Probability Book PDF
Probability Book PDF
Mathematical Models
What is a mathematical model?
Observation
Application
Deterministic models
Nondeterministic models
1
• the number of α-particles emitted from a piece of radioactive material in one
minute.
• the outcome of a single toss of a fair die.
• the number of students who will earn an A in EAS305.
Sets
Basic definitions
We will need to use set theory to describe possible outcomes and scenarios in our
model of uncertainty.
Definition 1.1. A set is a collection of elements.
A = {1, 2, 3, 4}
A = {x : 0 ≤ x ≤ 1}
or
A = {0 ≤ x ≤ 1}
A⊆B B⊇A
A = {x : x2 + x − 2 = 0}
B = {x : (x − 3)(x2 + x − 2) = 0}
C = {−2, 1, 3}
Then A ⊆ B and B = C .
Definition 1.6. A is a proper subset of B if A ⊆ B and A = B . We write
A⊆/ B.
Note: Some authors use the symbol ⊂ for subset and other use ⊂ for proper subset.
We avoid the ambiguity here by using ⊆ for subset and ⊆/ for proper subset.
Venn diagrams
Set operations
Complementation
Ac = {ω ∈ Ω : ω ∈ A}
Example: Here is the region representing Ac in a Venn diagram:
Ω
A Ac
Union
A ∪ B = {ω ∈ Ω : ω ∈ A or ω ∈ B (or both)}
A B
Intersection
A ∩ B = {ω ∈ Ω : ω ∈ A and ω ∈ B}
Example: Using the same Venn diagram as above the region representing A ∩ B
is given by:
Note that some textbooks use the notation AB instead of A ∩ B to denote the
intersection of two sets.
Definition 1.7. Two sets A and B are disjoint (or mutually exclusive) if A∩B =
Ø.
These basic operations can be extended to any finite number of sets.
A ∪ B ∪ C = A ∪ (B ∪ C) = (A ∪ B) ∪ C
and
A ∩ B ∩ C = A ∩ (B ∩ C) = (A ∩ B) ∩ C
Set identities
(e) A ∪ (B ∩ C) = (A ∪ B) ∩ (A ∪ C)
(f) A ∩ (B ∪ C) = (A ∩ B) ∪ (A ∩ C)
(g) A ∩ Ø = Ø
(h) A ∪ Ø = A
(i) (A ∪ B)c = Ac ∩ B c
(j) (A ∩ B)c = Ac ∪ B c
(k) (Ac )c = A
Venn diagrams can only illustrate set operations and basic results. You cannot
prove a true mathematical statement regarding sets using Venn diagrams. But they
can be used to disprove a false mathematical statement.
Here is an example of a simple set theorem and its proof.
Theorem 1.1. If A ⊆ B and B ⊆ C , then A ⊆ C .
Proof.
Discussion Reasons
(1) If x ∈ A then x ∈ B Definition of A ⊆ B
(2) Since x ∈ B then x ∈ C Definition of B ⊆ C
(3) If x ∈ A then x ∈ C By statements (1) and (2)
(4) Therefore A ⊆ C Definition of A ⊆ C
Note: A proof is often concluded with the symbol (as in the above proof) or the
letters QED.
Many mathematicians prefer to write proofs in paragraph form. For example, the
proof of Theorem 1.1 would become:
Proof. Choose an x ∈ A. Since A ⊆ B , x ∈ A implies x ∈ B , from the definition
of subset. Furthermore, since B ⊆ C , x ∈ B implies x ∈ C . Therefore, every
element x ∈ A is also an element of C . Hence A ⊆ C .
A proof of the distributive law for sets
Here is a proof of one of the distributive laws for sets. The proof is from a textbook
on set theory by Flora Dinkines2
Theorem 1.2. A ∩ (B ∪ C) = (A ∩ B) ∪ (A ∩ C).
Proof.
Discussion Reasons
(1) If x ∈ A ∩ (B ∪ C), then Definition of ∩.
x ∈ A and x ∈ B ∪ C .
(2) Since x ∈ B ∪ C , then Definition of ∪.
x ∈ B or x ∈ C .
(3) Therefore x ∈ A and x ∈ B , By (1) and (2).
or x ∈ A and x ∈ C .
(4) Hence x ∈ A ∩ B or Definition of ∩.
x ∈ A ∩ C.
(5) x ∈ (A ∩ B) ∪ (A ∩ C). Definition of ∪.
2
Dinkines, F., Elementary Theory of Sets, Meredith Publishing, 1964.
(6) If y ∈ (A ∩ B) ∪ (A ∩ C), Definition of ∪.
then y ∈ (A ∩ B) or
y ∈ (A ∩ C).
(7) If y ∈ A ∩ B then Definition of ∩.
y ∈ A and y ∈ B .
(8) If y ∈ A ∩ C then Same as (7).
y ∈ A and y ∈ C .
(9) In either case (7) or (8), Statements (7) and (8)
y ∈ A and y is an
element of one of the
sets B or C .
(10) Therefore y ∈ A and Since y is in B or C ,
y ∈ B ∪ C. it is in the union.
(11) Therefore y ∈ A ∩ (B ∪ C). Definition of ∩.
A ∩ (B ∪ C) = (A ∩ B) ∪ (A ∩ C).
Elementary sets
Let Ω be a universe with three subsets A, B , and C . Consider the following subsets
of Ω:
A B C
0 0 0 Ac ∩ B c ∩ C c = S0
0 0 1 Ac ∩ B c ∩ C = S1
0 1 0 Ac ∩ B ∩ C c = S2
0 1 1 Ac ∩ B ∩ C = S3
1 0 0 A ∩ Bc ∩ C c = S4
1 0 1 A∩B ∩C c = S5
1 1 0 A ∩ B ∩ Cc = S6
1 1 1 A∩B∩C = S7
The sets S0 , S1 , . . . , S7 are called the elementary sets of A, B and C .
Ω
B
A
A Bc Cc A B Cc Ac B Cc
A B C
Ac B C
A Bc C
Ac Bc Cc Ac Bc C C
Cardinality of sets
Using set notation, we can now introduce the first components of our probability
model.
Definition 1.8. A sample space, Ω, is the set of all possible outcomes of an
experiment.
Definition 1.9. An event is any subset of a sample space.
Example: Consider the experiment of rolling a die once.
3 What should be Ω for this experiment?
Ω=
Example: Consider the experiment of burning a light bulb until it burns out. The
outcome of the experiment is the life length of the bulb.
3 What should be Ω for this experiment?
Ω = {x ∈ R : x ≥ 0}
A = {x ∈ R : 0 ≤ x < 1.5}
B = {x ∈ R : x ≥ 5}
C = {24}
Sets of events
It is sometimes useful to be able to talk about the set of all possible events F. The
set F is very different than the set of all possible outcomes, Ω. Since an event is a
set, the set of all events is really a set of sets. Each element of F is a subset of Ω.
