; there is never a term propor¬
tional to time, as it does not make any sense to talk about the time for
which a lost sale exists. We do assume, however, that x includes the lost
profit.
Xbe only difference between the lost sales and backorders models
comes in evaluating the safety stock expression. The expected on hand'
inventory when a procurement arrives will be s, the safety stock, and the
expected on hand inventory immediately after a procurement arrives will
be Q -fi- s. Thus the expected on hand inventory varies between Q + s
and s in a cycle and averages to (Q/2) + s. Let e(x, r) be the on Wnd
inventory when the procurement arrives if the lead time Hp™^ is x
Then
r — x, r — x > 0
t{x, r) =
[0, r - x < 0
and the expected amount on hand when a procurement arrives is
S = Jo t('X’ ^ = Jo ^ ~ ^ (4-17)
(4-18)
= r — M + j xh(x) dx — rH(r)
SEC 4-4 DISCUSSION OF THE SIMPLE MODELS 169
It will be noted that the expected number of lost sales per cycle is pre¬
cisely the same as the expected number of backorders per cycle in Sec. 4-2,
i.e., is given by (4-8). This follows, since if the lead time demand is x, the
number of lost sales will berr-rif# — r > 0 and 0 otherwise. Averaging
this over x, we obtain (4-8). It follows that the average annual variable
cost in the lost sales case is
3C = | A + IC | + r - m
In this section we shall discuss some interesting properties of the two models
developed in the previous two sections, present a numerical technique for
determining the optimal Q and r values, and give a numerical example to
illustrate the use of the models.
First note from (4-12), (4-21) and (4-8) that in both the backorders and
170 LOT SIZE—REORDER POINT MODELS CHAP. 4
lost sales cases Q* > Qw, i.e.; the optimal Q value is never less than the
Wilson Q. In fact, for all normal cases, Q* > Qw for any r (not simply the
optimal r). The intuitive reason for this is that the expected number of
backorders or lost sales per cycle depends on r but is independent of Q.
oweyer, the average annual cost of backorders or lost sales is proportional
to 1 /Q. Therefore, for a fixed, specified value of r, when w > 0, it always
pays to increase Q somewhat thus incurring more carrying charges per
year m order to gain the benefits of correspondingly reducing the average
i
annual cost of backorders or lost sales. The value of Q should not be in¬
creased indefinitely because after a point the incremental savings in
expected stockout costs becomes less than the incremental increases in
expected carrying costs.
Secondly it will be observed that (4-13) does not make any sense if
; 7" > L ThlS WlU never occur for cases for which this model is in-
en e to apply, i.e., to high backorder costs. This anomaly arises because
the backorders term was omitted in determining the inventory carrying i
costs Note that a similar problem does not arise in the corresponding
equation (4-22) for the lost sales case.
Before going on to a third observation concerning the models, it will be
convenient to introduce a numerical procedure for solving the pairs of
equations (4-12), (4-13) or (4-21), (4-22). This procedure is as follows:
fltial estlmate of & i-e-> write Qx = Qw. Then use Q, in
j or v*“22) to compute n. The n so obtained is used in (4-12) or (4-217
o compute Q2. This Q2 is used in (4-13) or (4-22) to compute r2, etc. This
iterative procedure is continued until Q and r are obtained with sufficient
accuracy. If the equations have a solution at all, then the above iterative
scheme must converge to a minimum cost solution^
A proof of the convergence of the iterative scheme will be carried out
using a graphical argument for the backorders case. The corresponding
argument for the lost sales case can easily be made and is left for Problem
f'4^EnUat,10nS t12 and (4'13) can be thought of as describing two curves
whe^Or"Pnne' F°r th! CTe described by equation (4-13), we note that
££ Lkiis"a,ld when 8 ■A/IC-m -1—- «•
dQ t\ dr = IC 1
~dr ~ -fch(r)
or (4-24)
dQ 7rX h(r)
For the curve described by (4-12) note that when r = 0
'2\[A + Tji]
Q =
IC
and when r = =c, Q = Qw. Furthermore dr/dQ < 0 If
one plots the two *
curves something like Fig. 4-2 will be obtained.
.*
SEC 4-4 DISCUSSION OF THE SIMPLE MODELS 171
puted from the lost sales model and the backorders model when the same
values of the parameters, i.e., the same values of X, A, I, C, and x are used
in both, it will be found Q* for the backorders case is larger than Q* for the
lost sales case while r* for the backorders case is smaller than r* for the lost
sales case. This can be seen immediately if one follows through the iterative
computational procedure. At the first step with Qx = Qw> a larger n is
obtained in the lost sales case than the backorders case. This in turn gives
a smaller Q2 in the lost sales case than the backorders case. Consequently,
the ratio on the right of (4-22) will again be smaller than that of (4-13)
and r2 for the lost sales case will be larger than r2 for the backorders case.
Continuing in this way we see that the result stated above holds. In prac¬
tice the difference between the Q*, r* values for the lost sales and back¬
orders cases will generally be very small.
The intuitive explanation of the above behavior is roughly as follows:
Note that if the same values of Q and r are used in both models, the average
annual inventory carrying cost will be less for the backorders model than
for the lost sales case (since in the backorders model for those cycles when
backorders occur, the initial inventory in the next cycle is reduced by the
backorders to being less than Q, while it is never less than Q in the lost sales
case). Consequently, other things being equal, it pays to have a larger Q
in the backorders case than in the lost sales case. On the other hand, an
increase in r causes less of an increase in average annual inventory carrying
charges for the lost sales case than for the backorders case, while yielding
the same change in the expected stockout cost in both cases (provided the
Q’s are the same). Hence, other things being equal, it will be advantageous
to have a higher r in the lost sales case than in the backorders case.
Sometimes it is of interest to compare the average annual cost of opera¬
tion obtained from one of the two models discussed above with that for the
corresponding deterministic model using the same values of the parameters.
The average annual cost for the deterministic model will never be greater
than the average annual cost for the model which treats demand as a
stochastic variable. The difference in average annual costs of the prob¬
abilistic model and the deterministic model will be called the average
annual cost of uncertainty. This indicates what annual savings one could
obtain, on the average, if all uncertainties in the demand and lead time
variables could be eliminated, the mean values remaining unchanged.
We now give an example illustrating the use of the simplified models.
e, = e. - - V2(1Jf(5Cj = -1130
Then rx is computed from (4-22). In this case )
since the lead time demand is normally distributed with mean 750 and
standard deviation 50. Thus
n - 750
= 2.695
n — (i
iKn) = (m - ri)4>
174 LOT SIZE—REORDER POINT MODELS CHAP. 4
Q2 = , ^/2(1600)|"4000 + 20001.054361]
Then for n
* O
\
~ 750 \
50 )
L147
321.1 “
,
and
r2 - 750 0 fid
50 “ 2'69
or
ri = 750 + 134.5 = 884.5
There has been essentially no change in the safety stock. Additional itera¬
tions are not needed since the changes will be negligible. The optimal
values are Q* = 1147, r* = 884, s* « 134. Because of the high cost of
s ockouts a safety stock of 134 tubes is carried. The expected time between
the placing of orders is 8.60 months.
The average annual cost of ordering, carrying inventory, and stockouts
is easily computed from (4-23). It is
Tin , 2000(1600)1, .
+ L10 + 1147 J[- 134(0.00357) + 50(0.01071)]
since it is never optnnal to have lost sales in the certainty case Thus the
average annual cost of uncertainty is
which is $1/0 more than the average annual cost of $1340 incurred to carry
the 134 units of safety stock Of thp in 70 <kiaa .* ^
of Stock™,wh,l0 +n • „ *1/0’ $160 ls the average annual cost
of stockouts while the remaining $10 is incurred because Q* differs from
6 a W en usmg the °Ptimal Q and r values, the average annual
SEC 4-5 EXACT DEVELOPMENT—INTRODUCTION 175
stockout cost is only $160. This means that on the average a lost sale will
be incurred only once every 12.5 years.
Often it is desirable to know how sensitive Q* and r* are to changes in
the parameters such as A, /, or x. In Fig. 4-3, it is shown how Q* and r*
vary with x. It will be observed that Q* is very insensitive to x while r*
is slightly more sensitive but not really too sensitive, since a change in x
from $500 to $5000 increases r* by only about 40 units, i.e., increases the
safety stock by 40 units. Since the mean lead time demand is 750 units,
the safety stock changes from about 109 units to 149 units or a change of
less than 50 percent for a factor of 10 change in the cost of a lost sale. This
FIGURE 4-3.
points out that even if one cannot measure x too accurately, then provided
it is of the right order of magnitude, the Q and r values computed will not
be very far from the optimal ones that would be obtained on using the
correct value of x.
it yields directly the state probabilities, which are not obtained directly
from the other method of generating the average annual cost.
Before going on to develop the exact annual cost expression for the
backorders case with constant lead times, we shall first compute some time
averages that will be needed, and also say a word about the possible inter¬
pretations of the state probabilities and their use in computing the time
averages.
* The limits of random variables studied in this section cannot rigorously be considered
to be limits in the strict mathematical sense. The reason for this is that one cannot
guarantee that there exists a T0(e) so large that for all T > T0(e), the difference between
the random variable and its limiting value is less in absolute value than e, for any e > 0,
simply because of the stochastic nature of the process. The sort of statement which can
be made is that the random variable approaches a given limit with probability one, i.e.,
the probability that the random variable differs in absolute value from its limiting value
by any specified positive quantity approaches zero as T —» ». The limits employed in
this section will be interpreted in this way.
LOT SIZE—REORDER POINT MODELS CHAP. 4
lim? = * (4-25)
f—no f Q f-no f Q
iS f~K 0-26)
Hence we have shown that in the backorders case, the avm^e-numher^
qrders per year, i.e., the average number of cycles per year is.X/Q,. This is
independent of the specific nature of the demand or lead time distributions
provided that an order is placed after every Q demands.
The fraction of the time that the system is out of stock which will be
written Pout, is
n
Pout = (4-27)
This then is, by definition, the probability that the system is out of stock if
we observe the system at a random point in time. Also
n 71
Do* 2>
B = lim lim (4-28)
f—► OO
are the average unit years of stock held per year, the average unit years of
shortage incurred per year, and the average number of backorders incurred
per year. If the on hand inventory were always a constant D, then the unit
years of stock held per year would be D. Thus, Hk.the expected value of
th£xm hand inventory at any instant of time. Similarly, B is the wtperrtpH
number of backorders on the books at any point of time.
It is desirable to rewrite the expression for E. Note that since the limit
of a product is the product of the limits
2 T" 2 «<
E = lim ~ = lim J —-*=1—
f ^ L
r £T
i=i
*
n n (4-29)
I>" Dir
Dr"
SEC 4-6 COMPUTATION OF TIME AVERAGES 179
The second limit is X, since £?« 1 £* is the total number of units demanded in
the time i T". The ratio, in the limit as f —> °°, must be X by the defini¬
tion of the mean demand rate. We have just shown that the average
number of backorders incurred per year is the mean rate of demand times
the probability that the system has no stock on hand.
Another interesting result can be obtained using the fact that the limit
of a product is the product of the limits. We note that
where by definition
Q = lim (4-31)
is the average number of unit years of stock held per cycle. Furthermore,
by definition of X, limf_>» nQ/f = X. Thus (4-30) follows. The above
shows that the average number of unit years of stock held per year is \/Q
times the average number of unit years held per cycle.
The values of D, B, Ey and P0ut will depend on the nature of the sto¬
chastic processes generating demands and lead times. They can be com¬
puted by first computing the corresponding expected value per cycle and
then multiplying by \/Q. The expected values per cycle are computed by
using the information given about the demand and lead time distributions.
However, if we can determine the steady state probabilities $(x) that the
net inventory has the value x, then the values of D, B, E, and PQut may be
computed directly. Thus
Imax ~1 0
It is important to note that the steady state probability ^(x) can be given
several interpretations. It is the probability that the inventory position
is x if we observe the system at a random point in time (after the system is
in a steady state mode of operation). It is also the long run fraction of the
time that the net inventory will have the value x. Finally, it is the limit
as N —»co of the fraction of an ensemble of N systems in statistical equi¬
librium that will have the inventory position x at any given point in time.
The analysis for the lost sales case is precisely the same as for the back-"
orders case examined above. If the same definitions are used as in the
backorders case, it is immediately clear that the average number of lost
180 LOT SIZE—REORDER POINT MODELS CHAP. 4
sales per year is \Pout, and D, the average unit years of storage per year, is
given by (4-28) and (4-32). The average number of cycles per year is
different in the lost sales case than in the backorders case, however. We
shall now compute the average number of cycles per year in the lost sales
case. For this case, Q units are demanded in Tl, not in time T{ as in the
backorders case. The average number of cycles is, by definition
However, nQ is the demand in time £?= i Tl and hence the ratio of these two
quantities must approach X. Next observe that
(4-34)
i=l t=l i=l
and hence
I
lim = 1 - lim 1 - Po (4-35)
T—►« S f—f
Thus
and the average number of cycles per year is (\/Q)(l - Pout). It is also
true that
where T is the average length of time per cycle for which the system is out
of stock.
SEC. 4-7 EXACT FORMULAS 181
4-7 Exact Formulas for the Backorders Case with Poisson Demands
and Constant Procurement Lead Time
We shall now determine an exact expression for the average annual cost
in the case where a Poisson process generates the times between demands,
units are demanded one at a time, the mean rate of demand being X units
per year, and the procurement lead time is a constant r. In addition to
treating the demand variable as being discrete, the order quantity Q, the
reorder point r, and all the inventory levels will also be treated as discrete
variables.
As yet,- we have not discussed which inventory level (or levels) can be
used to define the reorder point. We can see at once that the on hand
inventory (or net inventory) cannot be used to rigorously define the reorder
point, since, if there was a very heavy demand during some cycle and a
huge number of backorders were incurred, then the arrival of whatever
outstanding orders there were might never bring the on hand inventory
back up to the reorder point again, and hence another order would never be
placed. This leaves only the inventory position (the amount on hand plus
on order minus backorders) as a suitable level for defining the reorder point.
The inventory position does provide a suitable level, since the difficulty
referred to above with the on hand level cannot occur. If during some cycle
there is very heavy demand and a considerable number of backorders are
incurred, it will also be true that a correspondingly large number of orders
will be placed, for the reorder point in terms of the inventory position will
be crossed a number of times.
If r is the reorder point in terms of the inventory position, then imme¬
diately after an order is placed the inventory position is Q + r. Thus the
inventory position must have one of the values r + 1,. . . , r + Q. It is
never in a state r for a finite length of time, since as soon as a demand occurs
which reduces the inventory position to state r an order is placed bringing
the state to r + Q. Note that specification of the inventory position tells
us nothing about the on hand inventory or net inventory. If the inventory
position is r + j, there may be no orders outstanding with the net inventory
being r + j, one order outstanding with the net inventory being r + j - Q,
etc. For Poisson demands, where there is a positive probability for an
arbitrarily large quantity being demanded in any time interval, it is the¬
oretically possible to have any number of orders outstanding at a particular
instant of time. (Of course the probability of a large number of orders
outstanding will often be very small.) In a case where there was a finite
upper limit to the amount that could be demanded in any time period,
then there could never be more orders outstanding than total to an amount
equal to the maximum lead time demand. On the other hand, specification
of the net inventory does uniquely determine the number of orders out-
182 LOT SIZE—REORDER POINT MODELS CHAP. 4
standing and the inventory position, since if x is the net inventory and n
orders are outstanding, the inventory position is x + nQ, and this can be
written r + j,j — . fQ. This restriction determines both n and
j. The value of n is the largest non-negative integer which makes
r<x + nQ<r + Q.
In the simple approximate model discussed in Sec. 4-2, we assumed that
there was never more than a single order outstanding, and that the reorder
point could be based on the on hand inventory. We have just seen that, in
general such restrictions cannot be made rigorously. It may be justifiable
to stipulate, however, that the probability that more than one order is
outstanding is very small.
In order to compute P0ut, the expected number of backorders on the
books at any point in time, and the expected on hand inventory at any
point in time we need the state probabilities for the net inventory. The
straightforward way to compute these would be to attempt to write down
the difference equations (3-124) which describe transitions between the net
inventory states. Difficulties are encountered with this procedure, however,
since the lead times are constant and are not generated by a Poisson proc¬
ess. We shall therefore use a different approach, and instead we begin by
hnding the state probabilities p(r + j) that the inventory position of the
system is r+j,j = 1,. . ., Q. A knowledge of these permits the compu¬
tation of the state probabilities for the net inventory. The advantage of
doing this is that the nature of the procurement, lead times dot's not, enter
into the computation of the p(r + j).
In a time dt, the system inventory position moves from state r j■ j to
r+J ~ 1,3 > 2, if a demand occurs. This happens when a Poisson process
generates demands and units are demanded one at a time with probability
X di. If the system is in state r + 1 and a demand occurs, it, names to state
1 e SmCe the demand triggers the Pigment of an order. Thus we have
the diagram representing the transitions shown in Fig. 4-4
The balance equations (3-124) then read for this particular case
SEC 4-7 EXACT FORMULAS 183
and since £?-i p(? + j) = 1, it follows that the unique solution for p(r + j)
is
We have shown that, in the steady state, each of the inventory position
states r + j has the same probability 1/Q which is independent of j. We
say that the states are uniformly distributed and that p(r + j) is a uniform
distribution.
To compute the state probabilities for the on hand inventory and the
number of backorders, consider the system at any instant of time t Also
consider the time t — r, r being the procurement lead time. Note that
everything on order at time t — r will have arrived in the system by time t
and nothing not on order at time t — r can have arrived in the system by
time t Thus, if the inventory position of the system was r + j at time
t — r, the probability that there are x units on hand at time t is the
probability that r + j — x units were demanded in the lead time r if
r + j — x > 0 and is zero otherwise. For Poisson demands, the probability
of having this number of units demanded is p(r + j — x; Xr) when
r + j — x > 0. However, the probability that the inventory position is
r + j at time t — r is 1/Q. Then if \f/i(x) is the state probability that x
units are on hand at any time t, to find \pi(x) we multiply p(r + j — x; Xr)
by 1/Q and sum over the j for which r + j — x > 0. Thus
1 Q
fifr) = q P(r + j - x; \t)
V j =1
1 r+Q^x
= q 22
u = r-\- 1—x
Xt)
(4-40)
= k Z)
^ Lu = r+1—x
Xr) ~ Z
u=r+Q+l—x
pfaXr)
q [P(r + 1 — X) Xr) — P(r + Q + 1 — x; Xr)], 0 < x < r+ 1
1 A \ r+Q—x
*&) = q Z p(r + 3 - Xt) = q Z
ti = 0
p(u> Xt)
(4-41)
= ^ [1 - P(r + Q + 1 - *; Xr)], r+l<x<r+Q
184 LOT SIZE—REORDER POINT MODELS CHAP. 4
. . l -A i »+r+Q
&(y) = q 22 p(y + r + j- Xt) = - ^ p(U; Xt)
y= l ^ u = i/-f-r+l
l (4-42)
= Q lp(y + r + 1; Xt) - P{y + r + Q + 1; Xt)], y >0
°° ^ i r*00 co
Pont = Z ^ = |_Z P(y + r+ 1; Xt) -J^P(y + r +
Q Q + l; Xt) J
If-" 00 00 (4-43)
=o Z L«=r+1
J2
t* = r+Q+l
^(«;Xt)1
J
and from A3-6
a(p) = W=»+l
Z P(u’ Xt) = XtF(v; Xr) - vP(v + 1; Xt) (4-45)
Thus E(Q, r), the average number of backorders incurred per year is
by (4-29) ’
B(Q> r) = J2 yi2(y)
y= 0
= Q^ Z0 yip(y + r + 1; Xt)
y=
- P(y + r + Q + 1; Xt)]
j r co ^
(4-47)
SEC 4-7 EXACT FORMULAS 185
»(p + 1)
P(» + 1;Xt) (4-49)
(» +1 E )
li —»+l
p(«; M = (Xt)(» + 1)P(»; Xt)
H-Q
+ E *[1 - P(r + Q + 1 - Xt)] l (4-53)
x —r+1 J
1 r r+Q r
E xP(r + Q + l — z; Xt)
186 LOT SIZE—REORDER POINT MODELS CHAP. 4
Now
1 r+Q
r* rv ! Q „ _
+ r («4>
Also
v v
«> w
m, r) = ^ + r
If" 00
oo
However,
oo ^
when ^ = Xr is the mean lead time demand. The last step follows from
A3-6. Then from (4-47), we see that
There is another simple procedure that one can use to compute the
expected value of the on hand inventory. By comparison of the result so
obtained with (4-57), an interesting result can be obtained. Note that the
expected value of the inventory position is the expected value of the on
and inventory plus the expected amount on order minus the expected
number of backorders, since the expected value of the sum of several ran¬
dom variables is the sum of the expected values. Thus the expected on hand
SEC 4-7 EXACT FORMULAS 187
and so
i.e., the expected amount on order is equal to the mean lead time demand.
This result, which we have just proved rigorously for the case under con¬
sideration, holds under much more general assumptions. The fact that the
expected amount on order should be the mean lead time demand can be
seen intuitively as follows. Imagine that orders flow into one end of a
pipeline and that procurements flow out the other end. Since all demands
are ultimately met, the mean rate of flow of units ordered into the pipeline
must be X. Since an order remains in the pipeline for a time r, the expected
number in the pipeline should be Xr = jjl.
All the terms in the average annual variable cost 3C have been evaluated;
3C consists of the expected ordering costs, holding costs, and backorders
costs. It is
(4-61)
= |4 + Jc[| + | + r-M] + tE(Q, r) + (* + IQBiQ, r)
where E(Q} r) and B(Q, r) are defined by (4-46) and (4-52) respectively.
We are here assuming that the unit cost C of the item is constant and is
independent of Q, so that the cost of the units themselves need not be in¬
cluded. We are also assuming that the cost of operating the information
processing system is independent of Q and r and hence need not be included
in 3C.
We have in the above derivations carried out the operations acting as if
r was positive. In actuality, with the proper interpretations, the above
formulas also hold when r is negative or zero. The proper interpretation
is to take P(v; Xr) = 1 if v < 0. We ask the reader to prove these state¬
ments in Problem 4-56. For example, if v is negative
188 LOT SIZE—REORDER POINT MODELS CHAP. 4
In problems of practical interest, it is usually true that the terms a(r + Q),
P(r + Q) in E(Q, r) and B(Q, r) respectively, are negligible. These terms
are important only if there is a significant probability that the lead time
demand will be greater than r + Q. If this happens, then there wfil stiff
be backorders on the books after the arrival of a procurement, i.e., the
procurement will not be sufficient to remove all the backorders. It will
never be optimal to have a sizeable probability that so many backorders
will be incurred that the arrival of an order will not be sufficient to meet all
of them unless it costs very little to incur backorders. This is seldom the
case for most real world problems, and hence in practice it is usually a very
good approximation to neglect a(r + Q) and /3(r + Q). When air + Q),
SEC 4-8 AN IMPORTANT SPECIAL CASE 189
Thus Q* is the largest Q for which AqX(Q, r*) < 0, or Q* = 1, and r* is the
largest r for which ArX(Q*, r) < 0.
When X is given by (4-64)
Thus
I2\A ,-
Qw —
-Jq = V1000 « 31.6
SEC. 4-9 THE NORMAL APPROXIMATION 191
b - liwif]<si ■ ^w-013475'+(aoo5383)
(31-6) (10)
= 0.152 > 0.126
50(50)
For large lead time demands the Poisson probabilities appearing in the
model developed in the last sections can be conveniently approximated by
a normal distribution. In such cases, it is usually a good approximation to
treat Q and r as continuous variables also. Here we shall develop the
equations corresponding to those in the previous section for the case where
the lead time demand can be considered to be normally distributed. Recall
from Sec. 3-14 that as v(xiu) approaches a normal distribution
with mean p and with variance a2 = p. In the development here, we shall
not specifically set cr2 = p in the normal distribution. The reason for this
is that later we shall want to consider the case of variable lead times. It
will be found that it is often desirable to represent the marginal distribution
of lead time demand by a normal distribution. However, in such cases, the
variance a2 will be greater than p because of the influence of variable lead
times. In this section, we shall imagine the lead times to be constant.
However, by not specifically setting a2 = p, and instead leaving <r as an
independent parameter, we shall simultaneously obtain the appropriate
192 LOT SIZE—REORDER POINT MODELS CHAP. 4
equations that will be useful when considering the lead times to be random
variables also.
Since r and Q are now being treated as continuous the inventory position
can take on any value between r and r + Q rather than only the values
r + 1,. . . , r + Q. The probability that the inventory position lies be¬
tween x and x + dx, r < x < r + Q is simply dx/Q, because in Sec. 4-7
we showed that p(r + j), the probability that the system was in the state
r+j was l/Q. This follows by using the same trick of drawing in rec¬
tangles used in Sec. 3-14 to show how the Poisson distribution approached
the normal. Note that if the same procedure was used as in Sec. 3-14 the
inventory position would range from r + §tor + Q + i This gives a
slightly better approximation than using the range r to r + Q, but since
for practical purposes the difference is negligible, and since in the physical
system the inventory position cannot get above r + Q, we prefer to use the
range r to r + Q. Note that
dx _
Q ~ (4-74)
as desired. In place of the Poisson density p(x; M) for the probability that
x units are demanded in a lead time, we shall now use
(4-75)
as the probability density for lead time demand, where p. is the expected
lead time demand. In the following, we shall make use of the fact that when
it is valid to use the normal approximation, then one can assume that
$(-M/<r) = 1 (see Sec. 3-14).
Then, if &(a0 is the probability density for the quantity on hand at any
time t, it follows that
Similarly, the density function My) for the number of backorders on hand
at any tune t is
SEC 4-9 THE NORMAL APPROXIMATION 193
Then
Now by A4-6
-r*(-t<ity~,i)»] <«•)
Therefore
where
Finally
fr+Q Q
D(Q, r) = Jo a^i(x) dx = | + r- iL1 + £(Q, r) (4-86)
All the terms needed in the average annual cost X have been evaluated;
X is
dX dX
dQ~ dr (4-88)
Problem 4-16 asks the reader to obtain all the first and second derivatives
of X.
In the usual case where the a(r + Q) and /9(r + Q) terms are negligible
and can be ignored, the equations of (4-88) reduce to
The same sort of iterative scheme used to determine Q* and r* for the
simple model of Sec. 4-2 can also be applied here. That is, we first set
Q = in (4-90) and determine rh the value of r which satisfies this
equation. Then rx is used in (4-89) to yield Qi, etc. Now, of course, it will
be somewhat harder to solve (4-90) than it was to solve the corresponding
equation for the simple case.
At this point we might note that when x = 0, a(r + Q) is negligible, and
S
the contribution of B(Q, r) to the inventory carrying charge is negligible,
(4-87), (4-89), and (4-90) reduce to (4-16), (4-12), (4-13) respectively. This
s ows us that the model of Sec. 4-2 is applicable, when the assumptions
stated above apply, for any number of orders outstanding, and is not
restricted merely to the case of never more than a single order outstanding.
It is not simple to make numerical computations using the exact form
for X obtained in this section. When the a(r + Q) and /3(r + Q) terms are
included, 3C is not generally convex. This nonconvexity makes it difficult
SEC. 4-10 AN EXAMPLE USING EXACT & 195
to rule out the existence of local minima which differ from the absolute
minimum. The nonconvexity has also caused convergence difficulties with
Newton’s method and the method of steepest descents in examples which
the authors have attempted to work out in which the a(r + Q) and
0(r + Q) terms were important. When the a(r + Q) and p(r + Q) terms
must be included, it would seem necessary to use a computer and an
appropriate search routine to determine Q* and r*. It is fortunate that in
practice it is seldom necessary to include these terms.
Consider an item for which the demand over any time interval can be
considered to be Poisson distributed with X = 400 units per year. The
procurement lead time is a constant, and r = 0.25 yr. The values of the
other relevant parameters are A = $0.16, C = $10.00, I = 0.20, x =
$0.10, X = $0.30 per unit year. It is desired to determine Q* and r* when
a Qr operating doctrine is used to control the system, and Q*, r* are
determined by minimizing the exact expression for X given by (4-61).
This problem was solved on the Burroughs 220 computer using the code
referred to in Sec. 4-7. The results obtained were
As was noted in Sec. 4-7, the code determines a relative minimum, but does
not guarantee that it is the absolute minimum, if there exists more than one
relative minimum. The authors could find no indication that there exist
other relative minima, and hence it seems very likely that the above values
are indeed the optimal values of Q and r. One finds that in X*, the various
cost components have the following values
^ A = 3.368 (4-92)
Observe that the backorder costs account for slightly over 40 percent of
the average annual cost. To show that the above answer does yield a
relative minimum, we present the values of X for neighboring (Q} r) points
in the following table
19 6 LOT SIZE—REORDER POINT MODELS CHAP. 4
TABLE 4-1
Values of X (Q, r)
Q 18 19 20
r
(4.96)
For this example, the a(r + Q), p(r + Q) terms are not completely negli¬
gible, but are not really large as compared to the a(r), fi(r) terms, when
evaluated at Q* and r*9 i.e.,
Even though the values of tt and f are absurdly small, the values of the
a(r "t" Q)j “b Q) terms evaluated at Q*, r* are not very large compared
to the a(r), p(r) terms. This suggests that the approximate model studied
in Sec. 4-8 which neglects the a(r -(- Q), fi(r + Q) terms should converge
and yield Q and r values close to Q* and r*. This is indeed correct. In fact,
the r value obtained from the approximate model is r*, and the Q value
obtained treating Q as continuous comes within one unit of Q*. The se¬
quence of Q and r values obtained starting with Qx = Qw = 8 in (4-72) and
then using (4-73) to obtain the new Q is: Qx = 8, rx = 105; Q2 = 12.9,
r2 = 101; Qz = 15, r3 = 99; Q4 = 17.5, r4 = 98; Q5 = 18.4, r5 = 97; Q6
= 19.4, r6 = 96; Qt = 20.3, 7*7 = 96. It should be pointed out that the
manual computations with the approximate model are quite arduous and
required about three hours of computation, whereas, working with the exact
cost expression, Q* and r* were obtained in only about 15 seconds on the
Burroughs computer (which is not a really high speed computer).
It is interesting that the approximate model did so well in this case
where t and f are exceptionally small. At the outset one might have
expected that it would do rather poorly. It would appear that the approxi¬
mate model of Sec. 4-8 is applicable under very general circumstances. It
might be noted that if we attempted to use the simple, approximate model
of Sec. 4-2, which neglects the backorders term in the carrying cost and
SEC. 4-11 THE LOST SALES CASE FOR CONSTANT LEAD TIMES 197
Interestingly enough, the exact equations for the lost sales case are much
more difficult to develop than the corresponding equations for the back¬
orders case. In fact, the exact equations have not been developed for the
lost sales case when more than one order is allowed to be outstanding
except for the case where Q = 1. In this section we wish to examine what
makes the lost sales case so difficult to treat, and to derive the exact
formulas for the case where only a single order can be outstanding. In
Sec. 4-13 we shall consider the case where units are ordered one at a time
(.Q = 1), but any number of orders are allowed to be outstanding.
Consider a system in which demands occurring when the system is out
of stock are lost. Again, as with the backorders case, we imagine that the
stochastic processes generating the demands and lead times are such that
the system can be described by a Markov process for which there exists
a steady state. To be specific, imagine that the procurement lead time is
a constant r and that a Poisson process generates demands with units being
demanded one at a time. As usual, assume that we desire to determine the
optimal values of Q and r when the system is operated using a lot size-
reorder point control policy.
If the inventory on hand plus on order is used to define the reorder point
r, then the quantity on hand plus on order fluctuates between Q + r and
r’+ i during each cycle. The states are not necessarily, however, uniformly
distributed. The reason for this is that the amount on hand plus on order
will not move from t -f- j to r -|- j — 1 when a demand occurs if the system
is out of stock. When the system is out of stock the amount on hand plus
on order does not change when a demand occurs. Unlike the inventory
position in the backorders case, it is not possible in the lost sales case to
treat the changes in the amount on hand plus on order independently of
the amount on hand. This is in part what makes it so difficult to treat the
lost sales case. It is necessary to take explicit account of the number of
orders outstanding and the times at which they were placed. However,
even if it was possible to obtain the distribution of the amount on hand
plus on order, additional difficulties would be encountered. The procedure
198 LOT SIZE—REORDER POINT MODELS CHAP. 4
used to compute the distribution of the on hand inventory from the inven-
tory position in the backorders case will not work here. It is still true that
everything on order at time t — r will arrive in the system by timp t and
nothing not on order at time t — r can arrive by t. However, it no longer
follows that if the amount on hand plus on order is r+j at time t - r, the
probability that x > 0 units are on hand at time t is the probability that
»• + 3 — x units were demanded in the time r. The reason is that it is possi¬
ble for the system to run out of stock and have one or more demands occur
while the system is out of stock in the period t - T to t. Those demands
occurring when the system is out of stock are lost, and hence x units can be
on hand at time t even if more than r + j — x demands occurred in the lead
time.
In the backorders case there could be any number of orders outstanding
when a Poisson process generated demands. Since there cannot be back¬
orders m the lost sales case, the number of orders outstanding cannot be
greater than the largest; integer less than or equal to (Q + r)/Q = 1 -)-
(r/Q). The maximum number of orders which can be outstanding is there¬
fore determined by the values of r and Q. If r<Q then there can never
be more than a smgle order outstanding. In the lost sales case then, it is
possible to stipulate that there is only a single order outstanding if one
requires that r < Q.
We shall now derive the exact equations for the lost sales case when the
stipulation is made that there is never more than a single order outstanding
I o obtain the average annual cost we shall here first compute the expected
cost per cycle and then multiply by the average number of cycles per year
rather than attempting to find the state probabilities and compute the
costs directly as was done in the backorders case. The reason for changing
the approach is that it is not too easy to compute the state probabilities
Let us first compute the expected length of time per cycle during which
the system is out of stock. Since there is never more than one order out¬
standing, nothing is on order when the reorder point is reached, i.e., at the
reorder point r units are on hand. If the system reaches an out of stock
condition in the time interval t to t + dt after the reorder point is hit, this
means that m the time 0 to t, r - 1 units have been demanded and the rth
one is demanded between t and t + dt. This probability is \p(r - 1; \t) dt.
If the system does reach an out of stock position between t and t + dt it
will be out of stock for a length of time r - t during the cycle. Hence the
expected length of time out of stock per cycle is
From (4-37) the average number of cycles per year is X/(Q + XT) where
T is given by (4-100).
It is very easy to compute the expected number of lost sales per cycle.
If the demand in the lead time r is x > r, the number of lost sales is x — r.
Thus the expected number of lost sales is
by A3-10. However, by A3-13 we see that the expected time out of stock
(4-100) is the expected number of lost sales times the expected time between
demands 1/X. This correspondence is what one would expect intuitively.
It remains to compute the expected unit years of stock held per cycle.
First note that the probability that w units are on hand when the order
arrives is P(r; Xr), for w = 0, and p(r — w; Xr), 1 < w < r. These are also
the probabilities that w + Q units are on hand after the arrival of a
procurement. Thus the probability that v units are on hand after the
arrival of a procurement is
j P(r; Xr), v = Q
(4-102)
{p(Q + r — v; Xr), Q < v < Q + r
The expected unit years of stock held per cycle will be computed in two
parts. First, for the time period up to the time the reorder point is reached
(this time is a random variable), and second for the period of fixed length
t from the time the reorder point is reached until the next order arrives.
Given that v units are on hand after an order arrives, there will be v units
in stock until the first demand occurs, and there will be v — 1 in stock
from the time that the first demand occurs until the second occurs, etc.,
and r + 1 in stock from the time the (v — r — l)th demand occurs until
the (v — r)th demand occurs which reduces the on hand inventory to the
reorder point. The mean time between demands is 1/X. Thus, given that
the on hand inventory is v, the expected number of unit years of stock held
until the reorder point is reached is
Averaging over the initial inventory, i.e., over v, it is seen that the ex¬
pected unit years of stock held until the reorder point is reached is
200 LOT SIZE—REORDER POINT MODELS CHAP. 4
2 Q+r -j
1 r_1
+ o7 X) t(r + C ~ w)(r + Q + 1 - «) — r(r + l)]p(w; Xr) (4-103)
A u= 0
The expected unit years of stock held from the time the reorder point is
reached until the next order arrives is the integral from 0 to r of the
expected amount on hand at time t, i.e.,
fr r— 1
On expanding out (4-103) and (4-104) and summing the results, we find
that the.expected number of unit years of stock held per cycle is'(to~Be'
worked -out-in Problem 4-62) . -— -^
where x is the cost of a lost sale including the lost profit and T is given bjr
(4-100). We ask the reader to prove in detail in Problem 4-74 that mini¬
mization of the average annual cost with the cost of a lost sale being defined
to include the lost profit will yield the same Q* and r* as the maximization
of the average annual profit.
It is interesting to note that when T is negligible, (4-106) becomes simply
the discrete version of the simple lost sales model studied in Sec, 4-3. In'
practice it is rather difficult to determine the optimal values of Q and r
by use °f (4-106). Fortunately, for most real world computations, it is
almost always true that the simple model will suffice, and it is unnecessary
to use (4-106). The reader should recall, of course, that (4-106) holds rigor¬
ously only if never more than one order is outstanding, and that the theory
for more than a single order outstanding has not been worked out.
of the (Q, r) models presented in the last several sections have assumed
that the lead time was constant. We would now like to investigate the
problems involved in attempting to properly account for lead time vari¬
ations in the exact models. As long as there is never more than a single
order outstanding, no theoretical difficulties are encountered. If it is
imagined that the procurement lead time can be described by a random
variable with density function g(r), it is only necessary to average the ex¬
pected annual cost for a given r over r to obtain the appropriate average
annual cost. Thus if 3C(Q, r, r) is the average annual cost for a given r,
the appropriate average annual cost averaged over r is
In practice, it is almost always true that orders are received in the same
sequence in which they were placed, so that orders cannot cross. If this is
true, then lead times cannot be considered to be independent random
variables, i.e., the time of an arrival of an order placed at time t can depend
on the times of arrival of the other orders on the books when the order is
placed at time t For example, if orders are backed up at the source for
some reason, new orders will also have to wait in line until the previous
orders have been processed. Unfortunately, there seems no easy way to
handle situations where the lead times are not independent. In general, it is
very difficult to describe the dependence between lead times without
developing a detailed model of the source. No work has appeared in the
literature as yet which deals with models having the lead times being
dependent random variables.
1 f* JL i Q
= QJ0 Z-/ P(r + i “ x'> Wfftf dr = - h(r + j - x) (4-108)
SEC. 4-12 STOCHASTIC LEAD TIMES 203
Then, for Poisson demands, the system can be solved by using Markov
analysis for a system continuous in time and discrete in space, since Poisson
processes generate all transitions. This has been carried out in [3]. The
details will not be presented here since the state probabilities become fairly
complicated. Furthermore the specific lead time distribution is not very
realistic, usually, from a practical point of view, and in addition, since
orders do not normally cross, the generalization is not particularly helpful
in the real world. We shall, however, illustrate the method of analysis for
a particularly simple case in the following section.
Often in the real world it is true that even although two or more orders
are outstanding at any point in time, the interval between the placing of
orders is usually large enough that there is essentially no interaction between
orders, and to a good approximation, it can be assumed that the lead times
are independent as well as assuming that orders do not cross. In this case
(4-108) is correct, i.e., to develop the appropriate expected values for the
cost expression it is only necessary to replace in the discrete case p(x; At)
by h(x).
When treating lead times as independent while simultaneously making
the assumption that orders do not cross, it is often convenient when demand
is Poisson distributed to assume that the lead time density can be described
by a gamma distribution. In practice, the lead time distribution, when
known, can often be fitted fairly well by a gamma distribution. From
Sec. 3-7 we know that if the random variable representing demand in any
time period is Poisson distributed and the lead time has a gamma distri¬
bution, then the marginal distribution of lead time demand has a negative
binomial distribution. Thus to compute the expected costs it is only
necessary to replace p(x; Ar) by bN[x; a + 1, + A)] in the expressions
for \pi(x) and (y)- For hand computations, of course, it is much more
difficult to work with the negative binomial distribution rather than the
Poisson since the negative binomial is not well tabulated. Intuitively, as
the variance of the procurement lead time distribution increases, the
variance of the distribution of states for the net inventory increases. This
in turn means that the safety stock and r must be increased. The compu¬
tation of Q and r treating the lead time as a constant equal to the mean
procurement lead time, when in actuality the procurement lead time is a
204 LOT SIZE—REORDER POINT MODELS CHAP. 4
random variable, can lead to carrying a safety stock which is much too low.
The amount of the error increases as the variance of the lead time distri¬
bution increases. When the model is being used which treats Q and r as
continuous and assumes the demand variable to be normally distributed,
one usually assumes that the marginal distribution of demand is also
normally distributed and the variance of the normal distribution is chosen
to give the proper variance. In this case the formulas of Sec. 4-8 can be
used for variable lead times without modification.
Similarly ^2 (y), the probability that there are y > 0 backorders at time t is
Thus Pout and the expected number of backorders B(R) are respectively
oo oo
Note that the average annual ordering costs \A are independent of R and
hence need not be included in 5C.
If R* is the smallest P which minimizes 3Z(R), then it is necessary that
R* satisfy
A3C(R*) < 0; AX(R* + 1) > 0 (4-115)
Now
A3C(R) — IC — xXp(P — 1 ,* Xr) — (IC + f)P(R; Xr)
P(P;Xr)>T^ (4-117)
Having obtained the average annual cost and the optimality condition
for the ease of constant lead times, we now turn to the case where we allow
the lead times to be random variables. To begin, we shall imagine that lead
times are generated by a Poisson process, i.e., the lead time density is expo-
206 LOT SIZE—REORDER POINT MODELS CHAP. 4
nential, with mean r = 1/8, and that the lead times are independent random
variables. This implies that orders can cross and need not be received in the
same sequence in which they were placed. The problem can be solved in
this case as a Markov process. Let the states be the net inventory (there
are an infinite number of these). The probability that the system moves
from state v to v — 1 in time dt is the probability that a demand occurs,
i.e., X dt. The probability that the system moves from state v to v + 1 in
FIGURE 4-6.
M = g = Xr
Since
- k) = = *■*(«) = 1
fc = 0 jfc = 0 /c*
it follows that
<K-R) = e-* (4-120)
and from (4-119)
Comparison with (4-109) and (4-110) leads to the interesting result that
precisely the same state probabilities are obtained as for the case of con¬
stant lead times, i.e., it makes no difference whether the lead time is a
constant r or whether the lead time is exponentially distributed with mean
r, provided that when the lead time is a random variable the lead times
corresponding to different orders are independent random variables. The
same optimal R will be obtained in either case. This result is peculiar to the
case where Q = 1. If Q > 1 it is not true that the state probabilities for
the constant lead time case are the same as where the lead times are
exponentially distributed, independent random variables. This is shown
in [3], referred to above, where the state probabilities are computed for
arbitrary Q when the lead times are exponential, independent random
variables.
An even more surprising result can be proved for the case where Q = 1.
It says that if demand is Poisson distributed, then when Q = 1, and the
lead times are independent random variables (i.e., orders can cross), (4-121)
gives the state probabilities for any lead time density g(r) with mean f.
In other words, the state probabilities and the optimal value of R are
independent of the nature of the lead time distribution if the lead times are
independent. The proof which we shall present is patterned after that of
Takacs [5].
208 LOT SIZE—REORDER POINT MODELS CHAP. 4
while event i does occur between k and k + dk. Equation (4-122) can be
condensed to
\ne~u dk ■ ■ . dtn (4-123)
We leave to Problem 4-63 the proof that n\/tn is a legitimate density func¬
tion. We note from (4-124) that since the density function is independent
of the k, the occurrences of the different events are independent random
variables. Furthermore, the probability that any one occurs between
k and k + dk is simply dk/t. To see how the n\ comes about in the joint
density it can be imagined that the time interval from 0 to i is divided up
to yield 2n + 1 boxes, n boxes corresponding to the time intervals k to
k + dk and the remaining n + 1 corresponding to the remaining n + 1
time intervals. The Poisson events can be imagined to be balls which are
tossed into the boxes. Now suppose that the n balls are tossed into the
oxes. All balls must go into one of the boxes, and for boxes corresponding
to the tune intervals k to k + dk only one ball can fit into a box. The
probability that one of the balls goes into the box corresponding to the
interval k to k + dk is then n dk/t. Given that one of the balls goes into the
box corresponding to the interval k to k + dk, the probability that one of
the n — 1 remaining balls goes into the box corresponding to the interval
k to k + dk is (n - 1) dk/t, etc. Thus the probability that one of the balls
goes into each of the n boxes corresponding to the intervals k to k + dk
SEC 4-13 MODELS WITH Q = 1 209
is (4-124). What we have shown is that if we are given that at least one
Poisson event occurs in the time interval 0 to t, then the probability that
any one occurs between £ and £ + d£ is d£/1, and this is independent of how
many events have occurred in the interval or of their times of occurrence.
Let us now return to the proof that the state probabilities are inde¬
pendent of the lead time distribution. The probability that a unit ordered
at time U will arrive by time t > U is
rt-u
S(t- ti) = I g(r) dr (4-125)
Then if we know that at least one demand has occurred in the time interval
0 to t, the probability that any one occurred between U and U + dti and
that the unit ordered arrives by time t is
Suppose now that the system has been operating for a length of time t.
We assume that R units were on hand initially. Let us compute the proba¬
bility that the net inventory is x at time t if there have been in total n de¬
mands since the system began operation. If the net inventory is x, then
n + x — R of the orders have arrived and R — x have not arrived.
The probability of this, which will be written q(x] n, t) is simply the
binomial probability of n + x — R successes in n trials when the proba¬
bility of a success is given by (4-126), i.e.,
q(x; n, t) =
(B-x)[fjfs®*r',[rjfii-s®^r ^
On weighting q(x; n, t) by the probability of having n units demanded, we
find that Wx(t) the probability that the net inventory is x at time t is
^*(0 = X)
n = R~x
n> 0
[1 - £(£)] d$x-*
e-*Jo (4-128)
(R - a) I
210 LOT SIZE—REORDER POINT MODELS CHAP. 4
“X *5{«-X (4-129)
i*e., the limiting state probabilities are Poisson and they depend only on the
mean procurement lead time f.
We have demonstrated above that when Q = 1, demand is Poisson dis¬
tributed, and lead times are independent (and orders can cross), then the
state probabilities and the optimal R are independent of the nature of the
procurement lead time distribution. Let us now examine what happens
if we do not allow orders to cross but instead require that they arrive in
the same sequence as placed. The procedure suggested in Sec. 4-12 will be
used in which it is imagined that lead times can be treated as independent
and that they arrive in the same sequence in which they were placed. Then,
if g(r) is the lead time density, the average annual cost is given by (4-107).
Suppose now that g(r) is a gamma distribution, i.e., g{r) = y(r; a, 0).
Then, from (3-73), we know that the marginal distribution of lead time
demand has the negative binomial distribution bN[x; a + 1, p/(p + X)].
Hence the average annual cost has the form
oo
00
The optimal R computed from (4-132) will in general be larger than the R
computed from (4-116). It can be considerably larger if the variance of
the marginal distribution of lead time demand is much greater than the
variance of the Poisson distribution for the mean lead time. Thus, for the
case of Q = 1, the procedure suggested in Sec. 4-12 yields a higher variance
SEC 4-13 MODELS WITH Q = 1 211
of the net inventory distribution, and will in general yield a larger R* than
would have obtained under the assumption that orders can cross.
To close this section let us examine briefly the lost sales case. Even when
Q = 1, it is not easy to treat rigorously the case where more than a single
order can be outstanding (here R would have to be one in order to specify
rigorously that no more than one order could be outstanding—this case is
of no interest). However, the problem can be solved if it is assumed that
the procurement lead time has an exponential distribution with mean
f = 1/8 and that lead times are independent random variables (and orders
can cross). Markov analysis can be used to solve the problem just as it
could for the backorders case under similar assumptions. If x is the value
of the on hand inventory at any point in time, x can only have one of the
values 0, 1, 2, . . ., R. Fig. 4-7 shows the transition diagram from which
the balance equations can immediately be written down.
If \f/(x) is the steady state probability that x units are on hand at time t
then
Thus
(4-134)
*(o) - (RT-
and since
M"
II
17 = 0
it follows that
Hence
*=[§(r;ri^r (4-135)
17 = 0 U'
p(R — v; Xf)
#(v) = v = 0, 1,. . ., R (4-137)
1 — P(R + 1; Xr)’
It has been shown in [1] that the are given by (4r-137) if lead times are
independent and orders can cross, for any lead time distribution g(r) such
that the cumulative distribution is continuous (the proof did not include
the case of constant lead times, however). Conny Palm [4] has also
presented a proof that the V'(j) are independent of the lead time distri¬
bution if lead times are independent and orders can cross. We shall not
present either of these proofs.
When the state probabilities are given by (4-137), the average annual cost
is
l-P(R;\t) ~ p CR;Xf)
3C = IC *4" Xt (4-138)
1 — P(R + 1; Xt) _ l-P(R + 1; Xf)
provided that the unit cost of the item is constant. We leave a discussion
of the determination of the optimal R to Problem 4-64.
The preceding analysis has assumed that the unit price of the item was a
constant independent of the size of the order quantity Q. In the real world
quantity discounts are frequently available on the purchase of large quanti¬
ties. Two types of discounts sometimes encountered are all units and
incremental quantity discounts discussed in Secs. 2-11, 2-12 respectively.
Precisely the same computational procedures that were valid in the de-
SEC 4-15 CONSTRAINTS 213
4-15 Constraints
The previous discussion in this chapter has assumed that the item under
consideration can be treated independently of any other items carried by
the system. When more than a single item is carried there will almost al¬
ways be some sort of interaction between the items. These interactions
may or may not have an important influence on determining the optimal
Q and r values. When they are not important, the models already discussed
may be used to determine Q* and r*. When the interactions are of conse¬
quence, things become more complicated, and some attempt must be made
to account for the interactions. Typical simple interactions are of the form
discussed in Sec. 2-8, i.e., competition for warehouse space, for the limited
number of orders that can be placed, or for the allowable total investment
in inventory. Formally, these constraints can be handled in much the same
way as in Sec. 2-8. However, as we shall see, this formal generalization
may not always be satisfactory. It will also become apparent that it is not
a simple matter to develop a satisfactory means for treating these con¬
straints when demand is a stochastic variable.
Assume that the system stocks n items. Consider first the case where
there is a limitation on the total number of orders (or setups) which can be
handled per year. Immediately a problem of interpretation arises. Because
of the stochastic nature of the demands, it is theoretically possible (if the
demand variable is being described by a Poisson or normal distribution) to
have any arbitrarily large number of orders placed in one year. Thus one
cannot guarantee that there will never be more than a given number of
orders placed per year. One can, however, talk about the probability that
no more than a given number of orders will be placed, or we can require
214 LOT SIZE—REORDER POINT MODELS CHAP. 4
that the expected number of orders placed be less than or equal to a given
value.
The situation where one places a limitation on the expected number of
orders placed is the easiest to handle. The expected number of orders
placed per year for all n items is
(4-139)
it follows that for any v, J will be minimized when each of the n expressions
rj) is minimized. The computational procedure then is to minimize
each 3j(Qj, Tj) for a given 17 and thus determine a set of Qj and rj; compute
N and compare it with h, select a new i) and repeat the process until N = h.
The set of Qj, rj so obtained will be optimal.
It must be recognized that the above procedure only sets an upper limit
on the expected number of orders placed per year. In any given year the
number of orders placed could be greater than h or less than h. The compu¬
tation of the probability that precisely M orders will be placed is not
especially easy, in general, when there are n items, although it is easy
for a single item in certain cases. Let us then first consider the case of a
smgle item when a Poisson process generates demands, units are demanded
one at a time, and all demands occurring when the system is out of stock are
backordered. Thus if the inventory position is r + j at the beginning of
the year, precisely m orders will be placed (to > 1) if the demand lies in
the interval (to - 1 )Q + j to mQ+j- 1. The probability of this is
1 Q
0W = 0 E {p[(m - 1 )Q + j; X] -P(mQ + j; X)}, m > 1
^ y=i
q1 E P(u;\)-2 P(u; X)
^ ^u=(m-l)Q+l u = mQ+l
0W=xt[i-P(i;x)]
(4-144).
{X - XP(Q; X) + QP(Q + 1; X)}
Equations (4-143) and (4-144) thus give the probability that precisely m
orders are placed in a one year period for a single item.
When there are n items it is much more complicated to find the prob¬
ability that precisely M orders are placed. Let be the probability
that precisely m,* orders are placed for item j in one year. Assuming that
demands for the various items are independent, we see that the probability
that mi orders are placed for item 1, m2 for item 2, etc., is
Where Q(M) is a function of Qi, . . ., Qn but not of the 77. In this case
there is no constraint. The cost function to be minimized is
n 00
X) Q(M)< a (4-148)
M = k+1
to be not too important and can be neglected. Hence each item can be
studied independently of the others and the models developed in the
previous sections can be applied.
where f is the upper limit to the average fraction of time for which the
system is out of stock. Now it is clear that if the costs (4-149) are to be
minimized, the average fraction of time out of stock should be as large as
possible. Hence the constraint (4-150) will be active, i.e., E(Q, r) = X/
(or when Q and r are treated as discrete, E(Q, r) should be as close to X/ as
possible while not being greater than X/).
To minimize (4-149) subject to E(Q, r) = X/, we know from the theory
of Lagrange multipliers (see Appendix 1) that we form the function
218 LOT SIZE—REORDER POINT MODELS CHAP. 4
F{Q, r, 8) =
§l = §l = dF_
dQ dr dd (4-152)
since we do not expect the optimal values of any of the variables to lie on
the boundaries.
Note, however, that minimizing F(Q, r, 6) for a given d will yield the
same Q*{6) and r*(6) as minimizing
Consider next the case where the constraint requires that the expected
number of backorders at any point in time be less than or equal to a
specified value. Again it is clear that the constraint will be active, and
will have the form B(Q, r) = 8. Thus we wish to minimize (4-149) subject
to this constraint. From the theory of Lagrange multipliers and what has
been shown above, this is equivalent to minimizing
with respect to Q and r for a given 77, thus yielding Q*(r)) and r*(j)), and
then determining the optimal value 77* of the Lagrange multiplier 77 such
that £[$*(77), t**(t7)] = 8. Then $*(77*) and r*(7j*) are the optimal values
of Q and r. The numerical procedure for making the computations is the
same as that outlined above. It follows that specification of the expected
number of backorders is equivalent to setting x = 0 and uniquely deter-
mining x. Thus by specifying 8 we implicitly determine a unique value
of x when x = 0.
A more general procedure than the two possibilities discussed above
would be to specify both the average fraction of the time which the system
could be out of stock and the expected number of backorders at any point
in time. The theory of Lagrange multipliers immediately shows that this
would impute nonzero values to both x and x, since in this case two
Lagrange multipliers would be needed.
REFERENCES
PROBLEMS
4-1. Derive the equivalent of the model presented in Sec. 4-2 when a
Poisson process generates demands, units are demanded one at a
time, and the procurement lead time is a constant r. Treat Q and
r as discrete variables. Show that the following numerical procedure
can be used to determine Q* and r* To begin, set Q = Qw in the
following expression and determine the largest integer r, call it r±
which satisfies this inequality
P(r;Xr)>^
7TA
2X
Q(Q !) < J£f + 7r[XrP(r; Xr) — rP(r + 1; Xt)]}
and determine the largest Q, call it Q2, which satisfies this expression.
Use Q2 in the previous relation and determine r2 by finding the
largest r which satisfies the inequality, etc.
4-2. Derive the equivalent of the lost sales model presented in Sec. 4-3
when a Poisson process generates demands and the procurement lead
time is a constant r. Develop a numerical procedure for determining
Q* and r*.
4-3. Modify the results of Problems 4-1, 4-2 for the case where Q can be
treated as continuous, but r is to be treated as discrete.
4-4. Draw the equivalent of Fig. 4-2 for the lost sales case and prove that
the iterative scheme suggested in Sec. 4-4 converges in this case.
4-5. A function of two variables X(Q, r) is said to be convex over some
region of the Qr-plane if for any two distinct points (Qlt n) (Q2, r2)
m the region and for any a, 0 < a < 1,
is 500 cartons per year. He estimates that the paper work cost in
preparing an order invoice is $1.50 and the cost of receiving an
order and placing it in the storage bin, along with the accompanying
paper work on receipt of procurement, is $2.00 plus $0.15 times the
size of the order. A carton of solder costs $12.00. The warehouse
uses an inventory carrying charge of I = 0.18. All demands occurring
when the warehouse is out of stock are backordered. It is estimated
that the goodwill cost plus the cost of writing a special letter to the
hardware store if a demand occurs when the system is out of stock
amounts to $25.00. Compute the optimal order quantity and re¬
order point for the warehouse. What is the cost of uncertainty?
4-13. For the example worked out in Sec. 4-4, determine howQ* and r*
vary with A. Allow A to vary from $,500 to $50,000 and plot the
results.
4-14. A large discount house is planning to install a (Q, r) system to con¬
trol the inventory of a particular model of an AM-FM radio. The
number of units demanded has been found to be essentially Poisson
distributed with a mean of 5 radios per week (this does not include
the Christmas season demand which will be handled separately).
The procurement lead time is essentially constant and requires 3
weeks. Each radio costs the store $40.00. An inventory carrying
charge of I — 0.20 is used. If a demand occurs when the model is
not in stock, the customer will almost always go to another of the
nearby discount stores rather than selecting a different model or
waiting until a new shipment arrives. The store has decided that
each such demand occurring when the model is not in stock costs
the store a goodwill loss of $20.00 plus a loss in gross profits of
$25.00. The total cost of placing an order is estimated to be $3.00.