For example, suppose we toss a two-sided coin twice. The sample space is
Ω = {HH, HT, T H, T T }.
If you were asked to list all of the possible events for this experiment, you would
need to list all possible subsets of Ω, namely
A1 = {{HH}}
A2 = {{HT }}
A3 = {{T H}}
A4 = {{T T }}
A5 = {{HH}, {HT }}
A6 = {{HH}, {T H}}
A7 = {{HH}, {T T }}
A8 = {{HT }, {T H}}
A9 = {{HT }, {T T }}
A10 = {{T H}, {T T }}
A11 = {{HH}, {HT }, {T H}}
A12 = {{HH}, {HT }, {T T }}
A13 = {{HH}, {T H}, {T T }}
A14 = {{HT }, {T H}, {T T }}
A15 = {{HH}, {HT }, {T H}, {T T }} = Ω
A16 = {} = Ø
In this case
F = {A1 , A2 , . . . , A16 }
Definition 1.10. For a given set A, the set of all subsets of A is called the power
set of A, and is denoted by 2A .
Theorem 1.3. If A is a finite set, then |2A | = 2|A|
Proof. Left to reader.
If Ω is a sample space, then we often use the set of all possible events F = 2Ω .
When Ω contains a finite number of outcomes, this often works well. However, for
some experiments, Ω contains an infinite number of outcomes (either countable or
uncountable) and F = 2Ω is much too large. In those cases, we select a smaller
collection of sets, F, to represent those events we actually need to consider. How
to properly select a smaller F, is found in many graduate level probability courses.
dominoes have each half of a tile painted with 1 to 9 dots (called “pips”) or a blank
for a total of 10 possible choices.
Of course, a domino painted (1,4) can be turned 180 degrees to look like a domino
painted (4,1). So there is no need to have one assembly line producing (1,4)
dominoes and another producing (4,1) dominoes.2 How many different dominoes
does your factory actually produce? (Try this problem on your friends and see how
many different answers you get.)
1
Chung, K. L., Elementary probability theory with stochastic processes, Springer-Verlag, New
York, 1975.
2
Just have a person at the end of the assembly line rotating half of the (4,1) the dominoes by 180
degrees.
19
In this chapter, we will present a systematic way to approach counting problems.
We will classify counting problems into four categories, and solve each type. In
practice, most counting problems can be approached using one or a combination
of these four basic types.
Fundamental Rule
The following rule forms the foundation for solving any counting problem:
Multiplication Principle: Suppose there are several multiple choices to be made.
There are m1 alternatives for the first choice, m2 alternatives for the second, m3
for the third, etc. The total number of alternatives for the whole set of choices is
equal to
(m1 )(m2 )(m3 ) · · · .
Example: Suppose a engineering class has five women and four men. You would
like to select a research team consisting of one man and one woman. There are
four choices for selecting the man and five choices for selecting the woman. The
total number of possible research teams is, therefore, 4 · 5 = 20
Sampling
The Urn Model: Imagine an urn that contains n balls marked 1 to n. There are
four disciplines under which the balls can be drawn from the urn and examined.
I. Sampling with replacement and with ordering.
A total of m balls are drawn sequentially. After a ball is drawn, it is replaced
in the urn and may be drawn again. The order of the balls is noted. Hence the
sequence 1, 3, 2, 3 is distinguished from 3, 3, 1, 2. Since there are n possible
ways to draw the first ball, n possible ways to draw the second, etc., the total
number of possible sequences is
nm .
Example: Suppose you give a quiz with 5 questions, each with a possible
answer of True or False. Therefore, there are two ways to answer the first
question, two ways to answer the second question, and so on. The total
number of possible ways to answer all of the questions on the exam is
2 · 2 · 2 · 2 · 2 = 25 = 32.
II. Sampling without replacement and with ordering.
A total of m balls are drawn sequentially with m ≤ n. After a ball is drawn,
it is not replaced in the urn and hence cannot be drawn again. The order of
the balls is noted. Since there are n possible ways to draw the first ball, n − 1
possible ways to draw the second, n − 2 possible ways to draw the third, etc.,
the total number of possible sequences is
For the special case when m = n, the total number of possible ways to draw
n balls is
n(n − 1)(n − 2) · · · (2)(1) ≡ n!.
Note that when m = n, this is equivalent to the problem of counting the
number of ways you can arrange n different balls in a row.
Computational Hint: We can compute
n!
n(n − 1)(n − 2) · · · (n − m + 1) = .
(n − m)!
By definition, 0! ≡ 1 so that when n = m, we have
n! n!
n(n − 1)(n − 2) · · · (n − n + 1) = = = n !.
(n − n)! 0!
Example: Suppose you have five different birthday cards, and you have
three friends (named Bill, Mary, and Sue) with birthdays this week. You
want to send one card to each friend. The number of possible choices for
Bill’s card is 5. The card for Mary’s must come from the remaining 4 cards.
Finally, the number of possible choices for Sue’s card is 3. Hence the total
number of ways you can assign five birthday cards to your three friends is
5 · 4 · 3 = 60.
It’s interesting to note that the sequence of friends (Bill, then Mary, then Sue)
is arbitrary and does not affect the answer.
Here is another way to look at this problem: Consider the empty mailboxes
of your three friends,
? ? ?
B M S
If the cards are labelled a,b,c,d,e, then one possible assignment is
c d a
B M S
There are 5 ways to fill the first position (i.e., Bill’s mailbox), 4 ways to fill
the second position and 3 ways to fill the third.
Example: Using the above example, we can compute the number of ways
one can arrange the three letters x, y, z. This time we have 3 positions to fill
with 3 different items.
? ? ?
1 2 3
So a possible arrangement is
y z x
1 2 3
n!
x · m! =
(n − m)!
In other words,
[Stage 1] · [Stage 2] = [Case II].
Solving for x then yields our result.
Computational Hint: Note that
n n! n! n
= = = .
m m!(n − m)! (n − m)!m! n−m
So, for example,
5 5 9 9
= and = .
3 2 4 5
Also
n
note that if n is large, brute force computation of all three factorial terms
of m can be a challenge. To overcome this, note that
n n! n(n − 1)(n − 2) · · · (n − m + 1)
= = .
m m!(n − m)! m!
with only m terms in both the numerator and the denominator. Hence,
10 10! 10 · 9
= = = 45
2 2!8! 2!
and
100 100 100! 100 · 99 · 98
= = = = 161700.
97 3 3!97! 3!
Example: A consumer survey asks: Select exactly two vegetables from the
following list:
◦ peas ◦ carrots ◦ lima beans ◦ corn
4
The total number of ways you can select two items from the four is 2 = 6.
IV. Sampling with replacement and without ordering.
This case is perhaps the most difficult to perceive as a sampling discipline,
yet it has important applications. A total of m balls are drawn sequentially
from the urn. Each ball is replaced after being drawn. However, the sequence
of the draws is ignored. Therefore, drawing 3, 3, 1, 2 is the same as drawing
1, 3, 2, 3. The number of ways that m balls can be drawn from an urn
containing n balls in this fashion is
m+n−1
.
m
Let’s prove that this formula works. First, note that to describe a possible
outcome from this scenario, we only need to know how many times each
ball was drawn. For example, to describe 3, 3, 1, 2 or 1, 3, 2, 3 we only need
to know that Ball 1 was drawn once, Ball 2 was drawn once, and Ball 3 was
drawn twice. This arrangement can be represented as follows:
1 2 3
√ √ √√
1 2 3 3
1 2 3
√√√ √
2 2 2 3
Now notice that any of these outcomes can be represented by only specifying
√
the arrangement of the four check marks ( ) among the inner two vertical
lines (|). For example,
Coded
1 2 3 Representation
√ √ √√ √ √ √√
1 2 3 3 | |
√√√ √ √√√ √
2 2 2 3 | |
? ? ? ? ? ?
√
Any four of these can be selected for the check marks ( ) and the remaining
two will be filled with the two vertical lines
(|). The number of ways we
can select 4 positions from 6 positions is 64 = 15. Therefore, there are 15
possible ways we can select 4 balls from an urn containing 3 balls if we allow
replacement and disregard ordering.
In general, if there are n balls and m drawings, we will need m check marks
√
( ). To represent all possible outcomes, our table will have n columns.
Coded
1 2 ··· n Representation
√√ √ √ √√ √ √
k1 k2 · · · kn ··· | |···|
The number of ways we can select m positions for check marks from (n −
1) + m positions is
n+m−1
m
Example: Mary wants to buy a dozen cookies. She likes peanut butter
cookies, chocolate chip cookies and sugar cookies. The baker will allow her
to select any number of each type to make a dozen. In this case, our table
has three columns with twelve check marks. So the following arrangement:
would represent the outcome when Mary selects 2 peanut butter, 3 chocolate
chip and 7 sugar cookies. In this case, we are arranging 14 symbols in a row:
2 vertical lines and 12 check marks. So the number of possible different
boxes of one dozen cookies is
12 + 3 − 1 14 14 14 · 13
= = = = 91.
12 12 2 2·1
An extension of case III is the situation where one has n balls with n1 of them
painted with color number 1, n2 of them painted with color number 2,. . . ,nk of
them painted with color number k . We have
k
ni = n.
i=1
Occupancy
The Token Model: In sampling problems, we think of n balls placed in an urn from
which m balls are drawn. A direct analogy can be made with so-called occupancy
problems. In such models, we are given n boxes marked 1 to n, and m tokens
numbered 1 to m. There are four disciplines for placing the tokens in the boxes.
Each of these four scenarios correspond to one of the four sampling cases.
The reason for introducing occupancy problems becomes clear as one encounters
different counting problems. Some may be easier to conceptualize as occupancy
problems rather than sampling problems. Nevertheless, the number of distinguish-
able ways in which the tokens may occupy the boxes is computed in precisely the
same manner as in the sampling models.
The connection between the two formulations is easy to remember if you note the
following relationships:
Sampling Problems Occupancy Problems
ball box
number of drawing number on token
j th draw yields ball k j th token is placed in box k
For example, suppose we have a sampling problem with 3 balls (labelled a, b, c)
being drawn 4 times. This would correspond to an occupancy problem with 3
boxes and 4 tokens (See Figure 2.2). Moreover, the following outcomes from each
Tokens Tokens
1 2 3 4 1 2 3 4
2 1 4
3
a b c a b c
I*. If more than one teacher can be assigned to a classroom, and it matters which
teacher teaches in a particular room, then the number of ways you can assign
teachers to classrooms is
nm = 53 = 125.
II*. If only one teacher can be assigned to a classroom, and it matters which teacher
teaches in a particular room, then the number of ways you can assign teachers
to classrooms is
n(n − 1) · · · (n − m + 1) = 5 · 4 · 3· = 60.
III*. If only one teacher can be assigned to a classroom, and it does not matter
which teacher teaches in a particular room, then the number of ways you can
assign teachers to classrooms is
n 5
= = 10.
m 3
IV*. If more than one teacher can be assigned to a classroom, and it does not
matter which teacher teaches in a particular room, then the number of ways
you can assign teachers to classrooms is
m+n−1 3+5−1
= = 35.
m 3
(C) 10!
(D) 1210
(E) none of the above.
S2.2 A teacher has a class of 30 students. Each day, she selects a team of 6
students to present their homework solutions to the rest of the class. How
many different teams of 6 students can she select?
30
(A) 6
(B) 306
(C) 30!/24!
(D) 630
(E) none of the above
S2.3 A box is partitioned into 3 numbered cells. Two indistinguishable balls are
dropped into the box. Each cell may contain at most one ball. The total
number of distinguishable arrangements of the balls in the numbered cells is
(A) 1
(B) 3
(C) 6
(D) 9
(E) none of the above.
S2.4 A professor wants to use three colors (red, green and blue) to draw an
equilateral triangle on a sheet of paper. Each edge of the triangle is a single
color, but any color can be used more than once. How many different
triangles can be drawn?
(A) 33
(B) 3!
5
(C) 3
5
(D) 3 +1
(E) none of the above.
S2.5 A professor has an equilateral triangle cut from a sheet of cardboard. He
wants to use three colors (red, green and blue) to color the edges. Each edge
of the triangle is a single color, but any color can be used more than once.
How many different triangles can be produced?
(A) 33
(B) 3!
5
(C) 3
5
(D) 3 +1
(E) none of the above.
S2.6 Two white balls are dropped randomly into four labelled boxes, one at a
time. A box may contain more than one ball. The number of distinguishable
arrangements of the balls in the boxes is
4
(A) 2
7
(B) 2
(C) 4
(D) 42
(E) none of the above.
S2.7 One white ball and one black ball are dropped randomly into four labelled
boxes, one at a time. A box may contain more than one ball. The number of
distinguishable arrangements of the balls in the boxes is
4
(A) 2
7
(B) 2
(C) 4
(D) 42
(E) none of the above.
S2.8 A class has ten students. The teacher must select three students to form a
committee. How many different possible committees can be formed?
(A) 103
(B) 310
(C) 10!/7!
10
(D) 3
(D) 3!
(E) none of the above.
S2.11 An ice cream shop offers 20 different flavors of ice cream. Their menu says
that their “Big Dish” contains five scoops of ice cream. You may order the
five scoops in any way, ranging from five different flavors to five scoops of
the same flavor. How many different “Big Dishes” are possible?