Determine Q* and r*. What is the cost of uncertainty? On the
average, how many lost sales will be incurred per year?
4-16. A military supply center stocks air bearing gyros for a ballistic
missile. The mean rate of demand for these gyros has been 100 per
year over the past three years and the number of units demanded
appears to be described quite well by a Poisson distribution. The
procurement lead time is essentially constant and has the value
6 months. The gyros cost $2000 each. The cost of placing an order,
incoming inspection, etc., is estimated to be $100.00. An inventory
carrying charge of / = 0.20 is used. All units demanded when the
system is out of stock are backordered. It is difficult to estimate
the cost of being out of stock. Instead, the requirement has been
made that the probability of being out of stock must not be greater
than 0.0005. If the system is to operate using a (Q, r) model, de-
224 LOT SIZE—REORDER POINT MODELS CHAP. 4
past year (excluding the Christmas season) has yielded the following
for weekly demand on a particular electric coffeemaker:
Frequency 2 6 3 8 2 4 2 1
There were never more than 7 units demanded in one week. The
unit cost of the coffee pot is $8.00 and the store uses an inventory
carrying charge of I = 0.13. The procurement lead time is one
week. The cost of placing an order is estimated to be $0.50. The
store feels that because of goodwill loss, the cost of a lost sale includ¬
ing lost profit is $10.00. Compute Q*, r* under the assumption that
the above data represent the true probability distribution for de¬
mand. What are the values of Q*, r* if one assumes that the
number of units demanded per week is represented by a Poisson
distribution whose mean is that obtained from the data given above?
4-23. An automotive repair shop installs new mufflers on autos. Past
history indicates that the number of units demanded for a certain
model of muffler is Poisson distributed with a mean of 1 per day.
The procurement lead time is always either 8 or 15 days, the prob¬
ability of 15 days being 0.7. A muffler costs the shop $6.00 and it
uses an inventory carrying charge of I = 0.20. The cost of placing
an order is estimated to be $1.00. Requests for muffler changes
which occur when the dealer is out of stock are taken elsewhere,
and the goodwill loss plus lost profits is estimated to be $25.00. If a
(Q} r) system is used to control the inventory of this muffler, what
are Q*, r*? What is the average annual cost of uncertainty? What
assumption was made about the nature of the lead times? Is it valid?
4-24. Work out the equations which determine Q* and r* for the simple
model discussed in Sec. 4-2 when
h(x) = \
0
given in the example apply here. Use the results of Problem 4-24
in making the computations.
4-26. Work out the equations which determine Q* and r* for the simple
model discussed in Sec. 4-2 when the marginal distribution of lead
time demand is gamma distributed with mean y, and standard
deviation a.
4-27. Solve the example presented in Sec. 4-4 when the marginal distribu¬
tion of lead time demand has a gamma distribution with the same
mean and standard deviation as the normal distribution in the
example.
4-28. Derive the exact cost equation and the equations which determine
Q* and r* under the assumption that the marginal distribution of
lead time demand is exponential. Assume that lead times are inde¬
pendent and that orders do not cross.
4-29. Derive the exact cost expression and the equations which determine
Q* and r* when the marginal distribution of lead time demand is a
gamma distribution with mean ju, and standard deviation <r. Assume
that lead times are independent and that orders do not cross.
4-30. Solve Problem 4-12 under the assumption that the marginal dis¬
tribution of lead time demand has an exponential distribution whose
mean is the same as that of the normal distribution referred to in
the problem.
4-31. Examine Molina’s tables and note that often by adding merely a
single unit of safety stock, the probability of incurring a backorder
or lost sale during the lead time can be reduced from a sizable value
to an exceptionally small value. Would one normally expect such
an increment in protection by the addition of a single unit in the
real world? Why or why not?
4-32. Is it possible when the demand variable is treated as continuous to
have the demand in any time period of length t described by a
gamma distribution with mean \ty instead of by a normal distribu-
tion, while still being able to describe the (Q, r) system as a Mark ov
process continuous in space and time? Why or why not?
4-33. A low demand, expensive spare part for an aircraft is stocked at a
large military depot. The average rate of demand is 3 per year.
The demands behave as if they were generated by a Poisson process.
Units are ordered one at a time as demanded. Each unit costs $2000
and the system uses a carrying charge of I = 0.20. Demands
occurring when the system is out of stock are backordered. There
is no fixed cost for a backorder, but the grounding of an aircraft for
PROBLEMS 227
lack of this part costs 110,000 per week. The procurement lead
time can be assumed to be a constant equal to 6 months. Determine
the optimal value of the inventory position which is to be main¬
tained at a constant value through time.
4-34. Solve Problem 4-33 under the assumption that the procurement
lead time has a gamma distribution with mean equal to 6 months
and standard deviation equal to 1 month. Imagine that lead times
are independent random variables and that orders are received in
the same sequence in which they were placed.
4-35. Solve Problem 4-33 assuming that in addition to the backorder cost
of $10,000 per week, there is a fixed cost per backorder of $5000.
4-36. Solve Problem 4-34 assuming that in addition to the backorder cost
of $10,000 per week, there is a fixed cost per backorder of $5000.
4-37. Plot a curve showing how in Problem 4-33 the optimal inventory
position varies with the backorder cost f.
4-38. Solve Problem 4-12 under the assumption that there is no fixed cost
for incurring a backorder but that instead a backorder costs $20.00
per week out of stock.
4-39. Solve Problem 4-15 under the assumption that instead of specifying
the probability of being out of stock, there is a cost of $10,000 per
unit week of shortage.
4-40. A paint store orders a type of one pint cans of red paint by the
carton. There are 24 pints in a carton and after each twenty-four
demands, a carton is ordered. The order size is always one carton.
One carton costs $12.00, and the store uses an inventory carrying
charge of I = 0.20. Demands occurring when the store is out of
stock are lost, and the owner estimates the cost of a lost sale to be
$1.00. Three weeks are required from the time that an order is
placed until the shipment is received. What is the optimal reorder
point?
4-41. Solve the example given in Sec. 4-4 under the assumption that
“all units” quantity discounts are available of the following sort:
the unit cost is $50.00 if 0 < Q < 1200; the unit cost is $49.50 if
1200 < Q < 2500; and the unit cost is $49.00 if 2500 < Q < °o.
4-42. Solve the example given in Sec. 4-4 under the assumption that
incremental quantity discounts are available of the following sort:
for all units between 0 and 1200 the unit cost is $50.00. For addi¬
tional units up to 3000 units the unit cost is $45.00. For all units
above 3000, the unit cost is $40.00.
228 LOT SIZE—REORDER POINT MODELS CHAP. A
4-43. Solve Problem 4-12 under the assumption that “all units” quantity
discounts of the following sort are available: the unit cost is 112.00
if 0 < Q < 50 and is $11.00 if 50 < Q < «.
4-44. Solve Problem 4-14 under the assumption that “all units” quantity
discounts of the following sort are available: the unit cost of a radio
is $40.00 if 0 < Q < 39, and is $37.00 if 40 < Q < 99 and is $33 00
if 100 < Q < oo.
4-45. Consider the problem of developing a (Q, r) system for an item when
the order quantity must be an integral multiple of a fixed package
quantity. Modify the simplified models developed in Secs. 4-2,
4-3 to take account of this, and obtain the equations which deter¬
mine Q* and r*.
4-47. For the model presented in Sec. 4-7 compute the probability that
precisely n orders are outstanding at any point in time. Also com¬
pute the expected amount on order and show that it is equal to the
expected lead time demand. Hint: Specification of the net inven¬
tory uniquely determines the number of orders outstanding.
4-48. For the model presented in Sec. 4-9 compute the probability that
precisely n orders are outstanding at any point in time. Also com-
pute the expected amount on order and show that it is equal to the
expected lead time demand. Hint: Same as for Problem 4-47.
4-49. Compute the mean of the distribution 6(m) defined by Eqs. (4-143),
(4-144) and show that it is equal to \/Q. ’
PROBLEMS 229
X) f Xir(T ~~ +3 - 1; K) d£
3-1 Jo
What then are the expected annual backorder costs? For the case
where ir(t) = w + ft show that the above formula leads to the same
results as those obtained in the text. Compute the average annual
backorder costs when
4-60. Discuss the use of floor space constraints in (Q, r) models with
stochastic demands when there are n items. Show that the same
sort of problems are encountered as those encountered in placing
a limitation on the number of procurements made.
4-64. How can R* be determined when the average annual cost is given
by Eq. (4-138)?
4-66. Obtain the E(Q, r) and B(Q, r) terms for the model of Sec. 4-7 by
first computing the expected number of backorders or unit years of
shortage incurred per cycle and then convert to an annual basis.
Hint: Take a cycle to be the time between the arrival of two suc¬
cessive procurements. Note that it is necessary to take explicit
account of the fact that the length of the cycle is a random variable.
4-66. Note that when units are demanded one at a time, there are pre¬
cisely Q demands in every cycle. Imagine here that a cycle refers to
the time between the placing of two successive orders. Then the
inventory position is r + Q until the first demand occurs, r + Q — 1
until the second demand occurs, etc., and r + 1 until the Qth demand
occurs. Compute the expected costs (displaced by a lead time) of
carrying inventory and of backorders incurred between the oeur-
rence of 'the jth and the 0’ + l)st demands. Then the corresponding
costs per cycle are the sum of these costs from j = 0 to Q — 1. In
this way compute the costs per cycle of carrying inventory and of
backorders when a Poisson process generates demands and the lead
times are constant. Hint: The time between demands is not con¬
stant but is exponentially distributed.
4-67. For any real world system the ultimate use of any item demanded
may vary widely from one demand to the next. It is conceivable
that a different backorder cost might be associated with each end
use. This is especially true in military supply systems where the
item may be used in a variety of ways. Suppose that there are N
different end uses and for end use i, the cost of a backorder is + iut.
Let fi be the fraction of the demands that will be for end use i (or
the probability that any given demand will be for end use i). Show
how to account for this sort of behavior within the framework of
the models developed in this chapter. Do they really implicitly
include this case also?
4-68. Combine the models developed in Secs. 4-2 and 4-3 to obtain a model
for the case where a fraction / of the demands occurring when the
system is out of stock are backordered while the remainder are lost.
Assume that the cost of a lost sale is different from the cost of a
backorder.
4-69. Solve Problem 4-12 under the assumption that instead of specifying
the cost of a backorder, it is required that the average fraction of
the time that the system is out of stock not be greater than 0.01.
What is the imputed cost of a backorder?
232 LOT SIZE—REORDER POINT MODELS CHAP. 4
4-70. Solve Problem 4-14 under the assumption that instead of specifying
the cost of a lost sale, it is required that the average fraction of the
time that the system is out of stock not be greater than 0.05. What
is the imputed cost of a lost sale?
4-71. Solve Problem 4-22 under the assumption that instead of specifying
the cost of a lost sale, it is required that the average fraction of the
time that the system is out of stock not be greater than 0.02. What
is the imputed cost of a lost sale?
4-72. Solve Problem 4-33 under the assumption that instead of specifying
the cost of a backorder, it is required that the average fraction of
the time that the system is out of stock not be greater than 0.005.
What is the imputed cost of a backorder?
4-73. Consider a system for which the mean rate of demand is constant
over time, and the number of units demanded in any given time
interval is Poisson distributed. The procurement lead time is always
a constant r, and demands occurring when the system is out of stock
are backordered. Imagine that there is no fixed cost of incurring a
backorder, and if there are any backorders on the books at time t,
the backorder cost incurred between t and t + dt is tt0 dt Note that
the backorder cost is independent of the number of backorders out¬
standing. Furthermore, assume that annual carrying costs are pro¬
portional to the maximum value of the inventory position. The
cost of placing an order is A, and the unit cost of the item is C
independent of the size of an order. Develop the average annual
cost expression for a (Q,r) model for the above item, and obtain
formulas for computing Q* and r*.
4-74. In the lost sales case studied in Sec. 4-11, prove that the minimization
of the average annual cost with the cost of a lost sale defined so as
to include the lost profit will yield the same Q* and r* as the maxi¬
mization of the average annual profit.
4-75. In order to treat r and Q as continuous and to differentiate 3C with
respect to these variables, what restrictions must be placed on h(x)
in the model of Sec. 4-2 if the derivatives are to exist for all Q > 0,
r > 0?
PERIODIC REVIEW MODELS
Newspaper Report
5-1 Introduction
develop and compare a set of models which use these three operating doc¬
trines in the case where the stochastic processes generating demands and
lead times are time invariant. The purpose of these models will be not
only to determine the appropriate values of r, R for a given T, but also to
determine the optimal value of T, the time between reviews, in cases where
this time is also a variable. Models which use an order up to R policy will
be referred to as (R, T) models, those which use an Rr policy as (R, r, T)
models, and those which use an nQ policy as (nQ, r, T) models. Of course,
any model which can be used to determine the optimal value of T as well
as R, r can also be used to determine the optimal values of R, r for
a given T.
In Chapter 4, the annual cost of operating the transactions reporting
system was not included in the average annual cost. The reason was that
it was assumed that this cost was independent of Q and r and hence did not
need to be included.
In this chapter, however, in order to determine the optimal value of T, it
is necessary to introduce the cost of making a review. The cost of making
a review will be assumed to be independent of the model variables. The
review cost will not include the cost of placing an order since, in general, an
order may not be placed at each review.
As in Chapter 4 we shall begin with two simple approximate models
which are also the most useful ones for practical applications.
In the real world the most widely used operating doctrine for periodic
review systems is the order up to R doctrine. We shall here present the
analogues of the backorders and lost sales models of Secs. 4-2 and 4-3
respectively, for periodic review systems using an order up to R policy.
Again all variables will be treated as continuous.
Consider first the backorders case. The time between reviews will be
denoted by T, and at each review time a sufficient quantity is ordered to
bring the inventory position of the system up to a level R. The problem
is to determine the optimal values of R and T. At the outset, the following
assumptions will be made:
1. The cost J of making a review is independent of the variables R and T.
2. The unit cost C of the item is constant independent of the quantity
ordered.
3. Backorders are incurred only in very small quantities. This will be
taken to imply that when an order arrives, it is almost always sufficient to
meet any outstanding backorders.
238 PERIODIC REVIEW MODELS WITH STOCHASTIC DEMANDS CHAP. 5
because of assumption 3 it must be true that the integral over time of the
net inventory must very closely approximate the integral over time of the
on hand inventory. Hence the expected unit years of storage incurred per
period is to a good approximation
T [| (R - M) + | (R - m - XT)] = T [i2 - „ - y]
IC [i? - M - y] (5-i)
It will be shown later when the precise expressions are derived that the
simple argument given above is correct in the case where the expected unit
years of backorders incurred per year is negligible with respect to the
expected unit years of storage. Note that the above argument applies even
if the procurement lead time is a random variable.
In order to compute the average annual cost of backorders, it is necessary
to compute the average number of backorders incurred per year. This will
be done by computing the expected number of backorders incurred per
period and multiplying by 1 /T. Consider first the case where the procure¬
ment lead time is a constant r. Then an order placed at time t will arrive
in the system at time t + r, and the next procurement will arrive in the
system at time t + r + T. After the order is placed at time £, the inventory
position of the system is R. We wish to compute the expected number of
backorders occurring between t + r and t + r + T. A backorder will
occur in this period under assumption 3 if and only if the demand in the
time period r + T exceeds 22. Assumption 3 assures us that after the
arrival of the order placed at time t there will be no backorders on the books,
and hence they must all be incurred between times t + r and t + r + T.
Hence the expected number of backorders incurred per period is(
with density g(r)1 and let rmm, rma3: be the lower and upper limits respec¬
tively to the possible range of lead time values. Then if n, r2 are the lead
times for the orders placed at times t and i + T respectively, the expected
number of backorders incurred per period must be
f Tnias CTnrni
/
/* <*>
(x — R)h(x; T) dx (5-3)
where
g(n) dn = 1
|| = 0 = IC-Z&(R;T)
where
T) = h (x; T) dx (5-8)
SEC. 5-2 SIMPLE, APPROXIMATE (R, T) MODELS 241
fi(R-, T) = — (5-9)
7T
R — n — XT + J (x — R)h(x; T) dx (5-10)
ic [s - m - y+ &] (5-n)
242 PERIODIC REVIEW MODELS WITH STOCHASTIC DEMANDS CHAP. 5
The average annual variable cost for the lost sales case is thus
3C = ^ + /C (x — R)h(x; T) dx (5-12)
ig* - 450
1.881
25.981
and
ig* = 450 + 48.8 = 498.8 « 499
From (5-5) and A4-25 the average number of backorders incurred per year
is
SEC 5-2 SIMPLE, APPROXIMATE <R, 7} MODELS 243
Thus the average fraction of time out of stock is --g-inr = 0.00196, which is
small as the use of the model requires. Note that since r = 6 months
and T = 3 months, there will always be precisely two orders outstanding.
Suppose now that the combined cost of ordering and reviewing is $25.00.
Let us then determine the optimal value for T, the time between reviews.
This will be done simply by tabulating 5C as a function of T, where for
each T the optimal R for that T is used. The cost expression for the case
under consideration is
25
X 3 [R - 300 300 T]
T'
~R - 600(0.5 4- T)'
+ 25T -f\/9oo(°- 5 + T) <j>
l - V900(0.5 + T) .
R - 600(0.5 + T)'
+ [600(0.5 + T) #i*rL V900(0.5 + T) .
TABLE 5-1
Data for Example
Average
Annual Average Average
Safety Review and Annual Annual Average
T R*(T) Stock Ordering Carrying Backorder Annual
(months) (units) (units) Cost Cost Cost Cost X
in Table 5-1 have been rounded off to the nearest dollar value. The back¬
order costs jump around somewhat because R is rounded to the nearest
integer.
The behavior of the safety stock in Table 5-1 deserves some comment.
At first it seems rather surprising that it should decrease with increasing T.
The reader can easily check that it continues to decrease even for higher T.
Intuitively, the reason for this is that the more frequently that the net
inventory gets down in the neighborhood of the safety stock, the greater
will be the required safety stock to prevent unduly large backorder costs.
However, m the periodic review model being studied, the net inventory
^ neighborhood of the safety stock once each period. Hence, the
,, TT T’ thf, fger the safety stock required. Another peculiar feature is
that the model has the safety stock approach infinity as T 0 We know
intuitively that this is not correct. The problem lies in the formulation of
the model. We have assumed that there are no backorders on the books at
the time an order arrives. This assumption is not valid when T is small
lne exact formulation given later will rectify this problem.
SEC. 5-3 EXACT FORMULATION OF THE <nQ, r, T) MODEL 245
5-3 The Exact Formulation of the (nQ, r, T) Model for the Backorders
Case with Poisson Demands and Constant Lead Times
At this point, instead of obtaining the exact form of the {R, T) models in
certain cases, we shall instead move on to obtain the exact equations for the
{nQ, r, T) model in the backorders case where the number of units de¬
manded in any time period can be described by a Poisson distribution and
the procurement lead time is always a constant r. Recall that an (R, T)
model is a special case of the (nQ, r, T) model with Q = 1 and R = r + 1.
Thus once having obtained the exact equations for the (nQ, r, T) model
we can immediately obtain the exact equations for an {R, T) model under
the same assumptions which apply in deriving the (nQ, r, T) model.
We shall in the derivations treat the inventory levels as discrete variables
as well as the demand variable. Recall that an nQ operating doctrine is one
such that, at a review time, an order is placed if and only if the inventory
position y of the system at the review time is less than or equal to r. If
y < r, then a quantity nQ is ordered, n = 1, 2, 3,. . ., where n is chosen so
that r<y + nQ<r + Q. Thus, immediately after a review, the inven¬
tory position of the system will be in one of the Q states r + 1, r + 2,
. . •, t + Q.
To begin, we shall compute the steady state probability p(r + j) that the
246 PERIODIC REVIEW MODELS WITH STOCHASTIC DEMANDS CHAP. 5
r + i - d + nQ = r + j, n = 0,1, 2,. . .
or
However, when j > i, the system cannot be in state j unless the demand
has been at least i — j -j- Q. Thus
From Sec. 3-10 we know that the p(r + j) must satisfy the equations
SEC 5-3 EXACT FORMULATION OF THE (nQ.r.T) MODEL 247
oo
= E P(fc; 20 = 1, 3 = 1, • • •, Q (5-18)
k=0
Thus in the periodic review ease the distribution of the inventory position
immediately after a review is uniform. Recall that for (Q, r) models the
distribution was uniform at each instant of time. Note that (5-19) hoi(js
for any distribution ot demand as long as the demands in different periods
are independent. Also observe that, each <Ui > 0, so that from the results
of Sec. 3-10 we know that (5-19) is the unique solution to (5-17)
We are now in a position to develop the average annual erst expression “
At this point we shall restrict our attention to the ease when- units are
demanded one at a time, and a Poisson process generates the demands
[In principle, we could consider a case where the order size is a random 7
variable, say the stuttering Poisson distribution, but t he equations become '
exceptionally complex and difficult to work with) We shall here also restriefr
our attention to the case where the unit cost of the item is a constant ^
independent of the quantity ordered. The cost, of a backorder will I*
assumed to have the form tt + n when* t is t he length of time for which the
backorder exists. The costs of review and placing an order will be denoted
by J A respectively. As usual, the inventory carrying charge will he de¬
noted by I. The average annual cost, will be computed by determining the
expected cost per period and multiplying by 1 T to obtain the average
annual cost. Problem 5-12 asks the reader to go through the same sort, of
analysis used m bee. 4-0 to show that this is correct.
Since there are on the average l/T reviews per year, the average annual
cost of reviews is J/T. The average annual ordering cost is not A/T si me
an order need not be placed at each review. Instead, the
Sll hi Sr? 7 ™Ap"/T> whcrti P" * tin probability that an order
will be placed at any given review time. Let us now compute [f the
inventory position of the system is r+j immediately after a review the
probability that it will be less than or equal to r at the time of the next
review is the probability that; or more units are demanded in the time T
tfiTESEVf» *****
is r + j immediately after a review is 1 /Q, Tims
Vor itl-(rAT)
and from A3-6
\T
Vor
Q [1 - -P(Q;XT)3 + P(Q + l; XT’) (5-20)
periodPZposyeTat1afevfet:tlkjs
it,view taKos pH<-e
place 7 timeTr'^
at f'"1 ha<’k<,nicr‘S
t. The next P<ir
review takes
SEC 5-3 EXACT FORMULATION OF THE (nQt r, T) MODEL 249
and at time t + r
On averaging over the states j and dividing by T, we see that the average
number of backorders incurred per year is
where
T (x - r - j)p(x; X£)
x = r+j
1 H~Q /*r+ T 00
B(Q, r,T)=—*r / y) (x - u)p(x; Xg) d£
^ u = r-\-l Jr x=u
or from A3-10
l HiQ rr+T
B(Q, r,T) = ~ 22 / - 1; X£) -uP(u] X?)] d| (5-25)
V u-r-fl Jr
However,
The detailed derivation of (5-26) through (5-29) is left for Problem 5-13.
The average annual cost of backorders is then irE(Q, r, T) + fB(Q, r,T).-
It remains to evaluate the average unit years of storage incurred per
year. Again the procedure used above will be applied. The expected unit
years of storage incurred between t + t and t + r + T will be multiplied by
1 IT to obtain the desired answer. If the inventory position of the system
was r + j immediately after the review at time t, the expected value of the
net inventory at any time | between t + r and t + r + T is
oo
J2 (r + j ~ x)p(x; X£) = r + j - X£
x=0
The expected value of the on hand inventory at any time, is the expected
net inventory plus the expected number of backorders. Thus at time £, the
expected on hand inventory is
oo
0 1 XT
= ^ + 2 + r-r> (5-31)
where n is the expected lead time demand. Of course, (5-31) is also the
expected value of the on hand inventory at any point in time. The pro¬
cedure used to compute E, B, and D here differs from that used to compute
the corresponding quantities for the (Q, r) model in Chapter 4, because here
252 PERIODIC REVIEW MODELS WITH STOCHASTIC DEMANDS CHAP. 5
x = ¥ + 1 Vm + IC [:2 + \ ~ 0 - y] + r, T)
+ (* + IC)B(Q, r, T) (5-33)
Just as with the exact form of the (Q, r) model of Chapter -1, it is difficult
to determine Q*, r*, and T* for the (nQ, r, T) model manually. Again it is
the terms in E(Q, r, T) and Ii(Q, r, T) involving r + Q which cause the
difficulty. Here, however, unlike the situation encountered for the (Q, r)
model, it is not in general a good approximation to drop these terms. It is
quite possible that the optimal Q will be small even though the average
quantity ordered is not; in such cases, the r + Q terms will not be negli¬
gible. To determine Q* r* and T*, a digital computer will usually be
required to perform the search needed to determine, the values of Q, r, and
T which minimize 3C.
SEC 5-3 EXACT FORMULATION OF THE <nQ, r, T> MODEL 253
A code to compute Q*(T) and r*(T) for the (nQ, r, T) model discussed
above has been written by P. Teicholz and B. Lundh of Stanford University
for the Burroughs 220 computer. The search procedure used in this code
determines a relative minimum, but does not necessarily determine the
absolute minimum if there exist a number of relative minima. The results
presented in the following example were computed using this code.
The above example seems to suggest that an order.up toJg policy will
be
have guessedjndgmalLy. It is only when the probability of placing an order
at any review is not too high that an {nQ, r, T) model can yield significantly
lower costs than an {R, T) model. There do arise cases where the review
interval will arbitrarily be set to a very small value so that the periodic
review system becomes almost a transactions reporting system (viz., a
computer will be used to process all transactions, but instead of doing it on
254 PERIODIC REVIEW MODELS WITH STOCHASTIC DEMANDS CHAP. 5
a transactions basis, it will be done once per day or once every couple of
days). In such situations Q* may be quite large compared to the average
demand in the time between reviews, and here, of course, an nQ operating
doctrine would be much preferred to an order up to R doctrine. However,
for these cases one could almost use the results of a (Q, r) model for the
periodic review Q and r. It may not always be possible to do this, though,
We shall here consider the case where Q is sufficiently large to justify ignor¬
ing the r + Q terms in E(Q, r, T) and B(Q, r, T) in (5-33). Then (5-33)
reduces to
„ j , A . TnrQ .1 . XTl
X = f + y Vor + rc|_2 + 2 + r - M - ~2\
+ | A(r, T) + T(r, T) (5-34)
Next consider Ar3C(Q, r). The following are needed in performing the
differencing:
In each of the above, the differences were converted to a form which involve
only p(r; \t) and P(r; Xi).
It follows that
We shall now work out the equations for the (nQ, r, T) model for the case
where the demand density for a time interval of length t is normal with
mean Xi and variance Dt. The demand variable in any relevant time inter¬
val will therefore be treated as continuous. Furthermore, Q and r will also
be treated as continuous variables.
To begin we shall show that if v(x; T) is the density function for the
demand per period (x assumed continuous), then, immediately after a
SEC. 5-5 NORMALLY DISTRIBUTED DEMANDS 257
p(r + y) dy = 1
it follows that
(5-50)
[0, otherwise
We are now ready to work out the various average annual costs for the
case where demand is normally distributed. The average annual review
cost is again J/T, and the average annual ordering cost is Apor/T where
the probability of placing an order at any review time is now
1i /■* /tf-xrw.
fQ^/y- \T\ , VDT f(Q-^/VoT
f .
" « j. 4 \VW) dV-—
*«)# (5-51)
where B(Q, r, T) is, of course, the average number of unit years of shortage
incurred per year, i.e., the expected number of backorders at any point in
time.
It remains to obtain explicit formulas for the average number of back-
orders incurred per year and the average number of unit years of shortage
incurred per year. The same method of analysis is used as for the discrete
case studied in Sec. 5-3. It is clear that the average number of backorders
incurred per year is
~ f5'54)
However, by A4-25
(5_55)
Thus from A4-6, A4-26,
Dt
2
and
Similarly
1 rQ rr+T
lj£-r-y)vsi*(tvm)did,d»
(5-59)
1 fr+Q fr+T r*
d£ dt du
qt L I L
We shall evaluate B(Q, r, T) in several steps. First, by use of the properties
in Appendix 4, we see that
~D2 + 2XV
U{a'r)- L L 4X3
_D3/2tI/2
X2
(5-60)
Next
+Si«iwiK^) <«»
Then
It is easy to determine the exact ^equations for an (Rr T) model when the
humber_of units demanded in any time period is Poisson distributed and
the lead time is constant, since it is only necessary to set Q = 1 and R —
r + 1 in the results of Sec. 5-3. First, from (5-20), it is clear that
por = XTp(0; XT) + P(2; XT) - p(l; XT) + P(2; XT) = P( 1; XT) (5-64)
as expected. It is easy to evaluate E(l, R — 1, T), B{ 1, R — 1, T), since
in (5-21), (5-25), when Q = 1, it is unnecessary to sum over u from r + 1
to r + Q; it is only necessary to set u = r + 1 = R. Thus
Note that (5-68) holds for all T while (5-69) holds only if T is sufficiently
large. To determine the optimal T it is probably simplest to'tabulate X
optimized with respect to R as a function of T.
Let us next obtain the exact equations for the (R, T) model when R is
treated as continuous and the demand in any relevant time interval can
be assumed to be normally distributed with mean Xf and variance
To obtain the desired formulas from the results of Sec. 5-5, we must set
r = R and take the limit as Q -»0 rather than setting r + l = F, Q=1
as in the discrete case. For the continuous case it is clear that por should
be unity, since the probability that a positive quantity will be demanded
in any time period is unity. This is indeed the result obtained from (5-52)
when Q = 0, and the assumption is made that the normal density function
is essentially zero for non-positive arguments.
To compute the average number of backorders incurred per year and
the average unit years of shortage incurred per year, i.e., E{0, R, T),
B(0, R, T), note that for any function/(y) that
262 PERIODIC REVIEW MODELS WITH STOCHASTIC DEMANDS CHAP. 5
^p(V)dy-m) (5-70)
I
lim^ f(y) dy = lim m)-F(O) _dF
= >o = /(0) (5-71)
Q-»0 y jo Q—*0 Q
+ (S_Xr)4(^i)} (w
(5-72)
3C = | + 7c[t2-m-^] + xl?(0, R, T)
+ (IC + *)B(0, R, T) (5-74)
where L = J + A. The optimal value of R for a given T can be found by
solving the equation
- n - tp -u ^ ^
- 0 - 7C + + (7C + x) rs (5-75)
where
57? _ 1
572 “ 7
5E 1 VdU(R,r+ T) dU(R, tY
dR T (5-77)
17(72,2) = (2
—— g2XR/n *
2X2 ® (^)
SEC. 5-7 LIMIT OF <nQ, r, T) MODEL AS T-»0 263
5-7 The (Q, r) Model as the Limit as 7—>0 of the (nQ, r, 7) Model
The total cost expression (5-33) for the {nQ, r, T) model holds for all T > 0.
As T —»0, the average, annual cost approaches infinity if the review cost
J > 0. However, if the review costs are omitted, then the average annual
cost should approach a finite limit as T —»0. In fact, this limit should be
precisely the average annual cost for a (Q, r) model, and Q*(T), r*(T)
should approach Q*, r* for the {Q, r) model with the same values of the
parameters. We shall now prove this. Recall that in the derivation of
(5-33) it was assumed that the demand over any time interval was Poisson
distributed and that lead times were constant. Thus the limit as T—>0
should be a {Q, r) model with demand being Poisson distributed and con¬
stant lead times, i.e., the model developed in Sec. 4-7. The result also
holds true when the lead times must be considered to be random variables,
provided the approximations introduced in Chapter 4 that the lead times
can be simultaneously treated as independent random variables without
any crossing of orders are valid. The proof in this case is left for Problem
5-18.
The continuity of transition between (Q, r) models and {nQ, r, T) models
referred to above is important, because by including the review costs in the
{nQ, r, T) model, and adding the average annual cost of operating the
transactions reporting system to the average annual cost of the (Q, r)
model, it is theoretically possible to determine whether it is better to use a
transactions reporting system or a periodic review system with an nQ type
of operating doctrine.
Let us now show that as T —»0, X for the (nQ, r, T) model with demand
being Poisson distributed and constant lead times approaches X for the
(Q, r) model when the review costs are omitted. It follows from this that
Q*(T), r*(T), the optimal Q, r values for a given T, approach Q*, r* for the
(Q, r) model, since the review costs do not influence the determination of
Q* and r*.
First observe from (5-33) and (5-20)
Inasmuch as Q > 1
lim P(Q; XT) = lim |p(Q + 1; XT) = 0
T—>0 T-K) 1
264 PERIODIC REVIEW MODELS WITH STOCHASTIC DEMANDS CHAP. 5
since p(j; XT) approaches zero as T—>0 for all j except j = 0 (which is
excluded in the above summations). Thus
Si If = Q (5'79)
and the ordering costs approach \A/Q, which is the correct term for the
(Q, r) model.
Consider next
-j r+Q «
lim E(Q, r,T) = lirn ^ Z (x - u) {p[x; X(r + T)] - p(x; Xr)}
u = r+l x=u
However
= Z (y — tt)p(j/;Xr) + Z
2/ = w—1 j/=w—1
p(t/; Xr)
00
Therefore
i HlQ /*r+r 00
hm B(Q, r, !T) = lim Wy J] / Z ~ u)p(x; X?) d£
Thus
1 H-Q j*
Markov process discrete in space and time. Let us then set up the equations
whose solution yields the p(r + j). The inventory position of the system
immediately after a review can only have one of the R — r values
r + 1, r + 2,. . . , To compute the transition probabilities a#,
assume that the inventory position immediately after a review is
r + i, i = 1,. . . , R — r. We must compute the probability that it is
in state r + j immediately after the next review. If j > i and j < R — r
this probability is zero, since there is no way in which such a transition can
be made. If j < i, the probability is p(i — j; T). When j = R — r and
i < R — r, the probability is P(i; T), where P(x; T) is the complementary
cumulative of p(x; T), since any number of demands which reduce the
inventory position by the next review to or below r will result in having the
inventory position immediately after the next review be R. When
i = R — r, then j = R — r if there are no demands or if there are R — r
or more demands, i.e., the probability of this is p(0; T) + P(R ~ r; T).
f
Thus
TABLE 5-2
R R - l R - 2 R - 3 r+1
P(0; T) +
R p(l; T) P(2; T) p(3; T) p(R - r - 1; T)
P(R - r; T)
►-*
4
9(1; T) P(2; T)
1
Position
After R — 2 P(R - r - 2; T) 0 9(0; T) 9(1; T) p(R - r - 3; T)
Review
at Time t R — 3 P(R - r - 3; T) 0 0 9(0; T) p(R — r — 4; T)
r+ 1 P(l; T) 0 0 0 9(0; T)
i
SEC 5-8 MODELS OF THE (f?,r,T> TYPE 267
R-r
^R (5'85)
X) ^=1
/=i
In this section we shall derive the exact average annual cost expression for
an (R, r, T) model by computing the expected cost per cycle and then
multiplying by the average number of cycles per year, hirst, we shall
obtain the form of the average annual cost without specifying in detail the
nature of the stochastic process generating demands. The only assumptions
that will be made concerning the stochastic process generating demands is
that it does not change with time and that demands in different periods are
independent. We shall also assume that the lead times are constant. It
will not be possible to write down explicitly all the cost terms since no
special assumptions are being made about the nature of the demand dis¬
tribution. However, from this cost expression, we shall be able to obtain
the general solution to (5-85) for the p(r + j), which we were unable to do
in the previous section. After this we shall restrict our attention to the case
where a Poisson process generates demands and units are demanded one
at a time. In this case we shall obtain explicitly the average annual cost as
a function of R, r, and T.
When an Rr operating doctrine is used, an order need not be placed at
each review. The time between the placing of two successive orders (i.e.,
the length of a cycle) will always be an integral multiple of the time T
between reviews; however, the number of periods included in a cycle will
be a random variable. Hence the length of a cycle is also a random variable
which can only assume the values nT, n = 1, 2, 3, ... .
Let p(x; T) be the probability that x units will be demanded in one
period. Then, since the demands in different periods are assumed to be
independent pM(x; T) is the probability that precisely x units are
demanded in n periods, where pin)(x; T) is the n-fold convolution of p(x; T).
SEC 5-9 DERIVATION OF AVERAGE ANNUAL COST 269
There will always be at least one period per cycle and the contribution to
the expected cost from the first period is simply H(R, T), since immediately
after placing an order the inventory position is R. The expected cost per
cycle of carrying inventory and of backorders is then found by summing
the expected contributions for period n over all n > 1. Thus the expected
cost per cycle is
Having found the expected cost per cycle of carrying inventory and of
backorders, we must next compute the average number of cycles per
270 PERIODIC REVIEW MODELS WITH STOCHASTIC DEMANDS CHAP. 5
year. This is simply the reciprocal of the expected length of a cycle. Since
the length of a period is T, the expected length of a cycle is T times the
expected number of periods per cycle. Let us then compute the expected
number of periods per cycle. A cycle will be precisely one period if the
demand in the first period is greater than R ~ r. The probability of this
is P(R — r; T), where P(x; T) is the complementary cumulative of p(x; T).
Consider now the probability that a cycle will contain precisely n(n > 2)
periods. If R — r — j units have been demanded in the first n — 1 periods
after the previous order was placed, then the cycle will contain precisely
n periods if the demand in period n is for j or more units. Thus the prob¬
ability that a cycle contains precisely n periods is
J2p*-»(R-r-j;T)P(j;T)
y-i
E PM& ~ r ~ r>
p(r + j) - „ ~k'-T° ’
EE -r-y, T)P(j; T)
n= 1 y= 1
y = 1,. . . , B — r (5-94)
We shall now show that this is indeed true. All that is required is to show
that the p(r + j) of (5-94) satisfy (5-85). When the a# given in Table 5-2
are substituted into the equations p(r + j) = p(r + they
become
satisfy (5-95), (5-96) for any value of S. Consider first the equations (5-96).
When p(r + j) is given by (5-97)
E
i-j
p(r + i)p(i — 3) T) = 5 E E P(n)CR — r — i; T)p(i — j; T)
n = 0 i=j
= 5 E E
n—1 w=0
Pin~1}(R — r - j — u; T)p(w; T) = 5 E
n=X
— r —j;T)
The last step follows from the definition of yM(R — r - j\T) in terms of
p(n_1) and p (seep. 119). However, when j = 1,. . ., R — r — 1, according
to the definition used above p(0)(i? — r — j; T) = 0 so that
« R—t—1
1 + ^ r) = £ E
E
n—1
£
=1 i
P{n)(R - r - i; T)P(i; T)
SfftP<r!(0-;P,r ^ p(0>(x; ^ =
P( ,(0; T), it follows1 that x ** Now because p(0;T) =
= I] P(B,(0; T)
n— 1
Cleariy (5-99) does sum to unity since it is merely the probability that one
nW rIqUired t0 W R~r™te demanded. The analytic
proof can be obtained as follows y
However
.R—r—1 R—r—v — 1
S5
n-u »-u «-u
pM(v’T^u’T) = E E E Vw(i~u;T)p(u-,T)
ji = 0 i=o u=0
« fi-r-1 t
— E 1=1
2D + E E E P(n-1)(^ — «; T)p(u; T)
n=2 i=0 u=0
- S-r-1 - R-r
=E E 20 = E E P(n)(B — r — i\T)
E E »P<*"U(B - r - j; T) P(j- T)
n=iy=i
= E E (» + !) P(n)(R ~ r ~
71 = 0 /= 1
r, D P(J; D
= EEV”>(fi-r- y; 2D PC;; T)
n=o y=i
+ E E ^’(B ~ r - r, D P(j; T)
n=oy=i
Now
” K—T
* R—r j— l
~ Z Z Z nPM(.R - r - j; T) p(u; T)
n=l 3 — 1 u=Q
= Z Z PCn)(-K — r — j; T) + Z Z wP(n+1)(jR — r ~ j) T)
n=1 £1 7=i
00 R—r w
- ZZ - r - 3; T) = £ £ PW(B - r - i; T)
fr+T “
= T (r + j - M + b(r + j, T)
(5-103)
In this section we shall develop the analog of the (R, r> T) model studied
in the previous section for the case where R, r, and the demand variable
are treated as continuous.
276 PERIODIC REVIEW MODELS WITH STOCHASTIC DEMANDS CHAP. 5
Let v(x; T) be the density function for the demand in a period. Demands
in different periods will be assumed to be independent random variables.
Again we shall assume that the lead time r is a constant. If the inventory
position immediately after a review occurring at time t is r + x, let
H(r + x, T) be the expected costs of carrying inventory and of backorders
incurred from t + r to t + r+T. Suppose that an order is placed at
time t0. The probability that at time to -f- uT) u > 1, the inventory posi¬
tion of the system lies between x and x + dx given that no order has been
placed since time to is v(n)(R — r — x; T)dx, where vin)(x; T) is the 71-fold
convolution of v(x; T). Thus the expected cost per cycle is
X(R, r, T) =
00 rR—r
(5-108)
v{R) =
(5-111)
E vW(R -r-y,T)
p(r + y) =
[e
Ln = 2
r
JO
nv(n~l)(R -r-x; T)V(x; T) dx +V(R-r;'.
We shall leave the proof that (5-111) and (5-112) are indeed the solutions
to (5-109) and (5-110) to Problem 5-23. For any given distribution v(x; T),
it is generally very difficult by use of (5-111) and (5-112) to obtain p(R)
and p(r + y) in a simple form not involving a sum over n.
When one deals with the inventory position at a review time prior to the
placing of any order the peculiar behavior at r + x — R is not encountered.
If 6{x) is the density function for the inventory position at a review time
prior to placing any order, then it is clear, on considering the possible
transitions from one review time to the next, that 6{x) must satisfy the
equations.
6(x) = v(R — X) T)
(5-113)
6{x) = v(R — x; T)
(5-114)
v(x; T) = fie*11*
Note that we do not implicitly indicate the time dependence here. We
imagine that T is fixed.
We shall first show how (5-113), (5-114) can be solve explicitly for B(x).
Note that
dv
— = -0* 6T* = -pv (5-116)
/
In order that 0(x) be a legitimate density function, it is necessary that
R
6(x) dx = 1 = e$r(R — r) + ki j
r
e$x dx = ki e$r ^ — r ^J
SEC. 5-10 THE <R, r, T) MODEL WHEN DEMAND IS CONTINUOUS 279
R — r +
and hence
r < x < R
•R — r +
(5-121)
\R-r + ■
6(x) dx
(5-122)
p(r + y) 0 < y < R — r
R-r +
Let us now compute p(r + y) and p(R) from (5-111), (5-112) to show
that the same result is obtained. From (3-108), it follows that if v(x; T) =
pe~fixj then v{n)(x] T) is a gamma distribution of order n, i.e.,
V(x; T) = e-e*
00 rl
Z
n-2 JO
-r-x; T)V(x; T) dx + V(R - r; T)
Finally
00
= 0 13 - »• - z)] = /3
n=0
Hence
which is precisely (5-122). Thus, we have obtained the same result in two
different ways. To solve (5-113), (5-114) explicitly for 8(x) is considerably
more difficult when v(x; T) is a gamma density of order two. We leave this
case to be worked out in Problem 5-32. The procedure for solving (5-113),
(5-114) when v(x; T) is gamma distributed by reducing them to differential
equations has been presented by Karlin m Chapter 14 of [1],
To have obtained simple expressions for p(r + y) and p(R) when the
distribution of demand for a period is exponential is not especially valuable
from a practical point of view for several reasons. First, the exponential
distribution is seldom a realistic representation of the density for the de¬
mand in a period. Secondly, in order to compute H(x, T) it would be
necessary to use some other distribution such as the normal since H(x, T)
involves the demand from r to r + T, and the exponential distribution
does not provide for introducing time explicitly. Finally, the average
annual cost even in this case does not reduce to an expression which can be
easily handled manually. While the result just obtained does not seem to be
of great practical value, it is of interest from a theoretical point of view to
show how p(R), and p(r + y) may be computed in two entirely different
ways.
When review costs are ignored, a transactions reporting system always has
lower average annual costs, all other things being equal, than a periodic
review system. The principal reason for this is that in the periodic review
system, sufficient safety stock must be held to offer protection for a length
of time t + T, while for a transactions reporting system r is the relevant
length of time. Thus the periodic review system will require higher safety
stocks and have higher costs. The thing that really determines which sys-
282 PERIODIC REVIEW MODELS WITH STOCHASTIC DEMANDS CHAP. 5
Except for the simple, approximate model of Sec. 5-2, all the models in this
chapter have concentrated on the backorders case. We noted in Chapter 4
that it was not possible to treat rigorously the lost sales case when more
than a single order was outstanding. For periodic review systems, things
are even more difficult, and trouble is encountered even in treating the case
where only a single order is outstanding.
To illustrate the difficulties, consider the simplest case, i.e., that where
the number of units demanded in any time period is Poisson distributed,
t e procurement lead time is a constant t, and an order up to R policy is
used. The additional stipulation will be imposed that there is never more
than a single order outstanding. This will be rigorously true if r < T.
In order to compute the average annual carrying costs and stockout costs,
it is necessary to determine the distribution of the on hand inventory, either
at a review time or immediately after an order arrives.
Let 6(x) be the probability that x units are on hand at a review time.
If x units are on hand, then R — x units are ordered, so that the quantity
on hand plus on order immediately after the order is placed is R. To set up
the equations which determine 6(x), it is convenient to divide the time
between two successive reviews into two sub-periods, the first extending
over a time period of length r from the time that the order is placed until
SEC 5-13 THE LOST SALES CASE 283
it arrives, and the other extending over a time period of length T — r from
the time that the order arrives until the next order is placed. It should be
noted that it is possible to incur lost sales in both of these sub-periods.
Let 4f(z) be the probability that z units are on hand immediately after an
order arrives. The time after an order arrives until the next review is
T — t. The probability that x units are on hand at the next review, if z
units are on hand immediately after an order arrives, is
m = E *(*)p
2=0
X(-T - ^ (5-127)
Next, by considering the time period extending from 0 to r, we shall
express \f/(z) explicitly in terms of Q{x). This can then be substituted into
the above equations to yield a set of equations involving Q{x) only (or
involving ip{z) only). If z units are on hand after an order arrives at time r,
the quantity on hand x at the time the order was placed could not have been
less than R — z> since the size of the order is R — x. If the quantity on
hand at the review time satisfies x > R — z, then in order to have z units
on hand after the order arrives, it is necessary that the lead time demand
be R — z. If x — R — z, then for any lead time demand > R — z, the
quantity on hand after the arrival of the order must be z. Thus we have
R
}p(z) = p(R — z; Xr) ^2 z; Xr)6(R — z),
x=*R—2+1
z > 0 (5-128)
o(x) =E
z~x i
~ Xt)p[z ~ x> x(r - T)1 E y—R—z-\-l
e^>
+E
z=x V
_ z> Xt)pCz - x> x(r - r)]0CS ~ z) * > 0 (5-130)
284 PERIODIC REVIEW MODELS WITH STOCHASTIC DEMANDS CHAP. 5
*(°) = 2=1
Z W ~ 2; Xr)P[2; \(T - r)]
^
J
y = R~z+l
%)}
J
+ Z0 \p(R
z= 1
~ z'> Xr)P[z; x(T — t)]8(r — z) (5-131)
“(*• y) = Z 2 =2/4-1
P(R - z; XtMz - x; \(T - r)] (5-134)
If the values of 0(2;) from (5-126), (5-127) are substituted into (5-128),
(5-129) we obtain the set of equations which involve ^(2) only. They are
R R
Hz) = P(R - z; Xr) J2 Hy)p\y - x; \(T - r)]
x = R — s-f-l 2/ = a:
R
+ P(P - 2; Xr) *(j/)p[y - P + z; X(T - r)],
y — R—z
z = 1,. . ., R - 1 (5-138)
and the average unit years of storage incurred per year would be
D(R, T) =
1
X Hz)
2=0
I T X (* - *)?(*; Xi) dt
XwsO
(5-143)
The average annual cost when the unit cost of the item is constant is then
We have now seen how to formulate the model if the distribution ^(z) was
available. Unfortunately, as we have noted, it does not seem possible to
solve for the tl/(z) explicitly, and thus the terms D(R, T) and E(R, T)
cannot be evaluated explicitly. Hence, for periodic review systems, it is
not possible to formulate exact models for the lost sales case even when
the requirement is made that never more than a single order is outstanding.
About the best one can do is to use the simple approximate treatment of
lost sales given in Sec. 5-2. Fortunately, for many practical applications,
this is adequate.
It has already been noted in Sec. 5-2 that for periodic review models, unlike
(Q, r) models, it is possible to treat rigorously stochastic lead times when
more than a single order can be outstanding at any point in time provided
that the range of lead time values is less than T. In this case, there can
286 PERIODIC REVIEW MODELS WITH STOCHASTIC DEMANDS CHAP. 5
be no overlap of the lead time densities, and one can rigorously require that
lead times be independent random variables and also that orders be re¬
ceived in the same sequence in which they were placed. The reason that
stochastic lead times can be treated rigorously in certain cases for periodic
review systems, while this is not possible with (Q, r) models, lies in the fact
that for periodic review systems, the time between the placement of orders
must be at least T, while for (Q, r) models there is a positive probability
that two orders will be placed in an arbitrarily small time interval.
Let us show how to incorporate stochastic lead times into the <nQ, r, T)
model of Sec. 5-3. The notation and terminology of that section will be
employed here. We shall assume that the lead times are independent ran¬
dom variables and that orders are received in the same sequence in which
they were placed. This can be done rigorously if the range of possible lead
time values is less than T\ otherwise, the treatment is only approximate.
If a review takes place at time t and an order is placed, and if an order is
also placed at time t + T, then let tj, t2 be the random variables repre¬
senting the lead times of the orders placed at t and t + T respectively.
The lead time density function will be written g(r). Thus the expected
number of backorders incurred in the period between the arrival of the two
orders under consideration is
1 f* /*» **+Q « f n
Q L L 23 23
Jv u=r+li = u
— w) i V[z; k(r2 + T)] — p(x; Xn) f g(.T2)g(Tj) dr2 dn
'v J
r+Q * r >
7) 23 23 (x ~ w) ]
^u = r+li=u 1
T) — h(x) y
J
(5-145)
where
The limits on the integrals have been taken to be 0 to oo. Of course, g{r)
may be 0 over most of this range so that the limits could be replaced by
'Tmin and rmax. Clearly, the assumption that lead times are independent and
arrive in the order placed could not hold rigorously if g(r) > 0 for all r
between 0 and co. In particular, the assumption could not hold rigorously
if g(r) was represented by a gamma distribution. In this case, the treatment
would at best be approximate.
The expected value of the unit years of shortage incurred between the
arrival of the two orders is
SEC 5-14 STOCHASTIC LEAD TIMES 287
1 f00 f 00 fT2*^” ^ Q 80
q / / £ X) (* - Wg(Ti)g(r2) d£ dn dn (5-148)
''■'JO Jo Jn
-j r+Q 00
m, r, T) = n £ £(*-«) [h(x; T) - h(x)] (5-149)
u=r+lx=u
r-f-Q so
1 C80 f00 f T2~^* ^
B(Q, r, T)
QTJo Jo L £ £
u=rJrl x = u
These changes allow the inclusion of stochastic lead times in the model
developed previously.
It is only for very special forms for g{r) that it is at all simple to work out
explicitly E(Q, r, T) and B(Q, r, T). This can be done quite easily if r is
gamma distributed. Problem 5-27 asks the reader to do this. Of course, for
a gamma distribution, it will never be rigorously true that the range of
lead time values will be less than T. When the demand variable is treated
as continuous and normally distributed, the usual procedure used to take
account of stochastic lead times is simply to change the variance of the
normal distribution, i.e., the value of D. In this case, the value of D for the
t terms need not be the same as the value of D in the r + T terms.
Stochastic lead times can be handled in (E, r, T) models in precisely the
same way as for the (nQ, r, T) model illustrated above. The details are
left for Problem 5-28.
* The reader should be careful to note that the use of this artifice in no way changes
the fact that an ordering cost is incurred only if a positive quantity is ordered.
288 PERIODIC REVIEW MODELS WITH STOCHASTIC DEMANDS CHAP. 5
When quantity discounts are offered which depend on the order size an
Rr or order up to R policy need not he optimal. To operate the system
optimally, a function Q(%) would have to he computed, using the techniques
of dynamic programming discussed in Chapters 7 and X, which gives the
quantity Q to order as a function of the inventory position £ at the review
time. In the real world, one will often use an order up to It, Hr or nQ policy
even when quantity discounts which depend on the order size are available
since it is too complicated to attempt to compute and use the Q(£) function!
The quantity discounts may, however, influence the optimal values of
R }t and 1 *. I he main e fleet of quantity discounts tin a periodic review
system, if any, is to increase 7’*. We shall illustrate how to account for
quantity discounts in an (It, T) model.
Let us study the backorders case under the assumption that the number
of units demanded in any time period is Poisson distributed. Ltd- (\x)
be the cost of the units when x are procured. The probability that, x units
are procured at any given review is the probability that x units wore de¬
manded in the past period, i.e., p(x; k'l). 'I hus the expected cost, of the
units procured at any given review is
J2 C(x)p(x; XT)
XmQ
Let us now evaluate this expression explicitly for the case of all
units quantity discounts. Assume that, there are m price breaks
?ij • • ■ > $>»(<?o = 0, jm+i = °o), and that the unit. cost, is (\ if the quantify
ordered x satisfies q{ < x < <•/,„. Then, the average annual cost of the
units procured is
PROBLEMS 289
The expected unit cost V of the item is simply the average annual cost
divided by the mean rate of demand, i.e.,
7ft
P - £ - 1 ;XT)- P(qt - 1; XT)] + CJP(qm - 1; XT)
i— 1
It is this expected unit cost that must be used in the inventory carrying
cost term. Note that R* for a given T can be found without including the
M(T) term in the cost expression. However, the expected unit cost V will
now be a function of T. To determine T* it is necessary to include the
M(T) term in the cost expression.
The computational procedure in the lost sales case is precisely the same
as in the backorders case, provided that the influence of quantity discounts
on the cost of a lost sale can be ignored. This is usually a valid approxima¬
tion in practice.
We shall leave for Problem 5-29 the development of the average annual
cost of units procured when incremental quantity discounts are available.
The same general computational procedure applies in this case. However,
(5-152) must be replaced by an equation of somewhat different form.
It is straightforward but tedious to compute M and (J for nQ and Rr
operating doctrines also. We ask the reader to attempt this in Problem
5-30.
REFERENCES
PROBLEMS
cost of placing an order is $2.00 and the cost of the review apportioned
to this item is $15. Demands occurring when the system is out of
stock are backordered and the cost of a backorder is taken to be
$100. The system uses an inventory carrying charge of I = 0.20.
The number of units demanded in any time period is normally dis¬
tributed with mean 240* and variance 500*, where * is the length of
the period in years. The lead time is constant at 2.5 months. What
is the optimal value of R for the presently used review period? What
is the optimal review interval and R*(T*)7 What average annual
savings could be obtained by using T* as the time between reviews?
6-2 Derive the discrete analogue of the model developed in Sec. 5-2 for
the case where demand is Poisson distributed. Consider both the
case of constant and stochastic lead times. What is the inequality
which is used to determine R*(T). Consider both the backorders and
lost sales cases.
6-3 A certain item carried in a large military supply system is reviewed
every three months. An order up to R policy is used. The number of
units demanded in any time period can be considered to be Poisson
distributed with the average annual demand being 15. The procure¬
ment lead time is essentially constant with the value five months.
The unit cost of the item is $500, the cost of a backorder is taken to
be $5000, and the cost of placing an order is $30. The system uses an
inventory carrying charge of 7 = 0.20. Determine the optimal value
of R for the specified time between reviews. What is the safety stock?
What is the cost of uncertainty?
6-4 Under what conditions will the solution to Eq. (5-9) not be unique?
Hint The solution is unique if h(x; T) >0 for all x > 0.
6-6 For the model developed in Sec. 5-2 develop the derivatives needed
for Newton’s method to solve for R* and T* simultaneously.
6-6 Apply the results of Problem 5-5 to the example given in Sec. 5-2 and
determine R* and T* simultaneously.
6-7 Discuss how the method of steepest descents could be used to deter¬
mine R* and T* for the model developed in Sec. 5-2.
6-8 Use the method of steepest descents to determine R* and T* for the
example presented in Sec. 5-2.
6-9 Prove that the cost function of Eq. (5-6) is a convex function of R
for a given T. Under what conditions is it strictly convex? Is 3C a
convex function of R and 77
6-10 Discuss the behavior of the safety stock for the model developed in
Sec. 5-2 as 7 —> 0. How do you explain this?
PROBLEMS 291
5-11 A department store reviews the stock of a particular white shirt each
week. An order up to R policy is used. The number of units
demanded in any time period for a particular size of the shirt can
be considered to be Poisson distributed with a mean of 25 per week.
Each shirt costs the store $3.00 and the selling price is $5.95. An
inventory carrying charge of I = 0.17 is used. Demands occurring
when the store is out of stock are lost, and the cost of a lost sale aside
from lost profit is $10.00. The procurement lead time can be assumed
to be constant at 10 days. Determine the optimal value of R.
5-12 Go through the same sort of analysis employed in Sec. 4-6 for the
(Q, r) model to obtain the various time averages for the {nQ) r, T)
and (R, r, T) models, and show their relation to expected values. In
particular, show that the average annual cost of carrying inventory,
stockouts, etc. is the expected cost per period times 1/T. Also show
that the average annual cost is the expected cost per cycle times the
average number of cycles per year.
5- 13 Derive Eqs. (5-27) through (5-29).
6- 14 For the (nQ, r, T) model compute the steady state probability $(x)
that the inventory position at a review time is x (before the place¬
ment of any order). Do this by making use of the p(r + j). Also
develop the set of equations from which the \f/(x) can be computed
directly by considering the related Markov process for generating
transitions between the inventory position states at the beginning of
a review. Can you solve these equations directly?