(A) 205
20
(B) 5
24
(C) 5
(D) 20!/5!
(E) none of the above
S2.12 There are 4 machines in a factory. If each is worker assigned to one machine,
how many ways can 4 workers be assigned to the machines?
(A) 42
(B) 4!
7
(C) 4
(D) 44
(E) none of the above
The scientific imagination always restrains itself within the limits of probability.
– THOMAS HENRY HUXLEY, Reflection (c 1880)
Probability Spaces
In Chapter 1, we showed how you can use sample spaces Ω and events A ⊆ Ω
to describe an experiment. In this chapter, we will define what we mean by
the probability of an event, P (A) and the mathematics necessary to consistently
compute P (A) for all events A ⊆ Ω.
For infinite sample spaces, Ω, (for example, Ω = R) we often cannot reasonably
compute P (A) for all A ⊆ Ω. In those cases, we restrict our definition of P (A) to
some smaller subset of events, F, that does not necessarily include every possible
subset of Ω. The mathematical details are usually covered in advanced courses in
probability theory.
Measuring sets
Definition 3.1. The sample space, Ω, is the set of all possible outcomes of an
experiment.
Definition 3.2. An event is any subset of the sample space.
How would you measure a set?
• finite sets
• countable sets
• uncountable sets
35
1. P (E) ≥ 0 for all events E
2. P (Ω) = 1
3. If E ∩ F = Ø (i.e., if the events E and F are mutually exclusive ), then
P (E ∪ F ) = P (E) + P (F )
The three specifications in the above definition are often called the Kolmogorov
axioms for probability measures. However, a great deal of probability theory has
its foundations in a mathematical field called measure theory with important results
from many mathematicians, including Emile Borel, Henri Lebesgue and Eugene
Dynkin.
Note that every event A ⊆ Ω is assigned one and only one number P (A). In
section 3, we will show that 0 ≤ P (A) ≤ 1 for every A ⊆ Ω. In mathematical
terms, P (·) is a function that assigns to each member of F, the set of events, a
single number from the interval [0, 1]. That is
P : F → [0, 1].
Such functions assign each number in R a single number from R. For example,
f (x) = x3 , maps x = 2 into f (2) = 8.
Here, the function, P (·), is a function whose argument is a subset of Ω (namely
an event) rather than a number. In mathematics we call functions that assign
real numbers to subsets, measures. That’s why we refer to P (·) as a probability
measure.
Interpretation of probability
|A|
P (A) =
n
For example, consider the outcome of rolling a fair six-sided die once, In
that case
Ω = {ω1 , ω2 , . . . , ω6 }
where
ωi = the die shows i pips
for i = 1, 2, . . . , 6. So we can assign, for example,
1
P ({ωi }) = for i = 1, . . . , 6
6
3
P ({The die is even}) = P ({ω2 , ω4 , ω6 }) =
6
and, in general,
|A|
P (A) = for any A ⊆ Ω
6
Logical probability is only useful when the mechanism producing the random
outcomes has symmetry so that one could argue that there is no reason for one
outcome to occur rather than other. Some examples where logical probability
can be used are:
tossing a fair coin The coin is symmetrical and there is no reason for heads
to turn up rather than tails.
rolling a fair die The die is symmetrical and there is no reason for one side
to turn up rather than another.
drawing a card randomly from a deck of playing cards Each card is phys-
ically the same, and there is no reason for one card to be drawn rather
than another.
As you can see, the notion of logical probability has limited applications.
Experimental probability This interpretation requires that the experiment can
be identically reproduced over and over again, indefinitely. Each replicate
is called an experimental trial. We also must be able to assume that the
outcome of any one of the trials does not affect the probabilities associated
with the outcomes of any of the other trials. As an example, suppose we
examine transistors being produced by an assembly line, with each transistor
an experimental trial. The outcome of one of our experimental trials is
whether that transistor is either good or bad. Consider the sequence of
outcomes for the repeated experiment, with
Ti = 0 if transistor i is bad
Ti = 1 if transistor i is good
T 1 , T2 , . . .
and assign n
i=1 Ti
P (A) = lim
n→∞ n
There are some important mathematical issues associated with defining P (A)
in this way. In particular, the type of limit that you defined in calculus cannot
be used. In calculus, there always existed some K , determined in advance,
such that for all n > K the quantity
n
i=1 Ti
n
was arbitrarily close to P (A). That concept is not suitable here.
Suppose, for instance, that P (A) equaled 12 . It is possible, although not
likely, that you could examine
billions of transistors and find that all of them
n
Ti
are bad. In that case, i=1 n would be zero, and quite far away from the
1
eventual limit of 2 as n gets bigger than billions.
In fact, it can be proven that for a prespecified small number δ > 0, there
does exist some integer K such that for all n > K
n
i=1 Ti
− P (A) < δ.
n
The value of K is, itself, random and cannot be predetermined. We can,
however, say that P (K < ∞) = 1. The proof requires measure theory, and
is typically covered in a graduate-level probability course.
Some examples where experimental probability can be used are:
quality assurance Specifying the probability that a single microprocessor
from a semiconductor fabrication facility will be defective.
interarrival times for queues Stating the probability that the time between
any two consecutive customers is greater than five minutes.
hitting a home run Specifying the probability that a baseball player will hit
a home run on any attempt.
Subjective probability There are some cases where we would like to use probabil-
ity theory, but it is impossible to argue that the experiment can be replicated.
For example, consider the experiment of taking EAS305, with the outcome
being the grade that you get in the course. Let
Example: A box contains three light bulbs, and one of them is bad. If two light
bulbs are chosen at random from the box, what is the probability that one of the
two bulbs is bad?
Solution 1: Assume that the bulbs are labelled A, B , and C and that light bulb A
is bad. If the two bulbs are selected one at a time, there are (3)(2) = 6 possible
outcomes of the experiment:
AB BA
AC CA
CB BC
Note that 4 of the 6 outcomes result in choosing A, the bad light bulb. If these 6
outcomes are equally likely (are they?), then the probability of selecting the bad
bulb is 46 = 23
Solution 2:
3
If the two bulbs are selected simultaneously (i.e., without ordering),
there are 2 = 3 possible outcomes of the experiment:
AB AC CB
and 2 of the 3 outcomes result in choosing bad bulb A. If these 3 outcomes are
equally likely (are they?), then the probability of selecting the bad bulb is 23
Example: A box contains two light bulbs, and one of them is bad. Suppose a
first bulb is chosen, tested and then placed back in the box. Then a second bulb is
chosen, and tested. Note that the same bulb might be selected twice. What is the
probability that exactly one of the two tested bulbs is bad?