5- 15 Compute the mean and variance of the density \j/(x) defined in
Problem 5-14.
6- 16 For the continuous version of the (I?, T) model presented in Sec. 5-6
determine the derivatives that would be needed if Newton's method
were to be used to determine R* and T* simultaneously.
6-17 Is the cost expression (5-74) a convex function of R for a given T?
Is it a convex function of R and T then. What are the implications
of this result for the determination of and consequences of R* and T*?
6-18 Carry through the limiting processes of Sec. 5-7 for the case of sto¬
chastic lead times. What approximations and/or assumptions are
needed?
6-19 Provide a step-by-step derivation of Eqs. (5-88), (5-89).
5-20 For a given value of r and R, develop a numerical procedure for
solving the set of equations (5-88), (5-89).
5-21 For a periodic review system determine an explicit form for the aver¬
age annual cost expression in the backorders case when an Rr policy
292 PERIODIC REVIEW MODELS WITH STOCHASTIC DEMANDS CHAP. 5
is used and the demand in the time between reviews has an exponen¬
tial distribution. Assume that the lead time is precisely one period
in length, x = 0, and that the inventory carrying cost for a period
depends only on the on hand inventory at the end of the period.
Determine R* and r*.
6-22 In Eq. (5-108) work out explicitly H(x, T) and v(n)(x; T) when v(x; T)
is a normal distribution.
6-23 Show that Eqs. (5-111), (5-112) are solutions to Eqs. (5-109), (5-110).
Also show that with the proper interpretation, p{R) and p(r + y)
represent a legitimate density function.
6-24 Derive in detail Eqs. (5-113), (5-114).
6-25 Evaluate explicitly the /3(z, y) in Eq. (5-141).
6-26 Develop a numerical procedure for solving the set of equations (5-132)
and (5-141) for a given value of R.
6-27 Work out the explicit form for Eqs. (5-149), (5-150) when the lead
time is gamma distributed.
6-28 Discuss the introduction of stochastic lead times into (R, r, T) models.
6-29 Develop the equations needed to handle incremental quantity dis¬
counts in order up to R models when demand is Poisson distributed.
5- 30 Study the problem of handling quantity discounts in (nQ, r, T) and
{R, r, T) models under the assumption that an nQ or Rr policy is to be
used.
6- 31 Show that the equations for quantity discounts developed in Sec. 5-15
reduce to the correct values when the unit cost of the item is a con¬
stant independent of the quantity ordered.
5-32 Obtain 6(x) and p(r + y), p(R) by solving Eqs. (5-113), (5-114), in the
case where v(x; T) is a gamma distribution of order two. Also obtain
p(r ”t~ y) and p(R) from (5-111) and (5-112). Hint: Use the proper¬
ties of v(x; T) to obtain second order differential equations for 6(x).
These equations are
5-47 For the simple model discussed in Sec. 5-2, show how to include an
inventory carrying cost which is based on the maximum inventory
level. How does this term change the equation which determines J?*?
PROBLEMS 295
PERIOD MODELS
Publilius Syrus
6-1 Introduction
The models to be discussed in this chapter are perhaps the simplest of all
inventory models in which demand is treated as a stochastic variable. The
essential characteristic of these models is that only a single time period,
usually of finite length, is relevant and only a single procurement is made.
A rather wide variety of real world inventory problems including the
stocking of spare parts, perishable items, style goods, and special season
items offer practical examples of the sort of situations to be studied here.
Since only a single time period is to be considered, there is no steady state
associated with the models analyzed in this chapter. These models then
provide a transition between the steady state models of Chapters 2, 4, 5
and the dynamic models to be presented in the next chapter.
* We here ignore the fact that in reality he will stock trees of different sizes and costs,
and that the demand variables for trees of different sizes will not be independent, since
if he is out of one size a customer will often accept a different size.
297
298 SINGLE PERIOD MODELS CHAP. 6
xp{x) + Sh Z p(x)
x=0 x = A-f 1
h-1
where p, is the expected demand for the period. The smallest h which max¬
imizes the expected profit is the largest h for which Ag(A) > 0. However
v>ru\ ^ ^ C — L
W>S + To-L- (S -C) + X„ + (C - L) M
is optimal, i.e., is h*. Note that C — L is the loss per unit on any units re¬
maining at the end of the period, S - C is the unit profit, and (S - C) + x0
is the cost incurred if a demand occurs when the system is out of stock.
Note also that no restrictions whatever are placed on the demand distribu¬
tion for the period p(x). For example, units need not be demanded one at
a time, and demand in non-overlapping time intervals need not be inde¬
pendent. If AS(fi* + 1) = 0, both h* and h* + 1 are optimal.
Often 4 is convenient to treat h and the demand variable as continuous.
SEC 6-3 EXAMPLES 299
Then if f(x) is the density function for demand in the period and F(x) is
its complementary cumulative, the expected profit for the period when h
units are procured is
+L J (h — x) f(x) dx — Ch — to J (x — h) f(x) dx
m s^~tL
=
In Problem 6-4 we ask the reader to show that (6-5) is a concave function
of h. Furthermore, 2(h) is usually strictly concave. These results imply
that any relative maximum of g is the absolute maximum. When g is
strictly concave, then the absolute maximum is unique. Problem 6-5 asks
for a discussion of the conditions under which the solution to (6-6) is not
unique.
6-3 Examples
Several examples will now be given to illustrate the wide variety of prob¬
lems to which the simple model developed in the previous section is
applicable.
1. A large supermarket must decide how much bread to purchase each
day. For shopping days Monday through Thursday, past history has
shown that the daily demand can be considered to be normally distributed
with mean 300 and standard deviation 50. A loaf of bread sells for $0.25
and costs the store $0.19. Any bread not sold by the end of the day is, on
the following day, placed on a counter where it is sold for $0.15 per loaf.
All bread on this counter can be sold at this price. It is desired to deter¬
mine the optimal number of loaves to purchase to maximize the expected
daily profit.
In the notation of the previous section, S = $0.25, C = $0.19, L = $0.15,
x0 = 0. Then h* is the solution to
300 SINGLE PERIOD MODELS CHAP. 6
h - 300
0.253 or h = 300 + 12.65 « 313
50
Now the expected number of loaves which will be placed on the stale bread
SEC 6-3 EXAMPLES 301
counter is 76.48. The expected daily profits using h* and /z are respectively
In this case there is a very large difference between the expected profit
when A* is used and the expected profit when is used. It should be
observed that in this example, the expected daily profit is not what would
be obtained by averaging the profit obtained from the accounting records
over a long time, since it includes the influence of the goodwill loss wrhich
would not appear in the accounting records. Indeed, on the basis of what
would be taken from the accounting records, the expected profit wrould be
lower using A* = 375 than using h = 313 or h = m- The reader should
check this by computing the expected profit with 7r0 = 0 when A* = 375.
3. A fashionable candy store does not make its own chocolates, but
instead buys from a large producer which caters to a number of different
stores. The store under consideration must decide how many large choco¬
late rabbits to order for the Easter season. They must be ordered two
months in advance and there is no possibility of placing a reorder. Each
rabbit costs the store $2.50 and sells for $7.00. Any rabbits not sold at the
end of the season are a total loss. The store feels sure that it can sell at
least 100 of these rabbits and not more than 500. Any number between
100 and 500 is felt to be equally likely, i.e., the store feels that the dis¬
tribution of demand is uniform over the interval 100 to 500. If the demand
variable is treated as continuous, this means that the density function
f(x) is
fix) = 100 < x < 500; fix) = 0, otherwise
Then, since h* = 357 and jjl = 300, the expected number of rabbits re¬
maining when h* are stocked is
The expected profit on the rabbits if A (100 < A < 500) units are stocked
is
8(A) = (S — L) p. — (C — L) h
Thus by carrying 57 rabbits above the mean demand, the expected profit
is increased by $28.57.
4. Certain spare parts for aircraft are made at the same time the aircraft
is produced. Once production on the aircraft has ceased, it is extremely
difficult to obtain these spare parts. Production is about to cease on a
certain military aircraft. It is desired to determine how many spares of a
certain low demand item should be available at the time production stops.
For the number of planes in the force, past data indicate that the demand
for the part will be Poisson distributed with the mean rate of demand
being 0.75 per year. It is uncertain precisely what the operational life of
the aircraft will be. However, it has been decided that the probability
density for the time until obsolescence can be described by a gamma dis¬
tribution with a mean of 6 years and a standard deviation of 1.5 years.
The parts cost $2000 each and their scrap value, if any remain unused at
the end of the plane's operational life, is $100 each. If a demand should
occur when no spares are available, the cost of obtaining one is estimated
to be $13,000.
In order to determine the optimal number of spares which should be on
hand, it is first necessary to determine the marginal distribution of demand
over the operational life of the aircraft. We know from (3-73) that this
distribution is the negative binomial a + 1, 0/(0 + X)], since de-
SEC 6-3 EXAMPLES 303
mands are Poisson distributed and the operational life has a gamma dis¬
tribution. To determine a, 0 we use (3-44), (3-45) which relate the mean
and standard deviation of the gamma distribution to a, p. Thus
00 1QOO
E bN(x; 16, 0.781) > = 0.1475
It is not too difficult to compute the optimal h, since once b&(x; 16, 0.781)
has been calculated for a given x, it is easy to compute successively values
for other x values using
Ifi 4- r
bN(x + 1; 16, 0.781) = (°*219) b*(x'>16> °*781)
subject to the constraint that the total volume of the spares does not
exceed F, i.e.,
n
Let us now consider how the h* may be determined. First, imagine that
each h* will be large enough that the hi can be treated as continuous.
Assume that p%(x) can be approximated by the continuous density fi(x),
whose complementary cumulative function is Fi(x). It is clear that when
we treat the hi as continuous, then the h* will satisfy (6-8) as a strict
equality. To solve the problem we can use the Lagrange multiplier
approach. We introduce a Lagrange multiplier 6 and form the function
[Sp<(1)]- Sp,(2)}
The next unit is assigned to the index where the maximum is taken on, etc.
This is continued until adding an additional unit would exceed the volume
restriction. We ask the reader to explain in Problem 6-28 why this pro¬
cedure should often yield a satisfactory computational procedure and under
what conditions it might not be expected to work too well. The problem
of finding non-negative integers which maximize (6-7) subject to (6-8) can
be solved exactly by the technique of dynamic programming to be intro¬
duced in the next chapter. We shall illustrate the procedure there. It
might be pointed out that even dynamic programming cannot give the
306 SINGLE PERIOD MODELS CHAP. 6
exact solution to the actual physical problem. The reason for this is that
specification of the unit volumes Vi and V is not sufficient. It is also
necessary to consider the actual shape of the units and of the space avail¬
able for storage. In other words it does not follow that simply because
(6-8) tells us that an additional 2 cubic feet of storage space are available
that an item having a volume of 1.75 cubic feet can be fitted into this
space. It depends on the shape of the available space and of the item.
In some cases there may be two constraints on the problem such as a
volume and weight constraint. The addition of one more constraint makes
the problem much more difficult to solve, but again the dynamic pro¬
gramming technique can be used.
Single period models involving a number of items connected by one or
more constraints can take a variety of forms. The following example
presents a situation which is slightly different from that given above, but
the method of analysis is just the same.
and the necessary conditions which the h* must satisfy (if they are all
positive) are
aw dQijhi)
= 0 = + 8C i
dhi dhi
P \ Ci - U -dCi . - n
Fi(hi) = g' i = 1, 2, 3
Using the appropriate values for the individual items, we see that the
necessary conditions become
In this problem, it does not follow that it will be optimal to use the
entire budget. However, one sees immediately from the coffee maker
equation that if there was no budget restriction (i.e., 6 ~ 0) an infinite
number of coffee makers would be purchased (why?); Thus 0 must be
different from 0. In particular, it is clear that 0 must be negative^ Table
6-1 presents the results of the numerical computations. To the accuracy
with which rounding to integral values permits the computation to be
made, one might select either 8 = —0.99 with h± = 42, h2 = 123, hz = 95
or 0 = —0.98 with hi — 43, h2 = 124, hz = 95. There would be no practical
value in attempting to be more precise.
TABLE 6-1
Cost of units
-8 hi h2 h (dollars)
The single period model introduced in Sec. 6-2 did not involve any costs
that were time dependent. We would now like to generalize that model to
include carrying costs which are proportional to the length of time that a
unit remains in inventory, and a stockout cost which is proportional to the
length of time from the point when the demand occurs until the end of the
period. As usual, the cost per unit year of keeping the item in stock will
be written IC, where C is the unit cost of the item. Furthermore, as usual,
7r will be the cost per unit year of a stockout. When including these costs,
one is forced to be somewhat less general than in Sec. 6-2, since it is neces¬
sary to introduce a distribution for the demand from the beginning of the
period to any time t in the period. The only case that can be worked out
easily when the variables are treated as discrete is that where demand is
Poisson distributed over any time period. This is the case we shall examine.
Consider first the case where the mean rate of demand X is constant over
time, and the length of the period is fixed and is not a random variable.
Then if h units are on hand at the beginning of the period, the expected on
hand inventory at time t is from A3-11
h
^2 (h — x)p(x-j \t) = h — \t + \tP(h; \t) — hP(h + l;\f) (6-12)
*=o
and the expected inventory carrying charges for the period are
by A3-18, A3-19.
If at time t the demand has exceeded h by y, then between t and t + dt
the variable stockout cost incurred is Tty dt. The expected value of y at
time t is by A3-10, A3-13
and the expected time dependent stockout cost for the period is ttB(K)
where
the Poisson terms in (6-16) the negative binomial distribution will appear.
We leave for Problem 6-13 the derivation of the equivalent of (6-17)
when the length of the period is gamma distributed. Frequently, it is
difficult to estimate reliably what the nature of a continuous density
function for the length of the period should look like. About, the best that
can be done is to select n times T, at which the period may end, and assign
probabilities wj that the period will end at times Tj. Then if t^ft, Tj) is the
expected profit when h units are on hand at the beginning'of the. period
and the length of the period is Th the expected cost- (.averaged over the
possible period lengths) is
is positive. Kadi 7;) has the farm (0-17) (with T replaced by Tj).
Of course, the determination of h* Ls now more complicated than in the
determination of h* using (1M7).
One final geneialization ol the model is ot ten needed. The mean rate of
demand will now be allowed to vary with time. Imagine a curve Ls avail¬
able giving the cumulative moan demand IHt) irom the beginning of the
period until time t. It will be assumed that the eumillative demand up to
time t is Poisson distributed with mean Then write hit) - t\{t).
If the rate oi demand remained constant at the value. XU) from the begin¬
ning of the period until time /, than there would be the mum* probability
of having x units demanded by time4 / as in the ease when the mean rate of
demand changes with time and the expected mean demand up to time, t Ls
D(t).
Let us first consider the ease where the mean rate of demand ean vary
with time, but the length of the period is fixed at the value T. To compute
the optimal quantity to have on hand at the beginning of the period, it is
convenient to subdivide the period into m subperiods, subperiod i extending
from time to U (4 (), tm =- 7*). Within each subperiod / the mean
rate of demand will be assumed constant at the value A, - Aitt). By
choosing U — 1 sufficiently small, the approximation made by taking the
and
u, Xi) — a(h, ti-1, X*)] (6-21)
where
These results follow from (6-13) and (6-14) on breaking the integral from
U-i to U into the integral from 0 to U minus the integral from 0 to U~i.
The expected profit for the period is then
1 m
8(A) = (-S - i)M - (C - L - ICT)h X<(i? - <?-i)
1=1
The optimal h is then the largest h which yields a positive value for the
following expression for Ag(ft).
(6-24)
m
- (t + IC)J2 [<iP(A; Xifi) - ii_iP(A; Xiii-0]
1=1
m ^
- (# + IC)h V 7- [p(X; Xjii) — p(A; X,-i,-_i)]
i=i Xi
Again the optimal h can be found by tabulation, although the task could
be quite arduous to carry out by hand. Several of the expressions in (6-24)
such as P(h; X,i») — P(h; XA-i) involve the difference of two terms which
are almost equal. In Problem 6-14 we ask the reader to show how Taylor's
expansion can be used to simplify such terms and thus yield a considerably
simpler expression for Ag(h).
A variable mean rate of demand can be combined with a situation in
which the length of the period is a discrete random variable having n
312 SINGLE PERIOD MODELS CHAP. 6
possible values Th with the probability of T} being Uj. For each Ts there
will in general be a different function describing the mean rate of demand
over time. We can then obtain expressions like (6-23), (6-24) for each possi¬
ble Tj. Let g(h, Tj) be the expected profit if the length of the period is Tj
To obtain g(A, Tj) from (6-23), T is replaced by Th X, by Xiy, and m by m,
(for each j we may wish to use a different number of subperiods). Then
n
The task of determining the optimal h from (6-26) will usually involve too
much work to be done manually. However, it can be done quite easily on
a digital computer. It is not easy to allow the mean rate of demand to be
time dependent if the length of the period is described by a continuous
density function. We ask the reader in Problem 6-15 to explain why this
Models of the type presented in this section have been discussed in
[4, 5],
the additional unit is demanded (i.e., the stockout cost is avoided on this
unit) and 0 otherwise. Hence the expected change in stockout cost is
— 7t0jP(/i) . Therefore
which is precisely (6-3). We have thus obtained &Q(h) without ever having
written down the expected cost.
What we have introduced above is, of course, nothing but the marginal
analysis so familiar to economists. Marginal analysis is a very useful tool
for deriving optimizing conditions, and is applicable to a very wide
variety of problems. In particular, marginal analysis can be used to obtain
the optimizing conditions for most of the models discussed in the previous
sections. The essential feature of marginal analysis is that one examines
what happens when a unit change is made in the variable under consider¬
ation. As another simple example, consider the order up to R model for
the backorders case discussed in Sec. 5-2. Observe that A3C(i?) is equal to
the expected change in carrying costs plus the expected change in backorder
costs on changing from R — 1 to R. Now when the time over which
backorders exist is negligible, then adding an additional unit implies that
it will be carried in inventory essentially all the time, and the expected
change in annual carrying costs will be IC. The change per period in the
backorder cost is — x if the additional unit is demanded in the period
and 0 otherwise. Thus, per period, the change in expected backorder cost
is —irH(R; T), and per year is —irH(R; T)/T. Therefore
l ~ — A ,4
XA _Q Q- i_ Q{Q - l)
and, since the average inventory is (Q - l)/2, the average annual change
in carrying costs is /C/2. Summation of these two terms yields the proper
expression for A3C(Q). In this case, however, marginal analysis did little
to simplify the analysis, since one essentially had to have the cos), terms
to difference them. In essence, one could not. analyze the effects of
adding one unit without considering its effects on the costs associated with
the other Q — 1 units.
The computation of ArIl{Q, r) for the model of Sec. -1-7 by marginal
analysis provides an illustration of a. ease when' it is easy to handle a time
dependent cost. Consider ArIi((J, ;•). If the inventory position of the sys¬
tem is j units above the reorder point at time / r, then in going from
r - 1 to r, the expected number of backorders at time t will be reduced
by one if the extra unit is demanded and will be unchanged otherwise.
The probability that the extra unit is demanded is /'(/• ( j; Xr). Thus
This expression can be .simplified using the Poisson properties. The direct
computation of AqB{Q, r) by marginal analysis is more* tricky and the
reader is asked to try it in Problem h-10.
Marginal analysis is olten a helplul way to look at a problem. In some
cases it (‘an simplify considerably the amount of work required to obtain
the optimization conditions. In other eases the analysis is so complicated
that, no groat .savings in time results. In principle, however, marginal
analysis could be. used to obtain the optimization equations for all the
models developed thus far in this book. Of course, often if. is desired to
have an explicit expression for the profit or cost, and in this ease, the
potential savings in labor by use of marginal analysis to avoid computing
the profit or cost expression am not realized. In such eases, however,
marginal analysis can be helpful in cheeking the result obtained by directly
dineienc.ing ^1<f cas(l or I,r<>fit expression. Infinitesimal marginal analysis
can be used when the variables are treated as continuous. Here one con¬
siders the change on going from say .r to x 1 tlx. Some examples of
infinitesimal marginal analysis are considered in the problems.
REFERENCES
PROBLEMS
6-7. Show how to incorporate “all units” quantity discounts in the model
of Sec. 6-2. Assume that the selling price remains constant. Devise
a procedure for computing the optimal order quantity. Sketch
curves equivalent to those of Fig. 2-13 for the case at hand.
6-8. Show how to incorporate incremental quantity discounts in the model
of Sec. 6-2. Devise a procedure for computing the optimal order
quantity. Assume that the selling price remains constant. Sketch
curves equivalent to those of Fig. 2-15 for the case at hand.
6-9. Solve Problem 6-1 under the assumption that “all units” quantity
discounts are available. If a quantity h is purchased, the unit cost
is $4.00 if 0 < h < 225, $3.00 if 225 < h < oo.
6-10. Solve Problem 6-1 under the assumption that incremental quantity
discounts are available. The first 225 trees cost $4.00 each and
additional trees cost $3.00 each.
PROBLEMS 317
6-11. Consider Problem 6-1. Assume that “all units” quantity discounts
are available. If a quantity h is purchased, the unit cost is $4.00 if
0 < h < q, and $3.50 if q < h < co. Determine the value of q such
that the optimal order quantity occurs at the price break. For what
range of q values will $4.00 be the optimal unit cost. For what range
of q values will $3.50 be the optimal unit cost.
6-12. The buyer of a large West Coast department store must decide what
quantity of a high priced women’s leather handbag to procure in
Italy for the coming Christmas season. The unit cost of the handbag
to the store is $17.50 and it will retail for $50.00. The buyer is
confident that all handbags not sold by the end of the season can be
disposed of at cost. However, for every dollar invested in a handbag
not sold at the end of the season, the buyer feels that he is really
losing $0.30 since the dollar invested in something else could have
yielded this gross profit. The buyer believes that he will sell more
than 50 of the handbags but not more than 250. Sales of any number
within these limits seem equally likely. How many handbags
should the buyer procure? Suppose now that, after discussions with
other buyers, he can sharpen his sales estimate to a point where he
believes sales will be normally distributed with mean 175 and a
standard deviation of 20. Now what is the optimal quantity to
procure? How much was it worth to the buyer to gain the additional
information?
6-13. Consider the model described in Sec. 6-5 whose expected profit is
given by (6-16). Suppose now that the length of the period is gamma
distributed. Determine the expected profit g(ft) and Ag(ft).
6-14. Recall that for a first order approximation in At = ti — to, Taylor’s
theorem reads f(k) - f(t0) = f(to)At, where f(to) is the derivative of
/with respect to t evaluated at to. Use this result to simplify P(h; \iU)
— P(ft; \%ti—i), tJP(h; \%ti) ti—iP(h) p{h) \{ti) p(Ji] \iii—i)
and tip(h)\iti) - U-xpQi) \iU~i). What does Eq. (6-24) reduce to in
this case? Why is it now much easier to evaluate numerically?
6-16. Point out the problems involved in attempting to allow the length
of the period to be a continuous random variable when the mean
rate of demand is also allowed to change with time.
6-17. Derive Eq. (6-6) by computing directly the change in expected profits
when the quantity procured is changed from ft to ft + dh.
318 SINGLE PERIOD MODELS CHAP. 6
6-18. The demand for a spare engine part is expected to behave in ac¬
cordance with a Poisson distribution with mean rate 0.01 units per
day. After 600 days elapse, it is known that the part will be obsolete
and will be scrapped for $200 less than its original cost. The costs
o carrying the unit in inventory amount to $0.10 per unit per day.
li parts are not available when engines break down, the daily costs
ot idle time are $900 per part needed and not in stock. What number
of spare parts should be acquired in advance for the 600 day period
of usage if all parts must be purchased at this time?
6-19. A farmer has to decide on which crops to take into the city on his
weekly trip. He has more than enough ripe crops to fill his truck
which can carry only 8000 lbs. of vegetables. Cabbage, cauliflower,
and tomato crops are the farmer’s leading candidates. The demands
i oooheon^ree Cr,°PS 'in P°unds) are normally distributed with means
, , and 2000 respectively. The corresponding standard
eviations are 200, 300, and 400. If the crops are not sold they are
worthless, while the profits per pound are 4 cents for cabbage, 6 cents
for cauliflower, and 7 cents for tomatoes. How should the farmer
load his truck to maximize expected profits?
6-22. Given that demand for a spare part is Poisson distributed with mean
0.1 umts per day and two alternative modes of delivery, namely two
ays and eight days at costs of $18.00 and $10.00 per unit respectively.
PROBLEMS 319
6-27. Assume in Example 3 of Sec. 6-3 that a goodwill loss of $1.00 per
rabbit demanded but not in stock is incurred. What is the increase
in optimal stock? What is the change in expected profits?
6-28. Consider the computational procedure discussed in Sec. 6-4 for
handling constrained problems when the variables are treated as
discrete. Under what circumstances is the procedure exact? When
it is not exact, under what circumstances should it yield a good
approximation?
6-29. Discuss the relationship between Eq. (6-4) and Eq. (4-116) with
7T = 0 for steady state models with Q = 1. Why should they be
of the same form?
6-30. Show for the general single period model discussed in Sec. 6-2 that
h* is the solution to
J
DYNAMIC
INVENTORY MODELS
in fewer offices?“
i
7-1 Introduction
In a strict sense, steady state conditions are a fiction in the real world.
The essential characteristic of all economic systems is that they are con¬
tinually changing with time. For inventory systems, the processes gen¬
erating demands and lead times change with time, as do the various costs
of interest, and even the items carried by the system. In many cases,
however, the changes occur slowly enough so that for considerable lengths
of time the system can be treated as if it were in a steady state mode of
operation. In other instances, however, the changes occur with such
rapidity that they must be explicitly accounted for. Usually it is the
changes in the process generating demands which are most important.
It is the purpose of this chapter to study multi-period models in which the
mean rate of demand changes with time.
As might be expected, the difficulty of formulating and obtaining
numerical solutions to realistic dynamic inventory models is considerably
greater than for the case where it was permissible to assume that the
system was in steady state. In fact, when demand is treated as a sto¬
chastic variable whose mean is time dependent, only the most trivial
problems can be solved manually. Usually a large digital computer is
needed to obtain numerical results. The natural formalism for setting up
dynamic models in a form for numerical computations is dynamic
programming.
Before turning to dynamic inventory models themselves, we shall first
introduce dynamic programming and cover the topics in this ‘subject that
will be needed here.
X
3 = 1
v>xi ^ V (7-2)
where the vj; V are specified constants, and fj{Xj) is a function of Xj only.
We shall assume that the vs, V are integers. There is no real loss in gen¬
erality in doing so, since by proper choice of the physical dimensions (i.e.,
perhaps using cubic centimeters instead of cubic feet), it is always possible
to obtain an arbitrary degree of accuracy with the v,, V being integers.
The single period, multi-item problems subject to a constraint, such as the
flyaway-kit problem discussed in Chapter 6, are special cases of the general
problem just formulated.
We shall introduce dynamic programming by studying how one might
attempt to solve numerically a problem such as that formulated above.
The important thing to notice is that we can minimize over the variables
in any way we choose, provided that the procedure we select does permit
the examination of all possible combinations of values for all the variables,
if necessary. Suppose then that we proceed as follows: We select a value
of x„ and we minimize g over Xi, ..., xn_i for this given value of xn. By
(7-2), we see that the variables xx,... , xn_t must satisfy
71 — 1
X
y=i
V’Xi ^ V - VnXn (7-3)
so that the allowable range of variation for Xi, . ... , xn-i will depend on
the value of xn selected. Note that the value selected for xn can only have
the integral values 0, 1, 2, . . ., [V/vn], where [V/vn] denotes the largest
integer less than or equal to V/vn.
Now*
n—1
min 9 = f(.Xn) + min V ffe) (7-4)
XI, , Xn-l XI, , Xn-1
and in computing
n— 1
min X fiixs) (7-5)
..
xh , xn^i must satisfy (7-3). We can observe that the minimum value
expressed by (7-5) will depend on V - vnxn because of (7-3). Denote (7-5)
by Zn-.i(V — vnxn). Then Zn-i(F — vnxn) is the minimum of 1 fj(xj)
* The notation min g will be taken to mean the absolute minimum of g(xu ...<xn)
for non-negative integers which satisfy (7-3). Equation (7-4) can there¬
fore be written
min q — fixn) ~i~ Zn—i(V VnXn) (7-6)
XI, ..., Xn-1
Hence, if we knew the function Zn-i(£) for all integral arguments £ from
0 to V, we could determine g* simply by computing /n(0) + Zn-i(F),
fn(l) + Zn-i{V - O, /»(2) + Zn-i(V - 2vn), etc., up to = [7/wn],
and picking the smallest of these. We would simultaneously determine xt,
the optimal value of xn, in doing so.