“Solution:” Suppose the two bulbs are labelled A and B and bulb A is the bad
light bulb, There are three possible outcomes:
So the probability that exactly one bulb is bad is 13 . (Actually, this solution is
wrong. Why it is wrong?)
Relationships between English and set notation
• P (not E) = P (E c )
• P (E or F ) = P (E ∪ F )
• P (E and F ) = P (E ∩ F )
A∪B = B since A ⊆ B
(A ∪ B) ∩ (A ∪ Ac ) = B since A ∪ Ac = Ω
A ∪ (B ∩ Ac ) = B DeMorgan’s law
P (A ∪ (B ∩ Ac )) = P (B) See Question 3.31
P (A) + P (B ∩ Ac ) = P (B) A ∩ (B ∩ Ac ) = Ø
P (A) ≤ P (B) since P (B ∩ Ac ) ≥ 0 (Axiom 1)
(A1 ∪ · · · ∪ Ak ) ∩ Ak+1 = Ø
since Ai ∩ Ak+1 = Ø for all i
P ((A1 ∪ · · · ∪ Ak ) ∪ Ak+1 ) = P (A1 ∪ · · · ∪ Ak ) + P (Ak+1 )
since (A1 ∪ · · · ∪ Ak ) ∩ Ak+1 = Ø
P ((A1 ∪ · · · ∪ Ak ) ∪ Ak+1 ) = (P (A1 ) + . . . + P (Ak )) + P (Ak+1 )
from induction hypothesis
P (A1 ∪ · · · ∪ Ak ∪ Ak+1 ) = P (A1 ) + . . . + P (Ak+1 )
Associative law
P (A ∪ B) = P (A) + P (B) − P (A ∩ B)
Proof.
A ∪ B = A ∪ (Ac ∩ B)
Distributive law for sets
and
B = (A ∩ B) ∪ (Ac ∩ B)
Distributive law for sets
Therefore
P (A ∪ B) = P (A) + P (Ac ∩ B)
from Axiom 3
P (B) = P (A ∩ B) + P (Ac ∩ B)
from Axiom 3
Subtracting the second equation from the first produces the result.
Example: A hydraulic piston can have only two types of defects:
1. a surface defect, and/or
2. a structural defect
A single piston may have one, both or neither defect. From quality control records
we see that
• The probability that a piston has a surface defect is 0.5.
• The probability that a piston has a structural defect is 0.4.
• The probability that a piston has some type of defect is 0.6
To find the probability that a piston has both a structural and a surface defect, first
define the following events:
We want to find P (S ∩ T ).
We know that
P (S) = 0.5
P (T ) = 0.4
P (S ∪ T ) = 0.6
P (S ∪ T ) = P (S) + P (T ) − P (S ∩ T )
0.6 = 0.4 + 0.5 − P (S ∩ T )
P (S ∩ T ) = 0.4 + 0.5 − 0.6 = 0.3
In order to allow the extension of Axiom (3) in Definition 3.3 to apply to a countable
collection of disjoint sets (rather than just a finite collection), the third probability
axiom is replaced by
3.* If A1 , A2 , . . . is a countable sequence of disjoint sets, then
∞ ∞
P Ai = P (Ai )
i=1 i=1
Example: Consider the experiment of counting the number of accidents in a factory
in a month. Let Ei be the event that exactly i accidents occur in a month with
i = 0, 1, 2, . . .. Suppose that
e−1
P (Ei ) = for i = 0, 1, 2, . . .
i!
and we want to find the probability that the factory has at least one accident in a
month.
That is, we want to compute P ∪∞
i=1 Ei .
∞ α i α
Since i=0 i! = e for any α, we have
∞
∞ −1
e
P (Ei ) =
i=0 i=0
i!
∞
1
= e−1
i=0
i!
−1 1
= e e =1
Using axiom 3.* and the fact that Ei ∩ Ej = Ø for any i = j , we get
∞ ∞
P Ei = P (Ei ) = 1
i=0 i=0
∞
P E0 ∪ Ei = 1
i=1
∞
P (E0 ) + P Ei = 1 Since E0 ∩ ∪∞
i=1 Ei = Ø
i=1
∞
P Ei = 1 − P (E0 ) = 1 − e−1
i=1
and we see that we need at least 23 persons to insure that P (E) is at least 0.50.
But why would we get the wrong answer if we don’t order the k individuals? Let’s
see what this would mean for k = 3.
Identifying the elements of Ω without ordering the individuals is equivalent to
listing all of the ways we can draw 3 balls from an urn containing 365 balls, with
replacement and without ordering. We can do this by creating a table with 365
columns (one for each calendar day) and asking the question, “how many ways can
we arrange 3 check marks among the 365 columns?”
So that
(n)!(n − 1)!
P (E) = 1 − P (F ) = 1 −
(n + k − 1)!(n − k)!
To show why the unordered individual assumption is wrong, let’s use all of our
formulas to solve a slightly modified version of the birthday problem:
There are two aliens from outer space in a room. They are from a planet called Yak.
Yak has only two days in its calendar year. The names of the days are “Heads” and
“Tails”.
Once again, assume that birthdays are randomly assigned, with each day equally
likely. Assigning birthdays at random to the aliens is the same as each alien tossing
a fair coin to determine it’s birthday. So we can use our formulas with values n = 2
and k = 2.
Under the ordered individual model there are four elements in Ω, that is
Ω = {HH, HT, T H, T T },
and
(2)!/(2 − 2)! 1
P (E) = 1 − =
(2)2 2
(1, 1) (1, 2) (1, 3) (1, 4) (1, 5) (1, 6)
(2, 1) (2, 2) (2, 3) (2, 4) (2, 5) (2, 6)
(3, 1) (3, 2) (3, 3) (3, 4) (3, 5) (3, 6)
(4, 1) (4, 2) (4, 3) (4, 4) (4, 5) (4, 6)
(5, 1) (5, 2) (5, 3) (5, 4) (5, 5) (5, 6)
(6, 1) (6, 2) (6, 3) (6, 4) (6, 5) (6, 6)
that is, “two heads”, “two tails” and “one head one tail.” And the corresponding
computation for P (E) yields
(2)!(2 − 1)! 2
P (E) = 1 − P (F ) = 1 − =
(2 + 2 − 1)!(2 − 2)! 3
And we see that the second (and faulty) formulation results in the same mistake
that Jean d‘Alembert2 made in 1754.
Moral of the story: Sampling Case IV (with replacement and without ordering) usu-
ally cannot be used to enumerate equally likely outcomes resulting from multiple,
repeated random assignments.3
Example: Roll a die twice. The sample space can be represented as in Figure 3.1.
Define the events
P (A ∩ B) = P (A | B)P (B)
P (A ∩ B)
P (A | B) =
P (B)
P (· | B) is a probability measure
P (E ∩ B)
P (E | B) =
P (B)
[something ≥ 0]
= ≥0
[something > 0]
Therefore P (Ω | B) = 1
• And for any two events E and F with E ∩ F = Ø
P ((E ∪ F ) ∩ B)
P (E ∪ F | B) =
P (B)
P ((E ∩ B) ∪ (F ∩ B)
=
P (B)
P (E ∩ B) + P (F ∩ B)
=
P (B)
P (E ∩ B) P (F ∩ B)
= +
P (B) P (B)
= P (E | B) + P (F | B)
Therefore if E ∩ F = Ø then P (E ∪ F | B) = P (E | B) + P (F | B)
Independence
P (A ∩ B) = P (A)P (B).
Note: If P (B) > 0 this is equivalent to saying that A and B are
independent if and only if
P (A | B) = P (A).
Bayes’ theorem
such that
1. B1 ∪ B2 ∪ · · · ∪ Bn = Ω, and
2. Bi ∩ Bj = Ø for any i = j .
Theorem 3.9. (Law of Total Probability) Given an event A and a complete set
of alternatives B1 , B2 , . . . , Bn then
n
n
P (A) = P (A ∩ Bi ) = P (A | Bi )P (Bi )
i=1 i=1
Example: (from Meyer4 ) Suppose that among six bolts, two are shorter than a
specified length. If two bolts are chosen at random, what is the probability that the
two short bolts are picked? Let
Although the answer is correct, the value of P (A2 ) is not 15 . But the value of
P (A2 | A1 ) is 15 .
To evaluate P (A2 ) properly, we note that {A1 , Ac1 } is a complete set of alternatives.
We can the use the law of total probability, and write
Example: Thousands of little snails live in the Snail Garden between the University
at Buffalo Natural Sciences Complex and Talbert Hall. A concrete stairway divides
the garden into two sections.
When it rains, the snails bravely travel from one section of the garden to the other by
crawling across the steps of the concrete stairway. It is a treacherous journey. Many
of the snails are crushed by the feet of unthinking students and faculty walking up
and down the stairway.
4
Meyer, P., Introductory probability theory and statistical applications, Addison-Wesley, Reading
MA, 1965.
There are two types of snails, smart snails and not-so-smart snails. A smart snail
crawls along the vertical portion of a step to avoid being crushed. A not-so-smart
snail crawls on the horizontal part of a step.
Suppose that half of the snails are smart snails, and the other half are not-so-smart
snails. The probability that a smart snail will be crushed is 0.10. The probability
that a not-so-smart snail will be crushed is 0.40. The event that any one snail will
be crushed is independent from all others.
3 To find the probability that a single snail (chosen at random) will successfully
cross the concrete steps uncrushed define the events
3 Suppose that after the end of a rainstorm, you select, at random, a snail that
has successfully crossed the concrete steps. What is the probability that it is
a smart snail? In this case, we want to find
P (A | S)P (S)
P (S | A) =
P (A | S)P (S) + P (A | S c )P (S c )
(0.90)(0.50)
= = 0.60
(0.90)(0.50) + (0.60)(0.50)
The probability of selecting a smart snail before the rainstorm was 0.50. The
probability after the rainstorm is 0.60.
What is the probability of finding a smart snails after two rainstorms?
Example: Suppose you roll a fair die repeatedly until a six appears. Assume all
the outcomes from each roll are independent. Let
Sk = {toss k is a six}
An = {the first six occurs on toss n}
Then
P (Sk ) = 1
6 P (Skc ) = 5
6
for k = 1, 2, . . ..
Furthermore,
An = S1c ∩ S2c ∩ · · · ∩ Sn−1
c
∩ Sn
for n = 1, 2, . . ..
for n = 1, 2, . . ..
We can also compute,
for k = 0, 1, 2, . . .. Which makes sense, since the first 6 occurs after toss k if and
only if you get k non-6’s in a row. So
for k = 0, 1, 2, . . ..
P (R = Heads | R = B)
is equal to
(A) 1/4
(B) 1/3
(C) 1/2
(D) 1
(E) none of the above.
S3.7 Two indistinguishable six-sided fair dice are rolled. What is the probability
that they both turn up the same value?
(A) 1/72
(B) 1/36
(C) 1/12
(D) 1/6
(E) none of the above.
S3.8 Suppose P (A) = 0. Then
(A) A = Ø
(B) P (A ∩ A) = P (A)P (A)
(C) P (A ∪ B) = 0 for any B ⊆ Ω.
(D) all of the above are true
(E) none of the above are true.
S3.9 Which of the following statements are always true?
(A) P (Ω) = 1
(B) If P (A) < P (B) then A ⊆ B
(C) If P (A) = 0 then A = Ø
(D) all of the above
(E) none of the above.
S3.10 Suppose P (A) = p, P (B) = q . If A ⊆ B and A = B then
(A) p ≥ q
(B) p ≤ q
(C) p > q
(D) p < q
(E) none of the above.
S3.11 There is one lock on a door and the key is among the six different ones you
usually carry in your pocket. Someone places a seventh, useless key, in your
pocket. What is the probability that the first key you try will open the door?
(A) 1/6
(B) 1/2
(C) 1/3
(D) 2/3
(E) none of the above.
S3.12 A multiple choice test has 12 questions, with each question having 5 possible
answers. If a student randomly picks the answer to each question, what is
the probability that the student will answer all questions correctly?
(A) 1/(125 )
(B) 1/(512 )
12
(C) 1/ 5
16
(D) 1/ 5
(B) 1/4
(C) 1/16
(D) it depends on whether or not the balls are different colors
(E) none of the above.
S3.14 A student club has four officers, three of them are seniors and one of them
is a junior. Three of the students will be picked at random to go on a trip.
What is the probability that the junior stays home?
(A) 1/4
6
(B) 1/ 3
(C) 1/4!
(D) 1/43
(E) none of the above.
S3.15 Suppose that five tiles labelled with the letters C, C, C, A, T are placed in a
container. If three tiles are selected from the container (without replacement),
what is the probability that they can be arranged to spell the word CAT?
(A) 1
(B) 1/3
5
(C) 1/ 3
5
(D) 3/ 2
S3.19 Suppose P (A) = 1 and P (B) = 0.25. Which of the following statements
are always true?
(A) A and B are independent
(B) A ∩ B = B
(C) B ⊆ A
(D) all of the above are true.
(E) none of the above are true.
S3.20 A multiple choice test has 5 questions, with each question having 4 possible
answers. If a student randomly picks the answer to each question, what is
the probability that the student will answer all questions correctly?
(A) 1/(45 )
(B) 1/4
5
(C) 1/ 4
8
(D) 1/ 4
(B) P (A) = 0
(C) A ⊆ Ac
(D) P (A) can be any value between 0 and 1
(E) none of the above.