The question then arises as to how we determine Zn-x(£) for any argu¬
ment £. By definition
We can now resort to the same trick as above. Suppose that we pick a
value of £»_i, and minimize /yfe) over a*, . . . , xn_2 for non-negative
integers rcy, j = 1, . . ., n — 2, satisfying
n—2
y ^ vjXj ^
i=l
2 < £ (7-10)
y=i
Then
2/n—1(£) ^ mm [/n—l(^n—l) “b 2(£ ^n—l^n—l)] (7-11)
Then we can compute a table giving Z3(£), etc., until finally from the
table for Zn^i(£) we obtain #* using (7-7).
Let us now repeat the computational procedure for determining g*, and
at the same time indicate how x*, . . . , x$ are determined. We define the
sequence of functions
k
Zk(£) = min ^fj(xj), k = 1,. . . , n (7-13)
XX, ... ,Xk j—l
X
J-l
VfXj < £ (7-14)
Then
g* = Zn(V) (7-15)
and the 2^(£) can be computed recursively using the recurrence relation
where in the minimization Xk can take on only integral values in the range
shown. In order to compute Z*(£) for a given £, we will quite possibly need
to know Zk~i(r}) for all 77 = 0, But then in order to compute
zk+i(f), we will quite possibly need Z*({) for ah £ = 0, 1, ..., f. Finally,
since g* = Zn(\), it follows that each Zk(£) may need to be known for all
£ = 0, 1, 2,. . ., F.
The computational procedure can be thought of as an n stage process.
At the first stage we compute a table of Zi(£) for £ = 0, 1, . . . , V, and at
stage 2 a table of Z2(£), and finally at stage n the single value Zn(V). Note
that once the table Zk(£) had been obtained the tables for Z*(£), i < k, are
no longer needed.
In computing Z^(£) from (7-16) or Zi(£) from (7-12) we also obtain for
each £ the value of xk which yields the minimum. Denote by £*(£) the
value of xk which yields Z&(£) in (7-16) or (7-12). Suppose now that as we
record the table of £&(£) we also record a table of £*(£). At the last step,
when Zn(V) is computed, we automatically determine x$ and no table is
SEC. 7-2 DYNAMIC PROGRAMMING 327
Thus in general
X% = xk (v — ]T} k = 1| • • • J n — 1 (7-17)
\ j**h+ i 7
Then, if 3C* is the optimal value of X, X* = Z3(9). For k> 2, the Zk(£)
can be computed sequentially from
TABLE 7-1
2716 1 4116 0
1
3 1213 3 2613 0
369.3 5 1769.3 0
5
76.29 7 1164 1
7
11.04 9 748.2 2
9
330 DYNAMIC INVENTORY MODELS CHAP* 7
Let us generalize the problem studied in the previous section to the case
where two constraints are present, i.e., we now wish to minimize
n
for non-negative integers xy, j = 1,. . ., n, which satisfy the two constraints
where the vj, Wj, V, IF are integers. Precisely the same procedure can be
used as before except that now Zk(£) must be replaced by a function of
two arguments 2jk(£, 77). Let us define the set of functions
k
Zk(£, v) = min ][]/y(xy), k = 1,. . . , n (7-22)
Xp .. .,xk y = 1
XI< £; X) ^ ‘n (7-23)
i=1 ;=1
Then g* = Zn(F, IF). Since the minimization can be carried out in any
way that we desire provided that all possible combinations of the values
of the variables are considered if necessary, it follows that for k > 2
SEC 7-3 FORMULATION OF OTHER PROBLEMS 331
i *~j
Finally
Zi(£, 7]) = mmfxixi) (7-25)
where
yic+1 = £ + %k — du
and
£«(£) = min/»(f, s») (7-32)
where
yk + Xk — dk = £
and 2/2,. . . , 2/fc satisfy (7-29). Then the recurrence relations take the form
and
J?i(£) = min/i(£ + dk — xh Xi) (7-37)
where 5Zy=i fj(y3j x3) is evaluated for a given set of y3 (the y3- being related
to the dj by (7-29)) and
n 3
p {d3) = Pi(di)p (d ) 2 2 . . . Pnidn) (7-39)
i=i
is the probability of the given set of dj, since the dj are independent
random variables.
As before, we assume that yi is specified. For a stochastic process it is not
possible in general to specify the state of the system at the end of the last
period, and hence a backward solution will always be needed in this case.
By analogy with the deterministic case we define the set of functions
Z [ n py(dy)l = n I" Z = 1
alldi>0 Li = * J j = Jc Lrf. = o ->
j = k, ... ,n
Hence
En ~] r -
min xk) + min n py(dy) Z
Xk l Xk+U...,Xn all flfj >0
u<ti>0 U=k JLyii+l -I
j = k, . . ., n
SEC 7-3 FORMULATION OF OTHER PROBLEMS 335
= min xk)
expression. The only holding costs that are relevant are those incurred
between h and f.
By period j we shall mean the time from t3 to ij+i (tn+1 = f). Write
T3 = tj+i — t3. Let yi be the on hand inventory at time h. The demand
rate as a function of time will be written \(t). Then the demand d3 in period
j is
rti+1
d3 = J A(t) dt
We are now ready to develop the total variable cost expression. All
variables will be treated as continuous. Let I3C be the cost of carrying one
unit in inventory for period j. Note that this is not an annual cost but the
cost for period j. If y3 is the on hand inventory at the beginning of period j
before any order placed at time fy has arrived, then the inventory carrying
charges for period j are
i c fti+i r n i t o fu+i r*
-jT Jt Vi + Qj ~ Jt Hu) du dt = I3C[y3 + Q3] - ~jT jt Jt Hu) du dt
Now note that I3Cy3+i is the inventory carrying cost in period j for those
units carried into period j + 1. The other term in (7-43) is the cost in
period j of carrying the d3 units which are demanded in period j. This
latter cost is independent of the Q3 and is unavoidable, since the d3 units
demanded in period j must be on hand at the beginning of period j. Hence
the carrying cost in period j for these d3 units need not be included in the
variable cost expression.
Thus the variable costs of ordering and holding inventory which are
incurred over the planning horizon are
n
where
0 if Q, = 0
$3 ~~
1 if Qj > 0
where yk+i = £ and (7-42) holds for the other ys. Then from (7-36), (7-37),
the recurrence relations are
where
'£ + di — yi if £ > 2/i — di (7-48)
.0 otherwise
We shall now restrict our attention to the case where yi = 0, i.e., the
I
time fi is chosen so that everything on hand at time fi will be used up just
as the first order arrives. There is normally no reason to have a procure¬
4
ment arrive before everything on hand is used up, and hence the case of
yx — o is the one that is usually of interest. If yi = 0, it is possible to make
several observations that will considerably reduce the computational effort
below that required if the problem was solved directly using the above
recurrence relations. Let us note first of all that if QI > 0 then yi = 0, i.e.,
the on hand inventory at the time of arrival of an order should be zero.
The reason for this is that if yk > 0, a saving in inventory carrying costs
SEC. 7-4 DYNAMIC-DETERMINISTIC LOT SIZE MODEL 339
(7-50)
Ak Zk-i(fi)
= min
Zk~i(dk)
where Zq(0) = 0, or if
Yk(w) = Aw + C 22 IJ i =j22+1 di
j = w L-
^w-i(O)
Note that Yk(w) is the cost for periods 1 through k under the assumptions
that the on hand inventory at the end of period k is zero, an optimal policy
is followed for periods 1 through w — 1 given that the ending inventory
for period w — 1 is zero, and an order arrives at the beginning of period w
which satisfies the demand for periods w through k.
We can now show that if for Zk—i{0), the order which satisfies the demand
in period k — 1 arrives at the beginning of period v, then in computing
Zk(0), the order which satisfies the demand in period k will arrive at the
beginning of period v or a later period, i.e., in computing Zk(0) it is un¬
necessary to consider the possibility that the order which satisfies the
demand in period k arrives before period v. Then in (7-53), it is only neces¬
sary to allow w to range from y to k. We shall prove this by contradiction.
Assume that
fc ^ — Jc —«
Zjk(O) = Yk(u) = Au + C h % di + 0) <
j=u L i=j+1 *-*
where
and v > u. However, on factoring out in Zk{0) the carrying costs for the
items carried from period u through k to satisfy the demand in period k, we
have
But
Yk(v) = Cdk 2 h +
since by assumption
k-1
F*-i(w) > F*_i(t;) = Z*-i(0) and Cd* ^ Jy > 0
j=u
Z»(£) = min
w
aw + cj_;[// J d* + f) Zw-1
- j=w «— ' —
and again w only needs to run from v to n, where in Zn-i(0) the order which
satisfied the demand of period n — 1 arrived at the beginning of period v.
The reason that the same computational procedure holds for Zn(£) as for
Zn(0) is that having £ units remaining at the end is equivalent, insofar as
the previous periods are concerned, to increasing the demand in period n
by £.
A convenient tabular format for making the computations is given in
Table 7-2. In the last line, the optimal Q3- values are indicated by enclosing
in parentheses all periods whose demands are met by the order arriving at
the beginning of the period whose number appears first in the parentheses.
Thus (3, 4, 5) would mean that it is optimal to meet the demands in periods
3, 4, and 5 with the order arriving at the beginning of period 3. Table 7-2
indicates the results which might be obtained in a typical case. The aster¬
isks on the Yk(w) indicate the minima over w, i.e., Zk(0).
In the above model the times at which orders could be placed were
specified. A more general problem would be to determine both the times
at which orders should be placed and the quantities to be ordered. The
problem of determining the precise times at which orders should be placed
when the demand rate varies with time turns out to be quite difficult even
in the simplest cases. Problem 7-6 will ask the reader to derive the equa-
i
SEC 7-5 DYNAMIC LOT SIZE MODEL-EXAMPLE 343
tions which determine these times for the case where there are no fixed
ordering costs, and to examine the difficulties involved in solving them.
It will be noted, however, that by allowing the times at which orders can be
placed to be sufficiently close together in the model developed in this
section, it is possible to come arbitrarily close to the solution where the
optimal times at which orders are to be placed are to be determined.
This shows that it is not optimal to have another setup at the beginning
of the second period, if the on hand inventory at the end of the second
period is to be 0. Next
= 800 = F3(3)
SEC. 7-6 STOCHASTIC DYNAMIC MODELS-FIXED HORIZON 345
Tj + ry+1 - ry /,
JM
Z-j (&* + Qi - x)v\*)
x~0
W ~ fy)] dt
JjCj J Tj+Tj+l
Tiy -f
+ Ty+i Tj Vi + Qi ~
(7-56)
+ L (p - Vi — Q])v(p', V) r dr
x=Vi+Qi J
where
(Ty + ry+i — Tj)B(y3 + Qy, T3 + ry+i, ry)
SEC 7-6 STOCHASTIC DYNAMIC MODELS—FIXED HORIZON 347
oo
E(Vi + Qi, Tj + r]+ , t/) can be written down explicitly by use of A3-10.
1
All the terms needed in the cost expression have now been evaluated.
The yj are related by the material balance equation
Vm = Vi + Qi - dj (7-59)
where d, is the demand from to fy+i- The probability of dj is p(dj) <jjT3).
It follows that the expected variable cost over the planning horizon of
procurement, carrying, and backorders, net of the expected gain from sales
at time f, is
(7-60)
3£ can also be interpreted as the discounted cost if it is imagined that the
discount factors are included in the As, Cj(Qj), Vj, fij and L.
Dynamic programming provides a natural way to determine numerically
the functions Q*(y,). Because of the stochastic nature of the process a
backward solution is needed. Let us define the set of functions
It should be recalled that a numerical value can be obtained only for Qi.
For the other Q/ s, we obtain a function Q*{ijj), and it is not possible to
determine Q* until y, is known, i.e., until the time is reached when Qj is to
be selected. In practice one will often only be interested in Q*, and will not
be interested in the Q%y3), j > 1. The reason for this is that the problem
will often be solved each time a procurement is to be considered, using
additional information obtained since the last time the computation was
made.
The procedure for numerically solving the problem has been outlined in
Sec. 7-3. In most cases the task is much too burdensome to carry out by
hand. It can be solved fairly readily on a large scale computer, however.
Even on a large scale computer, though, it is sometimes desirable to intro¬
duce shortcuts to save time. For example, if it is possible for Q* to be
quite large, say over 100, the task of minimizing over Qk when Qk is changed
by only one unit at a time may be excessively time consuming. To reduce
the time required to make these computations, one might, for example,
vary Qk one unit at a time between 1 and 25, five units at a time between
25 and 100, and ten units at a time over 100. In other words, Qk cannot
take on any integral value when Qk > 25, but instead can only take on the
values 30, 35, 40, etc., to 100 and then 110, 120, etc. The upper limit on £
will simply be the maximum level that it is ever expected that the inventory
position will reach. Again, to save computational time, it is advisable to
SEC 7-7 STOCHASTIC DYNAMIC MODELS—VARIABLE HORIZON 349
keep it as small as feasible. If the expected lead time demands are suffi¬
ciently large, it will be desirable to replace the Poisson terms by their
normal approximations.
It is also possible to give a formulation of the above model in the case
where the lead times are random variables. Problem 7-9 asks the reader to
do this. The lost sales case is, as usual, very difficult to handle. All the
problems encountered in the steady state period review systems are also
present here. In the event that more than a single order can be outstanding,
it is no longer possible to have the Zt be a function of a single argument
only. This means, essentially, that at the present time, it is impossible to
solve such problems numerically. We ask the reader to examine the lost
sales case in Problem 7-10.
to have more times at which procurements can be made than for the smaller
f(i)- In general, we shall suppose that if f(i) is the time of obsolescence,
there are n(i) times at which procurements will be considered, and these
times will be denoted ft(i), j = 2, . . ., n(i); ft = 0, which is the time that
the first procurement is to be considered will, of course, be the same for all i,
since at and prior to ft, it is not known which value of f (i) will actually
occur. Corresponding to the Xy, <ry, Th ry defined in the previous section,
we now have a corresponding set of these for each i, i.e., Xy(i), oy(i), 2y(t),
Tj(t). Also Aj(i), Ij(i), xy(i), £y(i), L(i) replace Ah ly, Ch xy, xy, L
respectively.
As before, y\ will be taken to be the inventory position of the system at
time ft. Let W i{yx + Qx) be the expected cost if Qx units are ordered at
tune ft, f (j) is the date of obsolescence, and an optimal policy is followed
after time ft. Then
and the other Zjp are defined by (7-62), (7-63), except that all relevant
parameters must now involve L
The expected cost for any choice of Qx is then
m
and the optimal Qx is that value which minimizes (7-65). To solve this
problem, we must essentially solve m problems of the type discussed in the
previous section, since we need the m functions 2^}(£). Then at the last
step these are joined together by (7-65) and Q* is determined. At each time
a procurement is to be considered the problem will be solved and any
additional information on the date of obsolescence obtained since the
previous solution will be incorporated.
In Secs. 7-6, 7-7 we assumed that the distribution of demand over any time
period was Poisson and that the mean rate of demand was known as a
REFERENCES 351
REFERENCES
PROBLEMS
7-1. An isolated air base is supplied with spare parts only once every
month when a plane from the main depot arrives to replenish the
inventory. There are n parts which are stocked at the air base. Just
prior to the time the base is to be replenished, it radios to the main
depot the on hand inventory of each of the n spare parts. Let y%
be the number on hand of item i just prior to replenishment. The
plane which brings the spares can carry a volume V and the unit
volume of item i is Vi. A cost n is incurred each time a demand occurs
when the base is out of stock. Monthly demand for each of the items
is essentially Poisson distributed with mean /it. Set up as a dynamic
programming problem the problem of determining the quantity Xi of
each item i which should be loaded on the plane, so as to minimize
the expected stockout costs for the coming month.
7-2. Consider a simplified version of Problem 7-1 which involves only
three items. Assume that: V = 15 cu. ft.; vx = 5 cu. ft.; v2 = 3 cu.
ft.; ^3 = 4 cu. ft.; yx = 1; y2 = 0; yz = 2; tti = $5000, tt2 =
$2000, tz — $10,000; mi = 1, /i2 = 5, /x3 = 3. Determine the optimal
values of xh x2, and xz.
7-3. A newsboy can carry only N papers with him in the evening. There
are two papers which he sells. The first is an out-of-town paper
which costs him Ci cents each and sells for px cents. Any papers not
sold can be returned for a credit of dx cents. The other is a local
PaPer which costs him c2 cents each and sells for p2 cents. Any papers
not sold can be returned for a credit of d2 cents. Assume that if yXi
y2 are the demands for the out-of-town and local papers respectively
then yXj y2 are independent random variables with densities pi(yi)}
V*(yz)- Formulate the problem of determining the quantities xXf x2
of each of the papers to procure as a dynamic programming problem.
Solve the problem in the particular case where cx = 7, px = 10,
di = 4, c2 = 5, p2 = 7, d2 = 3. The demand for each is Poisson
PROBLEMS 353
distributed, the mean being 7 for the out-of-town paper and 8 for
the local paper. He can carry no more than 15 papers.
7-4. Consider a flyaway kit problem in which there are only three items
to be considered. The total volume available is 11 cu. ft. The unit
volume of item 1 is 2 cu. ft., of item 2 is 3 cu. ft., and of item 3 is 4
cu. ft. The cost of stockout for item 1 is $500, for item 2 is $1000,
and for item 3 is $2000. The demand for each item is Poisson
distributed with mean 3 for item 1, 2 for item 2, and 1 for item 3.
How many of each item should be loaded in order to minimize the
expected stockout costs?
7-5. Consider the cost equation given by Eq. (7-44). Assume that a rate
of interest i is being used. Determine the discount factors which
should appear in the Ah if it is desired to have X be the discounted
cost. Include the discounted cost of the units, and work out a means
for solving the problem in this case.
7-6. Consider a situation in which the demand for an item can be taken
to be deterministic, with the cumulative demand as a function of time
being D(t). Assume that from t = 0 to t = T, n procurements will
be allowed. There is no fixed cost of making a procurement. Derive
the equations which determine the times U at which procurements
should be scheduled to arrive, so as to minimize carrying costs over
the interval 0 to T. Show that the U are solutions to
J = 2,3,.. ,n — 1
dti {j) tj-tw
lots, if 5000 capacitors are now on hand and it is desired to have 5000
on hand at the end of the year.
7-8. A manufacturer of actuators for rocket motors is under contract to
deliver the following quantities to a producer of rocket motors on the
first of each month for the coming year: 25, 50, 100, 175, 200, 200,
175, 150, 100, 75, 200, 150. The actuators are produced in lots and
the setup cost is $800. Each actuator costs $300 to produce. The
manufacturer uses an inventory carrying charge of I = 0.20.
Determine when setups should be made and the optimal lot sizes.
Ignore the time required to produce a lot.
7-9. Generalize the model developed in Sec. 7-6 to the case where the
procurement lead times are random variables. What assumptions
must be made?
7-10. Investigate the problems in treating the lost sales version of the
model developed in Sec. 7-6. Show that in the Zk, one must include
parameters for the size of each of the orders outstanding. Formulate
the model in the case where there is never more than a single order
outstanding.
7-11. What simplifications occur in the model developed in Sec. 7-6 if
it = 0, the terms involving r can be neglected in comparison to those
involving T + r, and the backorders term in the inventory carrying
charge can be neglected?
7-12. Demands for the three weeks of a season are deterministic and have
the values 1, 2, 2 units respectively. The total costs of the product
delivered at the beginning of each period are given by the following
table.
Weeks 12 3
No. Units
0 0 0 0
1 1.00 0.75 2.00
2 2.00 1.50 3.50
3 1.50 2.00 5.00
4 4.00 5.00 6.50
5 6.00 8.00 8.00
7-14. A retailer sells an item with the following characteristics. The item
costs $5.00, has a selling price of $9.00, and a liquidation value of
$2.00. The supplier arrives with his merchandise at the start of each
week of the three week season. The initial inventory is 0 and inven¬
tory carrying charges are negligible. Demand in the three weeks is
given by the following table of probabilities, the entry in each cell
being the probability that demand is exactly equal to the number
of units indicated in the left hand column. Determine the optimal
Week 1 2 3
No. Units
1
0 i 0 "3
i 1 1
1 4 2 3
1 1
2 0 2 1
number of units to order the first week and an optimal policy for the
remaining two weeks.
7-15. A manufacturer is able to increase production in any month at a cost
of $4.00 per unit increase in the production rate (when measured in
the dimensions of units per period). Reductions in the production
rate cost $3.00 per unit decrease. It is known that demands in the
next three periods are 2, 4, and 1 units respectively. Production
costs per unit are $6.00, $10.00, and $8.00 in periods 1, 2, and 3
respectively. Inventory carrying charges are applied to ending inven¬
tory and amount to $2 per unit. Assuming that demands must be
met as they occur and that initial and final inventories are zero,
determine the least cost levels of production in the three periods.
7-16. A camera store sells its Developo Camera at $120 per unit. Its
supplier appears each Monday morning and will sell additional sets
to the firm at a specified price S or buy sets back at price B. There
356 DYNAMIC INVENTORY MODELS CHAP. 7
Period 1 Period 2
S(S) 80 90
B(S) 70 70
Value ($)
No. A B C
1 6 4 5
2 8 6 7
3 10 7 9
Cost ($)
No. A B C
i
1 1 1 2
2 3 2 3
3 5 4 4
7-24. A business has the average cost and revenue schedules described
below, where x1} x2, xz represent the quantities produced in periods
one, two, and three respectively, and yh y2, yz represent the quantities
demanded in periods one, two, and three.
358 DYNAMIC INVENTORY MODELS CHAP. 7
IN THE ANALYSIS
where ICDt and tEt + tBt are the expected carrying and backorder
costs respectively incurred from f + rtof + r+ T (and discounted to
time f). Let a(T), 0 < a < 1, be the discount factor which discounts any
costs known at f + T to their present worth at time J.
We shall define £$■(£; T) to be the present worth (discounted cost) at
time f of all future costs (with the convention that no carrying or backorder
costs incurred from f to J + r are included) if an optimal quantity is
ordered at time f and at all future review times, £ is the inventory position
of the system at time f before any order is placed, and T is the time between
reviews. In order that £$•(£; T) will be finite, it is necessary to introduce
discounting and to assume that the interest rate is positive. Similarly,
2j-+r(£; T) will be defined to be the present worth at time J + T of all
future costs if an optimal quantity is ordered at time f + T and all future
times, and £ is the inventory position of the system at time f + T before
any order is placed. By convention Zf+r(£; T) will not include the carrying
or holding costs incurred between f + T and f + r + T. Then we see
from the methods introduced in the previous chapter that
T) = {J + A5 + CQ + /(£ + Q; T)
Q
where
if Q = 0
if Q > 0.
In other words, the present worth of all future costs at a review time, when
the inventory position is £ prior to placing an order and an optimal policy
is followed at that review time and all future reviews, is the same at any
review time. The reason is of course that we are considering the discounted
cost over an infinite planning horizon.
Therefore, if we define Z(£; T) to be the present worth at a review time
of all future costs (with the convention that no carrying or backorder costs
are included until a lead time has elapsed) when an optimal policy is
followed at the review time and at all future review times, the inventory
position of the system is £ at the review time prior to placing any order, and
the time between reviews is T, then (8-2) reduces to
A + mm F(t + Q; T)
Z(f; T) = J - Cf + min
F& T)
to y, and R*(T) the unique point where this minimum is taken on (R* will
be a function of T). Also let r* be the value of y(y < R*) for which F(y; T)
= A+F(R*;T), i.e.,
F(r*; T) = A + F(R*; T) (8-8)
be increased if an order was placed. On the other hand if £ < r*, say £2,
then it will be optimal to place an order, and the quantity ordered should
be sufficient to make F(y;T) as small as possible, i.e., R* — £2 should be
ordered. Then the costs of placing an order and future discounted costs
come to A + F(R*; T) < F(£2; T). Furthermore, for Q different from
Q* = R* — £2j the expected discounted costs would be greater than A +
F(R*; T). It is also clear that if £ > i2* an order should not be placed.
However, if the system is operated optimally, £ will never be greater than
R*. What we have shown then is that an Rr operating doctrine is the
optimal one to use in this case. It is interesting to note that the same sort
of analysis shows that if A = 0, then an order up to R policy is optimal.
366 DYNAMIC PROGRAMMING USED IN SS MODEL ANAL. CHAP. 8
position y is relatively low, the expected costs will be relatively high be¬
cause of high backorder costs in the present period. As t/ is increased the
expected costs should decrease until y is large enough so that the increase
in carrying charges outweighs the decrease in backorder costs. Then the
expected costs will begin to rise and continue rising as y is increased.
Addition of the linear yC term will not change the basic shape of the curve.
Hence we expect F(y; T) to yield a curve of the shape shown in Fig. 8-1
and not that shown in Fig. 8-2. Thus, we expect in general that, when the
unit cost is constant, an Rr policy should be optimal.
In the event that the unit cost of the item is not a constant, it is no
longer true that an Rr policy must be optimal. Now we can no longer
introduce a function F(£ + Q; T) which depends on { + Q alone. Instead,
the functional equation has the form
use of the observation that the steady state case with an infinite number of
periods can be thought of as the limit as n approaches infinity of an n
period system. Let Zn(£; T) be the discounted expected cost for an n
period system when £ is the inventory position at the beginning of period
n before any order is placed (period n, however, comes first in time) and
an optimal policy is used for each decision. Note that the periods are being
numbered in an order opposite to their time sequence. Then
and
Zn(i; T) = min [J + A5 + CQn + /(£ + Qn]T)
Qn
where
Fn(£ + Qn; T) = (£ + Qn)C + /(£ + ; T)
n = 2, 3,. . .
Also
Zi({; T) = min [J + A8 + CQ1 + /(£ + &; T)] (8-12)
Qi
The proof would be very simple if it followed that the Fn(£; T) were
convex if / (y; T) was convex. This is not in general true, however. The
sort of phenomenon that arises which prevents the Fn(£; T) from being
convex can be seen by studying just a simple two period case. Now from
(8-12)
ZlG; T) = J - Cl + min [Ad + Fl(i + Qi; T)]
Qi
where
Fi(r, T) = Cy +f(y; T)
and Fi(y; T) is a convex function of y, since f(y; T) and Cy are. Thus from
the discussion of the previous section, we know that an Rr policy will be
optimal for the one period case and hence
Zi(J; T) is plotted in Fig. 8-4 for the case where f(y;T) is differentiable
everywhere with respect to y. Note that Zi(£; T) is not a convex function
of £. Zite; T) is continuous, however, although there is a discontinuity in
the derivative at £ = r*. Hence, we see from (8-11) that since F%(y] T)
involves Zi(£; T), F2(y; T) cannot be expected to be convex, in general.
By continuing the argument, it can be seen that the Fn(y; T), n > 1 will
not necessarily be convex.
It is the fixed ordering cost A which causes Zi(£; T)} or more generally
Zn(£; T) and Fn{y\ T), not to be convex even though f(y; T) is convex.
In fact, one can prove that when A = 0, the Fn(y; T) and F(y; T) are
convex functions of y if f(y] T) is. This result will follow from what is
demonstrated below. It can be obtained directly, however, and we ask the
reader to do this in Problem 8-2.
The above shows that since we do not expect F{y) T) to be convex, the
proof cannot be based on demonstrating the convexity of F(y; T). Indeed,
F may have a number of relative maxima and minima. As before, denote
by E* the point at which F takes on its absolute minimum. If this point
370 DYNAMIC PROGRAMMING USED IN SS MODEL ANAL CHAP. 8
is not unique, E* will denote the smallest value of y for which F takes on
its absolute minimum. Also, let R denote any point at which F takes on a
relative minimum (E* will be one such point). Suppose now that the fol¬
lowing conditions are satisfied. For each E, it is true that there are no
points y < R with F(y; T) > A + F(E; T) at which F takes on a relative
maximum. This means that the curve z = F(y; T) can cross the horizontal
line z = A + F(R) T) only once for y < R- Thus if r* < E is the point
where the curve z = F(y; T) crosses the horizontal line z = A + F(R*; T),
Q* = 0 if £ > r* and Q* = E* - f if £ < r* so that an Rr policy is
optimal. When the above conditions are satisfied, cases like that shown in
Fig. 8-2, and cases where there exist relative minima with E > E* such
that for certain ranges of £ > E* it is optimal to order up to R (illustrate
this graphically), are ruled out.
We shall prove below that if f(y; T) is convex and differentiable every¬
where with respect to y, then F(y; T) is continuous and differentiable
everywhere with respect to y (even though Z(£; T) need not be differ¬
entiable everywhere with respect to £), and that for any a > 0
assume that F(y; T) has a relative maximum at yi < R with F(yi; T) > A
+ F(R; T). A contradiction is obtained on using (8-14), if we set R =
a + yh since at ylf F'(yi; T) = 0, and (8-14) becomes A + F(R; T) —
F(yi; T) > 0. Hence, the truth of (8-14) assures us that an Rr policy will
be optimal.
It remains, then, to prove that (8-14) holds if f(y; T) is convex and
differentiable everywhere. For carrying costs and backorder costs of the
type we use in this book f(y; T) is always differentiable everywhere with
respect to y when all variables are treated as continuous. We have previ¬
ously defined a function f(y) as convex if for any values yi, y2 and any a,
0 < a < 1, then
where f(y) = df/dy. We ask the reader to show this equivalence and to
give a geometric interpretation in Problem 8-3.
Scarf defines a modified form of convexity called A-convexity, which is
a useful concept in carrying out the proof. For any number A > 0, the
differentiable function/(y) is said to be A-convex if for all y and any a > 0
a J v(x)Zn-i(y — x; T) dx
is. However, from property (c) above, this will be true when Zn-i{y) T)
is A-convex. To prove that Zn-i(y, T) is A-convex, we observe first that,
since by assumption Fn-i(y] T) is A-convex, the optimal ordering policy at
stage n — 1 will be of an Rr type, i.e., there exist numbers rS_i and Rt-i
such that if £ < rS-1, QS-i = Rt-i - £ and if £ > rS-i, QS-i = 0. Thus
Z'n-i(£; T) = -C (8-19)
since the optimal Qn~i will yield a value less than or equal to that for Q„_i
= a. However, when £ + a > r£_i
Z_i(£ + a; T) = J + /(£ + a; T)
so that 0 is a lower bound on the sequence and hence (how do we know that
F' exists?)
0(y, a; T) = lim Gn(y, ID
n—>«
for any y and any a > 0. Thus F(y; T) is A-convex, and an Rr policy is
374 DYNAMIC PROGRAMMING USED IN SS MODEL ANAL CHAP. 8
optimal for the steady state periodic review system if f(y; T) is a convex
function of y and the unit cost is constant.
Geometrically, the fact that F(y; T) is A-convex rules out cases such as
those shown in Fig. 8-2 where an Rr policy is not optimal. It does not,
however, rule out behavior such as that shown in Fig. 8-5 where an Rr
policy is optimal.
The solution to the functional equation (8-5) will yield Z(£; T) and
Q*(£; T). If an Rr policy is optimal, Q*(£; T) will yield the values of R*
and r*, since Q* = 0, £ > r*, and Q* = R* — £ for £ < r*. Hence Q* will
have the form of the curve shown in Fig. 8-6. Thus r* is the point where
Q* changes discontinuously from 0 to R* — r* and the magnitude of the
jump then determines R*.
A possible procedure for computing numerically 2?* and r* for a given T
is to approximate the infinite period system using n periods, as was done in
the previous section in proving the optimality of Rr policies. The appropri¬
ate recurrence relations for the n period case are given by (8-10) and (8-12).
Then QJ(£; T) is used as an approximation to Q*(£; T), and R* and r* can
SEC. 8-5 USE OF DYNAMIC PROGRAMMING TO COMPUTE R AND 375
small values of the interest rate the results will be independent of the inter¬
est rate. However, when the interest rate becomes large, the computed
results may depend on the value of the interest rate selected. It will be
shown later that the results computed using the n stage approximation
will most closely approximate the R* and r* that would be obtained using
the average annual cost expression of Sec. 5-9 if the interest rate is taken
to be zero.
To determine T*, the computations would have to be made for different
T values. In order to select the optimal T it is necessary to have some costs
to compare. Now if the n used in making the computation is sufficiently
large, then the minimum expected cost Z»(£; T) will be essentially the same
for any £ lying between R and r since any one of these values are allowable
376 DYNAMIC PROGRAMMING USED IN SS MODEL ANAL CHAP. 8
at a review time when the steady state system is operated. Thus T* can be
determined by finding the T which minimizes Zn[R*(T); T]. If the reader
does not see why any Zn{£; T) for between r and R can be used, a more
rigorous justification will be presented later.
It would, of course, be an impossibly laborious task to compute manually
R*, r*, and T* by the method outlined above. However, it is quite feasible
to make the computation using a large scale digital computer. The method
outlined above will, in general, require more computer time than the
computational procedure suggested in Sec. 5-9. It does, though, provide
an alternative procedure for computing R* and r*.
demand (note that we can allow the number of units to depend on the time
since the last demand), and Vn(t) the probability that precisely n units are
demanded in a time interval of length t following a demand.
Suppose that a demand occurs at time 0, and after taking whatever action
is appropriate, the inventory position of the system is £. It will be supposed
that it is never optimal to have £ < 0. Assume also that the next demand
occurs between t and t + dt. The probability of this is g(t) dt. We would
like to compute the expected cost (discounted to time 0) of carrying inven¬
tory and of backorders in the time interval r to r + t. Imagine that the
demand at time t is for j units. The probability of this is d(j\t). We shall
perform the computation in two parts, depending on whether r > t or
r < i.
In making the computations, we must, as in the periodic review case,
introduce a discount factor in order that the cost will be finite. Here there
is no natural period, and hence it is convenient to use continuous discount¬
ing. By this we mean that if we have a sum S (t) at time t, and if the interest
rate is i, then at tune t At, the sum S(t -j- At) which we have will be given
by
S(t + At) = (1 + iAt)S(t)
for small At, or taking the limit as At —> 0 and solving the resulting differ¬
ential equation, we obtain
Thus the present worth 3(0) at time 0 of a sum S(t) at time t is S(t)e~u so
that the discount factor is e~u.
Consider first the case where r > t. The probability that x demands
have occurred between time 0 and f, r < f < t + t, is
X
The discounted expected cost of the unit years of shortage incurred be-
r and r + t is, when averaged over t < r
- X —
*i;p\ 2] o J-1
d(j\t)Vx—j(£ 0
J
Q(f) df dt (8-24)
n
* I
Jo Jr
/
T-\-t
E E E kW^x - j + k> r - 0 d(j\t) g(t)dt dt +
k=ix=£y=i
e_irE E Ex
1
x = lfc = £~x3=1
°° X
— y+*;r — od(j\t)g{t)d{at
The second term in (8-25) arises because x may be less than £ while x + k
is greater than £.
The Wk{n + fc; t) can be computed recursively just as the Vn(t). It is
clear that
n
when averaged over t < r
r+t £
£ (€ -
x=l
r-X -
E dO'IO^-i(r - 0
Li== 1
“I
0(f) # dt (8-28)
We must compute now the costs corresponding to those above for the
case where t>r. From r to t the on hand inventory is £ (and no backorders
exist for this time interval since by assumption £ > 0). However, from t
to r + t} the computation must be made as above. The discounted ex¬
pected cost of the unit years of shortage incurred from t to r + t averaged
over t > r is
n r-f-f w °° X
£-1 oo x
+ x
E E
a; = l &= £—x
(k + x — £)TF*0 — j + k; t — t)
d{j\t)g(t) dt dt (8-30)
The first term arises because of the possibility of incurring fixed backorder
costs when the demand occurs at time t.
The expected carrying cost from time r to t + t averaged over t > r is
ft r« rT+t £
J e-«m) dtdt + icj Jf £ (f - *)«r<r
D® = E (f - *)*(*) (8-33)
x=0
E® = £ £ (* + ft - £)*(M) + £ Z>(*, *)
x = 0 k—£—x =l
(8-35)
+ Z(*-0«(O.*)
a;-=£
and
= [f
Jo Jt
e *'f ^d0'|i)7*_y(f — f) g(t)dtdt
Lj_i J
nr-H r a;
+ e-if 2 d(i|i)k'I_y(r - i) g(«) df di, a; = 1, 2, 3 .
(8-36)
- e~ir]g(t) dt (8-37)
380 DYNAMIC PROGRAMMING USED IN SS MODEL ANAL CHAP. 8
and
fc = 1, 2, . . .
x = 1, 2, . . . (8-38)
a= dt (8-42)
Hence
Y, p(i) = 1
i
and the pO*) can be thought of as probabilities. Note that a is simply the
expected value of the discount factor over the time until the next demand
occurs.
The functional equation (8-40) differs from other dynamic programming
problems that we have studied in that the minimization over y must be
carried out before summing over j, i.e., y* is a function of £ — j, not of £.
The intuitive reason for this is clear. The optimal quantity to procure will
depend on the number of units demanded when the next demand occurs.
SEC. 8-8 OPTIMALITY OF Rr POLICIES 381
The solution to the functional equation will yield Z(£) and y*(£ — j). The
function y*(f - j) provides the optimal operating doctrine which we are
seeking.
Intuitively, one would expect that an Rr policy would be optimal for trans-
actions reporting systems of the type discussed in the previous section.
We here wish to show that, normally, an Rr policy will be optimal. From
the results of Sec. 8-4 we know that an Rr policy will be optimal if /(£) is
convex when the variable f is treated as continuous and the unit cost is
constant (proof?). To see that /(£) will normally be convex, note that
ICD(f) + tB(£) can be written
2
x=0
~ x)v(x) (8-43)
where
7C(f - *), £ - x > 0
KS - x) (8-44)
fi(x — £), f — x < 0
00
Z(Q = /(£) - aCf + min {aA5 + a(£ + y)C + aZ(£ + y - 1)} (8-47)
y>0
Now an Rr policy must be optimal in this case. We can write
R* = r* + Q*. The quantity procured will always be Q*. Let r* + 1 be
the largest value of £ such that in (8-47), y* > 0. Also, write
F(x) = Cx + Z(x — 1) (8-48)
SEC 8-8 EXPLICIT SOLUTION FOR POISSON DEMANDS 383
When the system is operated optimally, the inventory position will never
get above R*, since one never orders above R*. Now substitute (8-49) into
(8-50), the resulting equation into (8-51), etc. This yields
Zir* + k) =
Recall that Z(%) can be interpreted as the discounted value of all future
costs (with the convention that no carrying or backorder costs are incurred
until a lead time has elapsed) when the inventory position is £ immediately
after a demand, and an optimal policy is followed for all future times. We
have noted previously that an Rr policy is optimal, and hence, when the
384 DYNAMIC PROGRAMMING USED IN SS MODEL ANAL CHAP. 8
since, when this Rr doctrine is used, the only possible values of £ are
r + 1,. . ., R. Thus equations (8-49) through (8-52) are satisfied if Z is
replaced by % and the asterisks on r, R, and Q are omitted. Hence an
explicit solution for &(£; R, r) if given by (8-55), if Z is replaced by 3* and
the asterisks are omitted on r and Q. Note that
It will now be shown how to relate 3(£; R, r) to the average annual cost
3C(Q, r). To do this, consider an ensemble of identical systems in statistical
equilibrium. If over any given interval of time of length At, we determine
the ensemble average of the costs incurred over this time interval and then
divide by At, the result obtained, when expressed in the dimensions of
dollars per year, will be simply the average annual cost X (Q, r), for any
At > 0, however small. Furthermore, the rate of incurring costs X which
is obtained will be independent of time. Now consider the quantity £
defined by
When we compute
lim ib(r + Jc; R, r)
i—>0
it turns out that the same result is obtained for each k = 1,,Q. We
ask the reader to prove this in Problem 8-11. The intuitive reason for this
is that as i —> 0, the contribution of the costs which depend on the starting
inventory position become a negligible part of the total discounted cost.
Because of this we have the simpler result
a = X = r~~—: (8-61)
Jo X+*
Thus
lim a0~> = 1, j = 0, ... ,Q (8-62)
i—>0
and
% 1 X
(8-63)
SS 1 - o« “ SS - . Q . Q(Q + 1) , _ Q
•L J- T •> 1 o\ 2 * ' * •
rj(x) = X ff
J0 Jr—t
p(x — 1; XT)e~u dT dt + X f
Jt Jo
f p(x — 1; \T)e~u dT dt,
x = 1, 2, . . . (8-65)
We now change the order of integration of the first term by use of the
familiar geometrical analysis employed in Chapter 3. This yields for the
first term
X / j p(x - 1 ;XT)e-udtdT
JO Jr-T
+ r X) (x ~~ r — *)p(«; Xr)
A x=r-Hfe
(8-69)
rp *
= Y(r + fc-M) + T ^ p(a;; Xr)
x=r-f&
+ ^ 2 2>-*)p<*Xt)
V y=r+l
However
00
= ^ (w - v)P(w; Xr)
388 DYNAMIC PROGRAMMING USED IN SS MODEL ANAL CHAP. 8
It is possible to solve (8-40) explicitly for Z(£) in the general case, but the
task is somewhat more difficult than when units are demanded one at a
time. Assume that an Rr policy is optimal. Then let r* be the value of the
inventory position such that if £ < r* after a demand occurs an order will
be placed. Assume that when an order is placed the inventory position is
brought up to a level 22*. If the inventory position is r* + 1 immediately
after a demand occurs, it is certain that an order will be placed after the
next demand occurs. Similarly, if the inventory position is r* + 2 imme¬
diately after a demand, it will be r* + 1 after the next demand if only a
single unit is demanded and will be 22* if two or more units are demanded,
since in this case an order will be placed. Thus
Z(r* + 1) = /(r* + 1) + a[A + Z(22*)]
(8-73)
+ aC 2 (j + l)p(j)
(8-74)
SEC. 8-9 EXPLICIT SOLUTION IN GENERAL CASE 389
(8-75)
where P(j) is the complementary cumulative of p(j).
Let us now solve explicitlyfor Z(R*) by substituting (8-74) into (8-75),
then Z(R — 2) into the resulting expression, etc. On substitution of (8-74)
into (8-75) we obtain
and
Xi Qc) = apik), k = 1, 2,. . .
and
Xn(0) = 1, n = 1, 2,. ..
The quantity X«(fc) is almost the n-fold convolution of ap(k) as defined in
Sec. 3-9. It differs, however, in that we sum from j = 1 in (8-78) rather
than from j = 0, and that we use the artificial convention Xn(0) = 1,
whereas in reality, ap(0) = 0, because when a demand occurs at least one
unit must be demanded. Intuitively, Xn(k) is the probability, suitably
discounted, that k units are demanded in 1 or 2 or ... orn demands. Thus
a X(h)P(R* - r* -h)
h=0
R*— r* — 1 r- 00 —|
~R*—r*—l
+ 0 £ X(h)P(R* — r* — h) [A + Z(R*)] (8-81)
- h=0 J
or
Z(R*)
itT—r- —i
8-10 Explicit Solution of the Functional Equation for the Steady State
Periodic Review Model
The material developed in the last few sections suggests that it should also
be possible to solve explicitly the functional equation for a periodic review
system. We shall show that this can indeed be done. This in turn will
provide another means for computing the average annual cost of an
(R} r, T) model. It will, of course, be the same expression that was obtained
in Chapter 5. In order to solve explicitly the functional equation, it must
be formulated so that the £ in Z(£; T) is the inventory position at a review
time after any order has been placed, not before the placing of an order.
392 DYNAMIC PROGRAMMING USED IN SS MODEL ANAL CHAP. 8
Let us then define Z(£; T) to be the discounted expected costs over all
future time when an optimal policy is followed for all future times, and
we begin at a review time with the inventory position being £, after any
order has been placed. As usual, we do not include any carrying or back¬
order costs incurred until a lead time has elapsed. Then it is clear that the
functional equation becomes
Z((; T) =M)T)
which has precisely the same form as (8-40). We have omitted the J which
appears in (8-4), since it has no influence on the computations for a given T.
Now, of course, p(x; T) is the probability that x units are demanded in a
period, and /(£; T) is the expected cost of carrying inventory and of back¬
orders in the period from t to r + T, discounted to time 0, if the inventory
position is £ at time 0 (the review time) after the placing of any order.
Assume that an Rr policy is optimal. Then if the inventory system is
operated optimally, the inventory position at a review time after any order
has been placed will only have the values r* + 1, r* + 2,. . ., R*. Thus,
for these values, (8-84) becomes
Z(r* + 1; T) = f(r* + 1;T) + op(0; T)Z(r* + 1; T)
+ aP{ 1; T)[A + Z(R*-, T)]
+ aC ^ (x+l)p(x-,T) (8-87)
x = R*-f*-l
SEC. 8-10 EXPLICIT SOLUTION-PERIODIC REVIEW 393
aA ^ ^2 o"p<n)(A; T)P(R* — r* — h; T)
71 = 0 h — 0_
Z(B*; T) = oo i2* —7-*—1
1 - a J2 J2 a”pM(h-, T)P(R* - r* - h; T)
n=0 h=0
oo 1
We leave for Problem 8-14 the evaluation of the other Z(r* + k; T),
k = 1,. . ., R* — r* — 1. In Problem 8-15 we ask the reader to show that
lim iZ(r* + k; T), k — 1,. . . , R* — r*
x—>0
is the same for each k} and hence the average annual cost is this common
limit.
The above results can be used to obtain an explicit expression for the
expected annual cost. To do this we as usual define a function R, r, T)
to be the discounted expected cost at a review time if an Rr policy is used
with the critical numbers R and r (note that R and r can be different than
B* and r*), and if the inventory position is £ after taking whatever action is
appropriate. It then follows that equations (8-85) through (8-88) still
hold if Z(£; T) is replaced by 2(£; R, r, T) and the asterisks are removed
from R and r. Thus an explicit solution for d(R; R, r, T) is given by
394 DYNAMIC PROGRAMMING USED IN SS MODEL ANAL CHAP. 8
In this way one can obtain an explicit expression for the expected annual
cost of a periodic review system using an Rr policy. Problem 8-16 requires
the proof that the result obtained here is the same as that obtained in
Chapter 5.
Let Z(£, Wij, wm) be the discounted expected cost over all future
times when an optimal decision is made at time £ and at all future times,
and the time between reviews is T. Here we shall begin computing carry¬
ing costs and the cost of lost sales at time £, and the delay of a lead time
will not be introduced. The task of setting up the functional equation for
Z is not quite so straightforward as in the backorders case. Note first that
when lead times are constant, only the order placed at time £ — mT will
arrive between £ and f + T. Suppose that this order for a quantity
v>m > 0 arrives at time f + i. Let pi(xi) be the probability that xi units
are demanded from £ to f + t and p2(x2) be the probability that x2 units
are demanded in the interval £ + i to f + T. If xx < £ and x2 < wm
+ £ — %i, the on hand inventory at time £ + T is £ + wm — X\ — x2. If
xi > £ and x2 < wm, the on hand inventory at time £ + T is wm — x2.
Finally, the on hand inventory at time £ + T will be zero if x\ < £ and
x% > wm + £ — xi or if xi > £ and x2 > wm.
REFERENCES 395
Wm 00
+aL E
X2 = U>m+£~2l-f 1 Xl=0
E pi(x1)p2(a:2)
oo oo —j
where /(£, tum) is the expected cost of carrying inventory and of lost sales
incurred from f to f + T (not f + rtof + r+ T). We ask the reader in
Problem 8-17 to write out /(£, wm) explicitly.
Z is now a function of m + 1 variables rather than just a single variable
as in the backorders case. None of the methods used previously can be
applied to the functional equation (8-92). Indeed, it does not seem easy
to use it either in discussing the optimality of an Rr policy or as an aid in
obtaining the average annual cost. We shall not attempt to carry any
farther the analysis of the lost sales case. In Problem 8-18 we ask the reader
to obtain the functional equation in the case where no orders are outstand¬
ing at a review time.
REFERENCES
PROBLEMS
8-1. Write out explicitly Et(£ + Q; T), Brit + Q) T) defined in Eq. (8-1)
for the case where the number of units demanded in any time interval
is Poisson distributed.
8-2. For the model discussed in Sec. 8-2, prove that if A = 0, then Z„(£; T)
and Z(£; T) are convex, thus implying that an order up to R policy
is optimal. Carry out the proof directly, and do not simply specialize
the A-convex arguments to this case.
8-3. When f(y) is differentiable everywhere, show the equivalence be¬
tween the definition of a convex function given in Problem 2-6 and
that represented by Eq. (8-15).
8-4. Prove the four properties of A-eonvex functions listed in Sec. 8-4.
8-5. Prove that when f(y;T) is a convex function which is everywhere
differentiable, then Fn(y, T) defined by Eq. (8-11) is continuous and
differentiable everywhere. To do this, first show that Fi(y; T) is
continuous and differentiable everywhere. Then show that Zi(£; T)
is continuous everywhere. Thus show that Fi(y, T) is continuous and
differentiable everywhere, etc.
8-6. Consider the definition of ri(x) for x > 0 given by Eq. (8-36). Show
that if units are demanded one at a time
+ . . . + ap(R* — r* — 1; T)f(r* + 1; T)
+ . . . + aY2)(R* - r* - 1; T)Z(r* + 1; T)
+ a*Cp(l;D £ U* b T)
gm —I
+ «S(>(2; T) f] IX b 2)pu-; Tl f . . .
PROBLEMS
OF PRACTICAL APPLICATION
The old adage that there can be many a slip from the cup to the lip seems
to be especially descriptive of the problems encountered in any attempt
to implement an operating doctrine obtained by analysis of a mathematical
model in a real world inventory system. In this chapter we wish to turn our
attention from the main theme of this work, which has been the con¬
struction and analysis of mathematical models of inventory systems, to
problems of implementation, i.e., of practical application. We do not in¬
tend this to be a set of instructions on how a model should be implemented,
since it is not clear that such a set of rules could be devised to cover all
situations, and even if it could, more than a single volume would be re¬
quired to discuss these rules. Rather, we would like to point up the sorts
of problems that can arise, and insofar as possible, suggest ways to avoid
them or solve them. Unfortunately, however, there are no easy solutions
to most of these problems, and in a fairly large number of cases, no satis¬
factory solutions at all have been devised. We shall find it convenient to
discuss the various problems under the headings of: the relevance of the
model, data problems, multi-item problems, personnel problems and pro¬
cedural problems, and problems of evaluation. If nothing else, this chapter
should make the reader aware that a successful application will require
more than the formulation of an appropriate mathematical model.
to emphasize this point, and yet, the authors have encountered many
instances in practice where attempts have been made to use models m
situations where they were completely inapplicable, simply beca^m
model was available for use. One rather obvious misuse of a model which
the authors have encountered was in a military supply system where an
^t was made to apply a steady state model of the ^cussed m
Chapter 4 to spare parts for a military aircraft which would be pdased n
and out in a relatively few years. The model was completely inapplicable
since the mean rate of demand was continually changing with time, and
since explicit account should have been taken of the fact that the parts
would be obsolete when the aircraft was phased out. A more appropriate
model would have been some form of dynamic model such as those dis¬
cussed in Chapter 7. . , ,
The inadequacy of a model for some particular application is not always
so obvious as that referred to above, but nonetheless can lead to equally
disastrous results if applied. A good example of this concerns an apphcaUon
made in that part of a military supply system concerned with stockage of
electronic components. A large number of different electronic components
were carried in stock. The basic idea was to use a steady state model such
as the one described in Chapter 4, to set order levels and reorder points for
each of the items. Previously, the system had been operating using periodic
review. This in itself was reasonable enough, since the mean rate of demand
for many of the items was essentially constant over a fa r period of time,
and since the system had installed a digital computer which made it possible
to use what was, for practical purposes, a transactions reporting system.
However, it was recognized that the order quantities could not neces¬
sarily be used directly, since each year only a certain sum of money de¬
termined by the budget could be spent on procurements Hence it was
necessary to take into account, in one way or another, the budget con¬
straint on procurement expenditures. At this point, the following pro¬
cedures were attempted. It was decided to choose the order quantities Q,
and reorder points r* for each of the n items, by minimizing the average
annual cost X for all n items, in such a way that the expected procuremen
costs would not exceed the budgetary limitation. To dothis, the cost of a
backorder for unit j was assumed to have the form BtVCh where C,»the
unit cost of the item, t is the length of time for which the backorder exists,
and 0 is a parameter independent of j, to be determined in such a way that
the budget constraint was to be satisfied. Note that in this model there is
no fixed cost of a backorder. (The reasoning which led to the choice ot
= eVcTj is somewhat obscure, but is probably irrelevant, since the same
problems would have in all likelihood been encountered if was taken
equal to 0/3- for any arbitrarily selected number J>) Then the function to
hft mrniinized was
SEC. 9-2 RELEVANCE OF THE MODEL 403
where all quantities not defined above have their usual meaning. The
minimization was carried out for various values of 8 yielding X*(8), Qf(8),
r*(0)> j = 1,... ,n. Then 8 was chosen so that the expected procurement
expenditures were equal to the sum available in the budget.
It happened that, in the particular year in which the model was intro¬
duced, the budget was extremely tight. Because of this, 8 turned out to be
a very small number, so that the cost of a unit year of shortage was small.
For example, on some tubes, the cost of a unit year of shortage came to the
absurdly small value of $0.75. The reorder points determined in this way
were quite low. In many cases, they were so low that the safety stock was
negative (i.e., r,- was less than yty). The r, were in general considerably lower
than the on hand inventory at the beginning of the year. On the other
hand, the Qj was not especially small, since in the model, the contribution
of the Tj + Qj terms to the backorder cost were ignored, and thus Qj could
not be less than the Wilson Q.
The Qj and r,- values obtained from the above model were used in the
real world system. What happened was as follows: The system began
using up the on hand inventory for all items, and when (the very low)
reorder points were reached, orders for the rather large quantities Qy were
placed as needed. Before the end of the year, the entire budgeted sum had
been spent, and the system was about to run out of stock on many items—on
some because an order had been placed but the safety stock was negative,
and on others because the reorder points had to be passed because no funds
were available. At this point a crisis was reached which required special
legislation to obtain emergency funds for restocking through the sending of
expedited procurement orders to suppliers.
The application just discussed provides an excellent example of a case
where the results obtained using a mathematical model were much worse
than would have been obtained using the standard procedures that were
already in existence. It was entirely inappropriate, of course, to attempt
to apply a steady state model to a situation where there was a fixed a.nmia.1
procurement budget, and even worse to do it on the basis of introducing a
constraint on expected expenditures where the backorder cost was varied
to bring expected expenditures in line with the budget (since varying the
backorder cost did not have a great influence on the Qj). The proper
procedure for operating a system in the face of a tight budget is very com¬
plex. One must in general use hand to mouth buying (i.e., buy in very
small quantities), and insofar as possible attempt to postpone as many
procurements as possible until right after the beginning of the next budget¬
ary year when more funds will be available. None of the models developed
PROBLEMS OF PRACTICAL APPLICATION CHAP. 9
404
in this text are capable of handling such a problem and, indeed, no such
models have been developed. It is clear that on the average, the budget
must be sufficient to procure the average annual demand, or the system
cannot continue to operate. However, by clever manipulation, it can
survive particular years with tight budgets.
There exist many other ways in which a model can be improperly applied.
We shall now briefly mention several of these. For example, consider a
situation in which the mean rate of demand is fairly low while the number
of units demanded per demand varies widely. Here it would be very
dangerous to use the steady state model developed in Chapter 4 for Poisson
demands, even if the mean rate of demand remained constant over time,
since the variance in the order size could cause the safety stock determined
from the model to be much too small. Other misapplications have involved
attempts to apply a steady state model to a department store item, 50 per¬
cent of which was sold during the Christmas season, or the attempt to use
results computed from a transactions reporting model in a periodic review
system.
The above have illustrated actual cases where models have been mis¬
applied. The reader should not infer from the above that a model will never
be satisfactory unless it replicates very closely the behavior of the real
world. For example, it may be perfectly satisfactory to apply a steady
state model to a situation in which there is a moderate amount of season¬
ality in demand, i.e., the additional savings which could be obtained by
using an operating doctrine based on a dynamic model over the costs in¬
curred by using a {Q, r) model with a fixed Q and r (computed using the
average demand rate) may not justify its use, perhaps because it is too diffi¬
cult to get the clerical people to use a changing Q and r. It is quite easy to
obtain a rough estimate of the differences in ordering and holding costs ob¬
tained by using a fixed Q and one that varies with the season. All one needs
to do is compare the results of the simple deterministic lot size model of
Chapter 2 with the deterministic dynamic lot size model of Chapter 7.
Sample calculations quickly reveal that even with sizable seasonal patterns,
the cost differences can be very small. Thus in many instances the use of
a constant Q could be easily justified. It is somewhat more difficult to
determine whether or not it is reasonable to keep r constant at the value
determined from the steady state model also (i.e., have a constant safety
stock). This depends in part on the relative magnitude of the stockout
costs and carrying costs. A natural way to investigate whether or not the
use of a constant r value determined from the (Q, r} model would be
satisfactory would be to use simulation.
Simulation is a very useful tool for studying how any operating doctrine
can be expected to behave in the real world system. In particular, simu¬
lation is helpful in studying what sort of behavior can be expected when
SEC 9-3 DATA PROBLEMS 405
All inventory models require some information concerning the demand for
the item under consideration. In the simplest deterministic model only a
single parameter, the demand rate is needed. For other models it may be
necessary to determine the nature of the process generating demands, and
the nature of the distribution of the quantity demanded per demand. It may
even be necessary to predict how these will change over some future time
period. In order to make even the simplest estimate of the demand rate,
historical data are needed. It is surprising how often one finds that no
demand data whatever are available in usable form, and hence, before any
attempt can be made to apply the operating doctrine obtained from the
mathematical model, it is necessary to collect such data. In essentially all
cases, the data available will be sales data, and these will be equivalent to
demand data only if no sales are lost. When lost sales are possible, then
sales data will not yield precisely demand data. Even in the case where all
demands are backordered, sales data can distort the picture if they are
based on the time of filling an order, because, when some demands must
be backordered, the time pattern of filling orders will be different from the
time pattern of the original demands. One normally has no alternative but
to use the sales data as demand data. In the lost sales case, one could, if
the average fraction of the time out of stock was known, correct the mean
SEC 9-4 THE DEMAND DISTRIBUTION 407
rate of demand obtained from sales data by dividing by the average fraction
of time for which the system has stock on hand. Generally, however, the
fraction of the time out of stock is not available and is difficult to measure.