1 1
S3.25 If P (A) = 4 and P (B) = 4 then
1
(A) P (A ∪ B) = 2
1
(B) P (A ∪ B) ≤ 2
(C) A ∩ B = Ø
(D) A = B
(E) none of the above.
S3.26 The probability that the birthdays of twelve people will fall in twelve dif-
ferent calendar months (assuming equal probabilities for the twelve months)
is
(A) 12!/(1212 )
(B) 1/(12!)
(C) 1/12
(D) 1/(1212 )
(E) none of the above.
S3.27 In a screw factor, there are three machines A, B and C that produce screws.
Machine A produces 20% of the daily output. Machine B produces 30% of
the daily output. Machine C produces the remaining 50% of the daily output.
Of their output, 5%, 4% and 2% of the screws are defective, respectively. A
screw is chosen at random from a day’s production and found to be defective.
What is the probability that it came from machine A?
1. 0.01
2. 0.032
3. 0.2
4. 0.3125
5. none of the above.
S3.28 A box contains fifty light bulbs, exactly five of which are defective. If five
bulbs are chosen at random from the box, the probability that all of them are
good is equal to
(A) (0.01)5
50
(B) 1/ 5
50
(C) 5/ 5
Both lamps must be good in order for either to light. Two lamps are randomly
chosen from the box. The probability that the lamps will light is
(A) 2/9
(B) 1/4
(C) 1/2
(D) 3/4
(E) none of the above.
S3.34 Which of the following statements (A) (B) or (C) are not always true?
(A) P (E ∩ F ) = P (E)P (F ) for any events E and F
(B) P (Ω) = 1
(C) P (E) ≥ 0 for any event E
(D) all of the above are always true.
(E) none of the above are always true.
S3.35 Suppose
Then P (B | A) equals
(A) 0.2
(B) 1/3
(C) 0.5
(D) 0.6
(E) none of the above.
S3.36 A study of automobile accidents produced the following data:
Probability of
Model Proportion of involvement
year all vehicles in an accident
2002 0.10 0.02
2003 0.20 0.01
Other 0.70 0.05
You know that an automobile has been involved in an accident. What is the
probability that the model year for this automobile is 2003?
(A) 0.0200
(B) 0.0390
(C) 0.0513
(D) 0.3333
(E) none of the above.
4
RANDOM VARIABLES
Basic Concepts
Definition 4.1. A random variable is a function X(·) which assigns to each
element ω of Ω a real value X(ω).
Notation:
(1) P {ω ∈ Ω : X(ω) = a} ≡ P {X(ω) = a} ≡ P (X = a)
FX (a) ≡ P (X ≤ a).
1. lim FX (a) = 1
a→∞
2. lim FX (a) = 0
a→−∞
81
1
0.8
0.6
F(x)
0.4
0.2
0 1 2 3 4 5
x
Theorem 4.1. For any random variable X and real values a < b
Consider a sequence of n independent, repeated trials, with each trial having two
possible outcomes, success or failure. Let p be the probability of a success for any
single trial. Let X denote the number of successes on n trials. The random variable
X is said to have a binomial distribution and has probability mass function
n k
(2) pX (k) = p (1 − p)(n−k) for k = 0, 1, . . . , n.
k
n
Let’s check to make sure that if X has a binomial distribution, then k=0 pX (k) =
1. We will need the binomial expansion for any polynomial:
n
n k (n−k)
n
(a + b) = a b
k=0
k
So
n
n k
p (1 − p)(n−k) = (p + (1 − p))n
k=0
k
= (1)n = 1
Recall the “success/failure” model that was used to derive the binomial distribution,
but, this time, we want to find the probability that the first success occurs on the
k th trial. Let p be the probability of a success for any single trial and let Y denote
the number of trials needed to obtain the first success.
Then the probability mass function for Y is given by
and the random variable Y is said to have a geometric distribution with parameter
p where 0 ≤ p ≤ 1.
Some facts about geometric series When working with the geometric
distribution, the following results are often helpful:
Theorem 4.2. For any 0 < θ < 1,
∞
1
θj =
j=0
1−θ
Proof.
∞
∞
θj − θj = 1
j=0 j=1
∞
∞
j j+1
θ0 = θ − θ = 1
j=0 j=0
∞
∞
θj − θ θj = 1
j=0 j=0
∞
(1 − θ) θj = 1
j=0
∞
1
θj =
j=0
1−θ
Example: Using Theorem 4.2, we can show that the probability mass function for
a geometric random variable must sum to one. Suppose
then
∞
∞
pY (k) = (1 − p)k−1 p
k=1 k=1
∞
= (1 − p)j p
j=0
1
= p=1
1 − (1 − p)
αk
(3) pX (k) = e−α for k = 0, 1, 2, . . .
k!
0.8
0.6
F(x)
0.4
0.2
0 1 2 3 4 5
x
P (X < a) = P (X ≤ a).
x x+dx
pmf pdf
pX (x) ≥ 0 fX (x) ≥ 0
∞
pX (x) = 1 fX (x) dx = 1
all x −∞
a
FX (a) = pX (x) FX (a) = fX (x) dx
x≤a −∞
pX (x) = P (E) fX (x) dx = P (E)
x∈E E
The random variable X is said to have a uniform distribution if it has the proba-
bility density function given by
⎧
1⎨
if a ≤ x ≤ b
(6) fX (x) = b−a
⎩
0 otherwise
Then
+∞
E(X) = xfX (x) dx
−∞
0 1 +∞
= x0 dx + x2x dx + x0 dx
−∞ 0 1
1 1
2 2x3 2 2
= 2x dx = = −0=
0 3 0 3 3
Finding E(X) for some specific distributions
Binomial distribution
Let s = k − 1
n−1
n − 1 s+1
E(X) = n p (1 − p)(n−s−1)
s=0
s
n−1
n−1 s
= np p (1 − p)(n−s−1)
s=0
s
= np
Poisson distribution
αk
pX (k) = e−α for k = 0, 1, 2, . . .
k!
The expected value for X can be found as follows:
∞
αk
E(X) = ke−α
k=0
k!
∞
αk
= e−α
k=1
(k − 1)!
Let s = k − 1
∞
αs+1
E(X) = e−α
s=0
s!
∞
αs
= α e−α
s=0
s!
= α
Geometric distribution
Exponential distribution
P (X = −1) = P (X = 0) = P (X = 1) = 1/3
(B) P (X = a) = FX (a) − FX (a − 1)
(C) 0 ≤ FX (a) ≤ 1 for every a
(D) all of the above are true
(E) none of the above are true.
S4.8 The length X (in inches) of a steel rod is a random variable with probability
density function given by
⎧
⎨ x−9
if 9 ≤ x ≤ 11
fX (x) =
⎩ 0 2 otherwise
Then P (X ≤ 1) equals
(A) 1/8
(B) 1/4
(C) 1/3
(D) 1/2
(E) none of the above.