Fortunately, it is often small and hence not too much of a problem.
Let us now assume that some sort of demand (i.e., sales) data are avail¬
able, and let us then see what problems arise in attempting to use these to
obtain the parameters of the demand distribution needed in the model.
First, we shall study the case where the model being used assumes that the
stochastic process generating demands does not change with time. Before
any attempt at parameter estimation is made, one should examine the
situation to determine whether or not the assumption of a constant mean
rate of demand is reasonableT"Usually, in the real world, the mean rate of
demand will not be strictly constant, and what one really wants to decide is
whether over a suitable time interval (perhaps a year or so) the mean rate of
demand is essentially constant. It is very difficult to make this decision on a
quantitative basis since this would require comparing the cost using an
operating doctrine obtained from a model which assumes a constant de¬
mand rate and those using an operating doctrine obtained from a dynamic
model. To make such comparisons could require a great deal of work, and
might even require the use of a large computer to make a detailed simula¬
tion. To do all this is usually much too time consuming and expensive (and
even if it was done the results might be misleading because of the difficulties
in predicting how the mean rate of demand would change with time for use
in the dynamic model). Instead one makes the decision on a very qualitative
basis, often without a detailed study of the data. Frequently, it is sufficient
to know what sort of item is being dealt with. If it is a staple item, which
does not have an extreme seasonal pattern, and for which obsolescence is
not an immediate problem, one will usually use a model which assumes that
the mean rate of demand is constant, simply because the probable savings
that could accrue from using a more complicated dynamic model do not
justify the relatively high costs of its use. If the mean rate of demand is
changing somewhat with time, one can simply recompute the operating
doctrine at some regular interval, such as once a year. On the other hand, if
the item is a high fashion item, which is only sold during one period of the
year, or is an item which is to be phased in and out over a relatively short
period, it will frequently be ill advised to use a model that assumes the
mean rate of demand to be constant.
If one wants to determine the behavior of the mean rate of demand over
time in more detail, it is convenient to use a control chart type of analysis
in doing so. As a measure of the demand rate one will use the demand per
time period, i.e., the daily, weekly, or monthly demand. This will be
plotted on a graph as a function of time. Using the demand per time period
data, one will then compute from the data available the average demand
408 PROBLEMS OF PRACTICAL APPLICATION CHAP. 9
per period 1 and the standard deviation of the average demand jper period
Sd- On the graph where the demand data are plotted, lines for d, d + ZSd
and d — 3Sd are also plotted. This may yield results such as those shown
in Figs. 9-1 and 9-2. The chances that a point will lie outside the control
limits d + ZSd or d — 3Sd are very small if the mean rate of demand is
indeed constant.
If the data behave like those shown in Fig. 9-1, then statistically there
is no reason to believe that the data were not generated by a stochastic
process in which the mean rate of demand is constant. On the other hand,
if the data behaved like those in Fig. 9-2, it seems clear that the mean rate
of demand is increasing with time. In both cases, however, one might use
an operating doctrine obtained from a model which assumed that the mean
rate of demand was constant. Indeed one could use graphs such as Figs.
9-1, and 9-2 to indicate when it was desirable to change the parameters in
the model and recompute the operating doctrine. The recomputation (or
at least an investigation of whether or not a recomputation should be made)
would be triggered by having a point fall outside the control limits. For
such purposes, however, it would not necessarily be desirable to have the
control limits lie at d + 3Sd and d — ZSd- It might be noted that the sort
of results one obtains on graphs like Fig. 9-1 can be strongly dependent on
the time period chosen, i.e., whether one uses the daily, weekly, or monthly
demand. For example, the daily demand on Mondays may always be very
low while it is always very high on Fridays. Hence for this unit of measure,
a control chart would indicate that the mean rate of demand was not
SEC. 9-4 THE DEMAND DISTRIBUTION 409
constant. On the other hand, the mean rate of demand on a weekly basis
might be quite constant. The interval should be chosen to be as large as
possible consistent with the data available, and any other requirements
such as those that will be discussed later. It should not, however, be larger
than either the lead time or the average time between the placement of
orders.
We shall now turn to the details of estimating the demand parameters
or the demand distribution needed in the model. If the model is a determin¬
istic one which requires only the mean rate of demand, we would divide
the total demand, over the longest possible time period for which it is be¬
lieved that the demand rate is representative of the current demand rate,
Weekly i
demand
•
•
J, oo
CL + uOjjr
•
• # *
d -.-.•. '■ 1 ■
•
d-SSd •
•
0 Time
FIGURE 9-2.
useful data into intervals of length r. For each such time interval we de¬
termine how many demands occurred. Then we construct a histogram by
finding the fraction of the time intervals which have demands lying in each
of some suitably chosen demand intervals of length Ax. When plotted
graphically one might get a result such as that shown in Fig. 9-3.
Unfortunately, the procedure just suggested often cannot be used. The
reason for this is that if r is fairly long, only a very small number of time in¬
tervals of length r will be available in the time period over which useful data
are available. In order to use the above technique, one should have some¬
thing like at least 50 intervals of length r in the demand history, if one is to
obtain a histogram which contains any detail. Often this cannot be done.
FIGURE 9-4.
Central to any attempt to use a dynamic model is the procedure used for
making predictions or forecasts. The nature of the predicting technique
used can vary widely. It may, for example, involve only the use of histor¬
ical data on the item itself, it may involve predictions on the general state
of the economy, or it may be based on future planned requirements (in the
case of spare parts, etc.).
We shall not discuss prediction techniques in detail, since different
circumstances require widely varying approaches. We shall, however,
briefly examine several of the more simple ones. Consider first the problem
of predicting the demand for things like spare parts, which depend on the
usage of some other piece of equipment. This is an important problem in
any military supply system. To be specific, imagine that we are attempting
to predict demand for a spare part for a certain aircraft over a given future
period of time. The typical procedure which has been used in the military
to do this is to establish, on the basis of historical data, a usage rate, given
as the average number of spares needed per flying hour. Then the total
number of expected flying hours for all the aircraft of the type under con-
414 PROBLEMS OF PRACTICAL APPLICATION CHAP. 9
In the above, dj is the predicted demand for period j and dj is the actual
demand for period j. Assume that a, b are to be determined by using the
demands in the previous N periods. We shall imagine that the current
time £ is a review time, and that the time period from ttot — T,T being
the time between reviews, will be referred to as period 0, the period from
t - T to t - 2T as period -1, etc. Thus the numbers of the N periods
to be used in determining a and b will be 0, — 1, . . . , —-(A — 1). To
determine a and 6, we set dF/da = dF/db = 0 and solve the resulting
equations for a and b. Then from (9-1)
„ -(V-1)
7- = -2 L j(dj - aj - £>) = o,
= -2 L — aj - b) = 0
— (JV—1> 1
. -N(N - 1) ; V—' J-
L ji = ^ N(N — l)(2iV — 1)
L 3= 2 ’ j=0 D
i=0
Also, let
— UV— l)
U= L v= £ 3di
r JV - 1 TT1 , U , (N - 1\
N(N -1)(N + 1)
416 PROBLEMS OF PRACTICAL APPLICATION CHAP. 9
SO that
$=f+
u + 12 V + ^-^U
[j + ^y-1) i = 1) 2, 3,. . . (9-4)
N
N(N - 1 )(N + 1)J
Equation (9-4) tells how to compute the predicted demand for period j,
j = 1,2,, using data for the demands in periods 0, — 1,. . ., — (N — 1).
Note that period 1 occurs between t and t + T, period 2 between t + T
and t + 2T, etc. Observe also, that U/N is simply the average demand in
periods 0, — 1,. . ., — (N — 1), so that (9-4) indicates that the predicted
demand is the average over the past N periods plus a trend correction.
In (9-4) only U and V depend on historical data. It is unnecessary to
recompute U and V from their definitions each time a new prediction is
made. Instead, it is easier to compute the new values of U and V from the
previous values. Let & and V’ refer to the values of these variables com¬
puted at time t + T, i.e., using di, do,..., d-(N-2), and U and V the values
computed at time t, i.e., using do, d-i, . . ., d-w-1>. Then it is easy to see
that
ty = + U - d-(w_ i); V = v-U + Nd-ur-1) (9-5)
It is much easier to compute t) and V from (9-5) rather than directly from
their definitions.
By use of (9-4) and (9-5), it is quite simple, either manually or on a com¬
puter, to predict the demand in each of the future periods needed and also
to predict (at least approximately) the lead time demands needed.
The uncertainty in the predicted demand will increase with j, i.e., the
farther one moves into the future. It is desirable when possible to get some
estimate of this uncertainty as a function of j (the number of periods in
the future for which prediction is made). If sufficient historical data are
available, this can be done by computing a histogram for the distribution of
forecast errors. This is accomplished by using (9-4) to predict the demand
j periods in the future and then comparing it with the actual demand.
The difference between the actual demand and the forecast demand is
called the forecast error. By doing this for as many historical points as
possible, one can obtain a histogram of forecast errors for each value of j.
Let Sf be the variance of the distribution of forecast errors when predicting
j periods into the future. If the dynamic model uses a theoretical distri¬
bution such as the normal, then for period j one can use Sf for the variance
of the period's demand. Similarly, be adding the appropriate variances,
one can at least roughly estimate the variance of the lead time demands.
We gave no indication above as to how to choose N} the number of past
SEC. 9-5 DEMAND PREDICTION All
periods demand that are used in making the predictions. One way to select
N is to pick the N which minimizes the variance of the distribution of
forecast errors for a given /, or minimizes some combination of the variances
over all /. This could be a tedious computation to make manually, but it
could easily be carried out on a computer.
One difficulty with using a least squares line for prediction is that it is
always necessary to have available the demand N periods back. This can
take up an unnecessarily large amout of storage in a computer when many
items are being handled. The exponential smoothing method of predicting
eliminates this problem. Given any time sequence of data with/,- being the
value of the variable for period /, then the smoothed value of the variable
for period /, denoted by /,-, is
U = afj + (1 - «)//_i, 0 < a < 1 (9-6)
i.e., the smoothed value for period / is a times the value of the variable for
period / plus (1 — a) times the smoothed value for period / — 1, where a
is a positive number less than unity. Since /,- involves /,*_i, it is clear that all
previous data are included in obtaining /,-, i.e., from repeated substitution
(9-9)
a
418 PROBLEMS OF PRACTICAL APPLICATION CHAP. 9
1 — a
P
a
dj = do + ~ J po (9-12)
/ = 7 j[ e-y<f-nf(r)dt (9-14)
remarkably constant from one season to the next. In this way, the total
sales for the season can be predicted with some accuracy after only the first
few weeks. At least one very large retail chain makes use of these seasonal
percentages to make early estimates of the total season’s sales for purposes
of placing additional orders, or for markdowns.
It was pointed out in the previous section that it is often true that demand
data needed in the application of an inventory model are not available.
This is even more true with lead time data, especially if lead times are
fairly long and orders are not placed too frequently. It is rare indeed when
sufficient data are available to yield a detailed histogram for the lead time
distribution. Sometimes, about the best one can do is to obtain crude
estimates of what the maximum and minimum lead times are. In such a
case, about all that can be done is to average them to yield the mean lead
time, and to estimate the standard deviation to be the range divided by
six (since often six standard deviations, three on either side of the mean,
will include essentially all of a density function).
In practice many other problems may arise with lead times besides those
of trying to estimate the lead time distribution. There may not, in fact,
be any stationary lead time distribution, since lead times may be continu¬
ally changing with time. Furthermore, it may turn out on occasions that
orders are split and the entire order is not shipped at one time. In addition,
there may exist the possibility of expediting orders if an out of stock con¬
dition appears imminent. Although some of these additional complications
can be included in the mathematical model, it usually turns out that the
model becomes so much more difficult to work with and requires so much
additional data which are hard to obtain, that it is not worthwhile to
attempt to include them rigorously.
We have noted in the previous section the difficulties involved in at¬
tempting to use an empirically determined demand distribution in a mathe¬
matical model. The difficulties are compounded, of course, if one attempts
to use both empirically determined lead time and demand distributions,
since then one must generate numerically the marginal distribution of lead
time demand. This in turn requires that the demand distribution for each
possible lead time must be found. It is only in very special cases that this
marginal distribution can be determined with sufficient accuracy to
warrant the effort required to obtain it. One can, however, use an empiri¬
cally determined lead time distribution with a theoretical, discrete, demand
distribution such as the Poisson, for if it is assumed that the lead time r
can only assume one of a finite number of values U with probability Z(k),
420 PROBLEMS OF PRACTICAL APPLICATION CHAP. 9
The types of costs which are relevant in working with inventory models
have been discussed in Chapter 1. These costs include the costs of the units,
the fixed ordering costs, carrying costs, stockout costs, and the cost of
operating the information processing system. All these costs can be difficult
to determine in certain cases, and some are almost always difficult to
determine. Regarding this determination, there is not a great deal that
can be said to be really useful in a wide variety of situations; so we shall
only make a few brief remarks.
Stockout costs are often the most difficult to determine. They cannot
normally be measured directly, since they usually include such intangibles
as good will losses. Consequently, the usual procedure, if stockout costs
are specified at all, is for someone to make a guess as to what they are.
Fortunately, as we have illustrated previously by example, the optimal
policy tends not to be very sensitive to these costs, and hence an estimate
that is of the right order of magnitude will often suffice. In the text it was
SEC 9-7 DETERMINATION OF COSTS 421
world where discounts of this sort are important and must be taken into
account. Fortunately, however, there is another type of quantity discount
which seems to be more popular, and which is also easier to handle. This
discount is based on the total quantity purchased per year—not on the size
of any individual order. For situations where the mean rate of demand is
fairly constant, the average amount procured per year will be fixed inde¬
pendently of the operating doctrine (at least for the backorders case, and
also for the lost sales case if the system is seldom out of stock) so that the
average discount is predetermined. In this case one can treat the unit cost
as a constant, independent of the operating doctrine. The unit cost should,
however, be modified to reflect the discount.
Finally, let us consider the costs involved in operating the information
processing system. In general, these costs will include all the costs associ¬
ated with maintaining inventory records, including wage costs, computer
costs, material costs, etc. Also included will be the costs of making physical
inventory counts and the cost of making demand forecasts. In the text, it
was assumed that the cost of a review for periodic review systems and the
cost of operating a transactions reporting system were independent of the
operating doctrine. This need not- be true when the parameters of the
operating doctrine are varied over a wide range. If one could determine
precisely how these costs varied with the parameters of the operating doc¬
trine, these appropriate functions could be included in the cost expression.
Generally, however, it is adequate to assume that they are independent of
the operating doctrine.
text have clearly indicated that different items should be treated differ¬
ently, depending on the nature of the costs and stochastic processes in¬
volved. However, it can become impossibly expensive if one attempts to
develop and use sophisticated operating doctrines on each of 100,000 items.
For example, if 10,000 items were being controlled, one might need a very
large computer. If 100,000 items were being managed, several of the largest
computers available might be needed. Just the operation of several com¬
puters could easily amount to an expense of two or three million dollars
per year. The answer to this problem lies in dividing the items up into a
number of groups, with items in the different groups being treated differ¬
ently.
Recent studies in both the military, retail stores, and industry have all
reached the same interesting conclusion that, in general, a very small
fraction of the total number of items stocked account for a very large
fraction of the dollar volume of business involved. Frequently this will be
something like ten percent of the items, accounting for eighty to ninety
percent of the dollar volume. These studies have led many large inventory
systems to change the way in which they control the items they stock.
Now, the items are broken down into several categories, usually three, and
items in different categories are treated differently. The items in the three
different categories might be referred to as high, medium, and low value
items. In the military, for example, the high value items would be con¬
trolled very closely, using the best means available. For items with a
relatively constant demand rate, transactions reporting might be used along
with a (Q, r) model. For items with strongly time dependent demands, the
best available dynamic model might be used. In the case of medium value
items, somewhat less costly control procedures would be used. For example,
periodic review might be used for all items, with only two or three possible
review intervals allowed. Here also, an effort would be made to account
for the nature of the demand distribution and the costs in determining the
safety stocks to be kept. In the case of low value items, no attempt would
be made to use a sophisticated operating doctrine. All items might be
reviewed once per year. The safety stock for all these items would be more
or less arbitrarily set at k months of supply. By use of a control system of
this sort, one keeps down the cost of control while still doing a good job of
controlling the most important items. A similar sort of procedure might be
used in a department store. In this case, however, periodic review would
probably be used for all items, with shorter review periods for the high value
items. The review periods might range from one day to two weeks for high
value items, one week to one month for medium value items, and two weeks
to six months for low value items.
We have previously suggested that there will often be interactions be¬
tween items carried by the inventory system. Almost any interactions
SEC. 9-9 PERSONNEL AND PROCEDURAL PROBLEMS 425
stand them. Frequently, one will have to break down the details for using
them to the point where even the color of pencil to be used in making
certain notations is indicated.
Some examples, taken from actual real world situations, may help to
emphasize some of the above points. One large department store instituted
a lot size-reorder point operating doctrine for a great many items in the
housewares department where the mean rate of demand is reasonably
constant over time, with the exception of the Christmas season. Initially,
the buyers were strongly opposed to the introduction of such a system.
Usually, the buyer in a department store has a great deal of freedom in
determining what to buy, when to buy, how much inventory to carry, when
to take markdowns, etc. The buyer himself makes all these decisions for
the items under his jurisdiction. With the new system, the buyer had no
control over how much to stock or when to place orders. In fact, all the
ordering decisions which were previously made by the buyer were, under
the new system, made by clerks in the accounting office which previously
had nothing to do with these decisions. The buyers no doubt opposed the
introduction of the new system partly because they were afraid that their
jobs were being downgraded. Probably another reason that they opposed
it was because they were afraid of what would be discovered. On making
checks prior to introducing the system, all sorts of shocking things came to
light, such as the appearance of a fifteen years7 supply of some items on
hand, the stocking of a huge number of different brands of a particular item
such as coffee pots (in different parts of the store, without one buyer know¬
ing what the other was doing), items that could no longer be sold because
they were obsolete and the buyer had not taken a markdown at the proper
time, etc.
Initially, things did not go at all smoothly on introducing the new system.
The greatest difficulty was in getting people to do the right things. The
clerks would mix up stock numbers and count the wrong items, or they
would not make the counts correctly. In the accounting department,
orders would not be placed when the reorder points were reached, the
wrong item would be ordered, or the wrong quantity would be ordered.
In time, these difficulties were corrected to a considerable extent. Further¬
more, most of the buyers did become more enthusiastic about the system
since it relieved them of a considerable amount of work and, in addition,
made it possible for them to blame someone else if things went wrong.
The operation of some military supply systems provides an excellent
example of situations where the system of rewards and punishments for
the operating personnel tends to make them behave in a way completely
contrary to the behavior desired. Just like the buyer in a department store,
the base supply officer at a military installation has considerable freedom
in deciding how much of various items to stock despite rules and regulations
SEC 9-10 THE EVALUATION PROBLEM 427
which, theoretically, determine what the inventory levels should be. The
supply officer’s commanding officer is the line officer in command of the
base. The main concern of the base commander is that all planes should
be in flying condition, or that all ships be seaworthy and with all systems
operational. Consequently, if a plane is grounded because there is a stock¬
out of some spare part, the base supply officer will be heavily penalized
and may receive a black mark on his record. On the other hand, there are
no equivalent demerits for keeping too much inventory. Thus, the logical
thing for the supply officer to do is to keep as much of everything on hand
as possible. This is precisely what he does do, and this leads to much higher
stocks at bases than are really needed.
Another example of this same sort of phenomenon occurs when some
part is in short supply throughout the entire military supply system. Then
the item is placed on what is called a critical list, and all base supply
officers are supposed to report how much they have of this item, so that
if someone needs it desperately, the item can be shipped from one base to
another. What actually happens when an item is placed on the critical list
is that all base supply officers attempt to hoard the item, and will not part
with any of it under any circumstances. Thus the situation is really made
worse rather than better.
To get some feeling as to the problems that can be involved in attempting
to write down a precise set of instructions as to how a system should be
operated, one only need look at a set of military supply manuals. These
consist of a set of thirty or more volumes that presumably state in detail
precisely how the system should be operated. However, on attempting to
determine what should be done by reading these manuals, one soon becomes
hopelessly confused. It is essentially impossible to discover from them how
the system should be operated. This is in part due to the complexity of the
whole system, and in part due to the way in which the manuals are written.
No one person knows enough about the whole system to say how it should
be operated. Different people write different volumes in the set, and,
consequently, this makes it difficult even to have them consistent, let alone
coherent.
and they will make every effort possible to cover up any defects in the
system. On the other hand, if it really is a spectacular success, then in some
industries the management will also attempt to keep the entire operation
a secret from the outside world and their competitors.
9-11 Summary
REFERENCES
PROBLEMS
9-1. The daily sales of a certain item in a department store have been
observed for 60 days. It has been found that never more than three
units are demanded and the frequencies with which 0, 1, 2, 3 units
were demanded are 15, 10, 20, 10, 5 respectively. The procurement
lead time is always four days. Determine from the empirical data
the marginal distribution of lead time demand.
9-2. Suppose that in Problem 9-1, one additional day is observed, and the
demand is for 6 units. How does this change the marginal distribution
of lead time demand?
9-3. Consider the demand data given in Problem 9-1. Assume that the
lead time is always either 2 days or 3 days, with the probability of
two days being 0.3. Determine the marginal distribution of lead time
demand.
9-4. The weekly demand for an item has had in the recent past the follow¬
ing values on successive weeks: 25, 10, 15, 16, 9, 30, 17, 18, 8, 25, 26,
14, 12, 9, 15, 18, 7, 11, 19, 15, 28, 17. By use of a moving average
having N = 3 and without a trend correction, estimate the demand
two periods in advance for each week, beginning with the third in the
demand history, and compare with the actual demand. Repeat the
procedure using Eq. (9-4), which includes the trend correction, to
make the prediction. Plot the results obtained, showing the predicted
values and actual demand as a function of time. Compute a histo¬
gram of forecast errors for the case where the trend correction was
included in making the prediction.
9-5. Repeat Problem 9-4 using exponential smoothing with a = 0.15.
9-6. Assume that for all past time the demand per period has been 10
units. Then beginning at time 0, the demand per period begins to
increase at the rate of 2 units per period, so that dy = 10 + 2j. Plot
dj and dj = U/N when N = 4. Also plot dy and dj = dj = adj +
(1 — a)d3-1, a = 0.20. Does the lag dj — dj behave in the way one
would expect it to?
9-7. Assume that for all past time, the demand per period has been 10
units. Suddenly at t = 0, the demand per period jumps to 20 units
and remains constant at that value. Plot dy and dj, where dj is com-
PROBLEMS 431
puted from Eq. (9-4) for N = 4, and the prediction being made for
the next period.
9-8. Repeat Problem 9-7 using Eq. (9-12) with a = 0.2, and again making
the prediction for the next period.
9-9. Solve Eq. (9-13) when
0, f<0
/(f) = 6f, t > 0
ia+ia.^ = 0 or — = - ^0
dXj dxj dg/dxn
JL
3^ _ j9f _ dXn d£_ = 0, j = 1,.. (Al-2)
~ dXj dg dXj
dXn
Now if rt . . ., x* are the minimizing values, write y = -df/dxn/dg/dxn
where df/dxn and dg/dxn are evaluated at x\, ...,xt Therefore the opti¬
mal Xj values must satisfy the n + 1 equations
Let us return to the study of the problem which seeks to find non-negative
values of xu . . ., xn which minimize z = /(mi, ...,Xn) subject to the
constraint g(x1,If 2* is the minimum value of z subject to
the constraint, then z* will in general depend on the value of a, i.e.,z is a
function of a. Similarly, the Lagrange multiplier v will also depend on a,
i.e. v is a function of «. We would now like to show that dz*Jda = -fl
where both sides of the equation are evaluated at the same value of a.
To prove that dz*/da = -v, let x\,. . ., xt be a set of */ values which
yield z*. These x*} values will also be functions of a. Thus
da fzi $x* da
However, because of (Al-3)
dxj V dxf
so
dz*_y' dg dx* (Al-8)
da~ n 34i dx* da
Next we note that the constraint g(xi,. .., a;») - a must always be satis¬
fied (i.e., it must be an identity in a). Hence, differentiating the constraint
with respect to a we obtain
436 APPENDIX 1
y-* $9 dxj __ l
dx*J da (Al-9)
J =1
A2-1 Introduction
x — Xq = Ax = -~f(xo)/(df(xo)/dx) (A2-2)
dfl(,Xn) y-n)
0 = fi(xnj yn) + {Xn-ri Xn)
dx
+ dy
yn) fl(Xn+l} yn+l)
A _ i a/i/dz -A
Aj/„ - yn+1 yn- j d^bx —y2
For most of the models formulated in this text, the optimal operating
doctrine was determined by minimizing some cost expression such as t e
average annual cost X. Newton’s method provided a way of solving equa¬
tions such as dX/dQ = dX/dr = 0, which are necessary conditions that Q
and r must satisfy if they minimize X. The method of steepest descents
works with X directly, and proceeds to minimize X rather than to solve
440 THE METHOD OF STEEPEST DESCENTS SEC A2-3
In the following properties, p(r; p), P(r; p), y(t) a, 0), bw(j; n, p) are
defined by
00
mYe-e*
7it', «, P) r(a + 1)
O' > ~ 1, jS > 0
The properties hold for all m > 0 and all non-negative integers r*. The
differencing operator A{/(r)} is defined by A{/(r)} =/(r) —/(r 1).
T(x) is the gamma function of argument x.
1. rp(r; p) = pp(r — 1; p)
Pm(r) = P L (m a; X) m = X’ 2’ • ■ ■
3- p) = pP(r - p)
y=r
00
Then,
- (r - l)m+1P(Y; m = 1,2,
+ ^P(r-2;M) + ^P(r-3;M)
eo eo 00
16.
Jo
f p(r; Xt) dt = 7 P(r + 1; XT)
X
rpn+l
U\T)=^P(r-,\T)
IS. f*l-P(r;U)dt-±£^P<.* +
_ _J_ / 1 \ (n + r)\
Xn+1 \n + 1 y (r — 1)!
n = 0,1, 2,. .
(x + r)!
P(x + r + 1; XT)]
(r - 1)!
23. f“ y(t; <x, 0)p[r; X(t + T)] * = £ P(r “ P> xT>6" (* ° + lj M^s)
Jo 3=0
n
24. JJ J*q r(*;«>0)p(r; Xt) dT * = x [x “ %blT ij'’a + 1’ x + *)]
t+T
y(t- a, P)p(r; Xt) dr dt
- iA [i
L
- j»o»=o
Z £ rO - <; («■«
v
+ !■ r+l)]
26. A{j*i(r)} = -(r - l)p(r ~1;m) = ~MP(r ~ 2> **)
1
e~rl/2; $(r) = £ 4>{x) idx
1. J x4>(x) dx = <t>(r)
/ * i
xn<3?(x) dx = ———rn+1$(r) + n
1 z*00
± J xn+14>(x) dx,
n = 0,1,
6. y* <£(x) etc = <^>(r) — r$(r)
^ ^ ^ [|^] - I ^ [^]
APPENDIX 4 445
“■ J"'-w'(x’T‘>’ T*T‘>0
, m if ,, sdx . 3Dn. rx-xri
w2(x, T)=^|* + X + X2
X ^ X2 J^Lc^)1/2]
_
_2CP?y/2rT - aoi .nx — xr~i
x2 |_ x2 J L (DT)1,2J
+
1r ox~u
6z2Z> ,, 15xD2
±ojsJL/ , 15Z)3
W.(*, T) = + — + “5?“ + 14^1
14. Let
Rn{x, Ti, T2) = JJ* [|r^] dt, n = 0, 1, 2,... ; Ti, T2 > 0
Then _ XT 1
«.(*, ^1, Ts) = r§+1 $ [4rp] “ ^+1n+1 $ [w4J
and
RJx, Th T2) = - £ Ml Th T2) d,Z
y (x f\ — ir f2 _ — _ $ r ~ ^~i
x.
J> - 2L X2 X3 2X4J L (DT) «* J
, (DT)ll2V,rr , 3D , 1 . fx - xn
+ 2X2 LXT+ X +Xrl(DT)u*_
D_ _ MTI ,n ^ fx + XT1
2X3 [X
i 2XJ'e L (DT)1’2]
Tx + xn [x - xn
18. eMD
*L(Dtyi*] “ *L(Dt)"*]
dMf)
19.
dr
= ~4>(r)
Vx - xn i rx - xn
20. j-$
dx L(D<)1/2J_ (doi/20L(doi/2J
26. j x$ dz
+ | (r2 + Mr + M2 + 2<r2)4> ^^