S4.12 Let X be a discrete random variable with probability mass function
e−1
P (X = k) = for k = 0, 1, 2, . . .
k!
Then P (X > 1) equals
(A) e−1
(B) 1 − e−1
(C) 1 − 2e−1
(D) 1
(E) none of the above.
S4.13 You have two light bulbs. The life length of each bulb (in hours) is a
continuous random variable X with probability density function
2e−2x if x > 0
fX (x) =
0 otherwise
The two bulbs are allowed to burn independently of each other. The proba-
bility that both bulbs will each last at least one hour is
(A) 0
(B) (1 − e−2 )2
(C) e−4
(D) 2e−2
(E) none of the above.
S4.14 Let X be a continuous random variable with cumulative distribution function
⎧
⎪
⎨ 0 if x < −1
1
FX (x) = 2 (x + 1) if −1 ≤ x ≤ 1
⎪
⎩ 1 if x > 1
Let fX (x) denote the probability density function for X . The value of fX (0)
is
(A) 0.00
(B) 0.50
(C) 0.75
(D) 1.00
(E) none of the above.
S4.15 Let X be a random variable with cumulative distribution function FX (x).
Suppose FX (0) = 0.3 and FX (1) = 0.5. Then P (0 < X ≤ 1) equals
(A) 0
(B) 0.2
(C) 0.3
(D) 0.5
(E) none of the above.
S4.16 Let X be a continuous random variable with probability density function
3x2 if 0 ≤ x ≤ 1
fX (x) =
0 otherwise
An integrated circuit considered defective if its life length is less than 3000
hours. The life lengths of any two integrated circuits are independent random
variables. A batch of 100 integrated circuits are tested. The expected number
of defective items is
(A) e−3
(B) (1 − e−3 )
(C) (1 − e−9 )
(D) 100(1 − e−9 )
(E) none of the above.
S4.18 Let X be a discrete random variable with cumulative distribution function,
F (a) ≡ P (X ≤ a), given by the following graph:
0.8
0.6
FX(a)
0.4
0.2
-2 -1 0 1 2 3
a
Basic Concepts
Two of the primary reasons for considering random variables are:
1. They are convenient, and often obvious, representations of experimental out-
comes.
2. They allow us to restrict our attention to sample spaces that are subsets of the
real line.
There is a third, and probably more important reason for using random variables:
3. They permit algebraic manipulation of event definitions.
This chapter deals with this third “feature” of random variables. As an example,
let X be the random variable representing the Buffalo, New York temperature
at 4:00 pm next Monday in degrees Fahrenheit. This random variable might be
suitable for the use of Buffalonians, but our neighbors in Fort Erie, Ontario prefer
the random variable Y representing the same Buffalo temperature, except in degrees
Celsius. Furthermore, they are interested in the event
A = {ω ∈ Ω : Y (ω) ≤ 20}
where Ω is the sample space of the underlying experiment. Note that Ω is being
mapped by both of the random variables X and Y into the real line.
In order to compute P (A), we could evaluate the cumulative distribution function
for Y , FY (y), at the point y = 20. But suppose we only know the cumulative
distribution function for X , namely FX (x). Can we still find P (A)?
105
The answer is “yes.” First, notice that for any ω ∈ Ω,
5
(1) Y (ω) = 9 X(ω) − 32 .
A = {ω ∈ Ω : Y (ω) ≤ 20}
5
= ω∈Ω : 9 (X(ω) − 32) ≤ 20
= {ω ∈ Ω : X(ω) − 32 ≤ 36}
= {ω ∈ Ω : X(ω) ≤ 68} ,
So, knowing a relationship like Equation 1 between X(ω) and Y (ω) for every
ω ∈ Ω, allows a representation of the event A in terms of either the random
variable X or the random variable Y .
When finding the probability of the event A, we can use the shortcut notation
presented in Chapter 5:
P (A) = P (Y ≤ 20)
= P ( 59 (X − 32) ≤ 20)
= P (X − 32 ≤ 36)
= P (X ≤ 68)
= FX (68).
X( )
Real line
Y( ) 212
Real line
100
)
h(x
y=
20
x
68
y
y= h(x)
B
x
= h-1(B)
Some examples
Discrete case
Continuous case
Example: Let X be a uniformly distributed random variable on the interval (−1, 1),
i.e., the density for X is given by
1
if −1 ≤ x ≤ 1
fX (x) = 2
0 otherwise
FY (y) = P (Y ≤ y )
= P X2 ≤ y
√ √
= P − y≤X≤ y
√ √
= FX y − FX − y
√ √
y+1 − y+1
= −
√ 2 2
= y.
So, ⎧
⎪
⎨ 0 if y < 0
√
FY (y) = y if 0 ≤ y ≤ 1
⎪
⎩ 1 if y > 1
Finding the probability density function for Y is now easy. That task is left to you.
Example: Let X be a uniformly distributed random variable on the interval (0, 2),
i.e., the density for X is given by
1
if 0 ≤ x ≤ 2
fX (x) = 2
0 otherwise
Let h(x) be the greatest integer less than or equal to x, i.e., let h be given by the
step function shown in Figure 5.4. Find the distribution for Y = h(X).
Solution: Since X takes on the values ranging continuously from 0 to 2, Y can take
on only the integers 0, 1 and 2 (why 2?). We immediately know that the random
variable Y is discrete, and we have our first example of a continuous random
variable being mapped into a discrete random variable. Look at Figure 5.4 again
and try to visualize why this happened.
To find the probability mass function for Y , first let’s try to find pY (0). We know
that Y takes on the value 0 when X takes on any value from 0 to 1 (but not including
y
3
1
x
-2 -1 1 2 3 4
1). Hence,
1
1 x 1 1
pY (0) = dx = = .
0 2 2 0 2
Similarly,
2
1 1
pY (1) = dx =
1 2 2
and 2
1
pY (2) =
dx = 0,
2 2
which is good thing because we ran out of probability mass after the first two
computations.
So we have the answer,
pY (0) = pY (1) = 12 .
FY (y) = P [Y ≤ y] = P [h(X) ≤ y]
FY (y) = P [h−1 (h(X)) ≤ h−1 (y)]
FY (y) = P [X ≤ h−1 (y)]
FY (y) = FX (h−1 (y))
To get the relationship between the densities, take the first derivative of both sides
with respect to y and use the chain rule:
d d
FY (y) = FX (h−1 (y))
dy dy
d
fY (y) = fX (h−1 (y)) h−1 (y)
dy
We will leave it to you to find a similar result for the case when h(x) is decreasing
rather than increasing.
Note: This transformation technique is actually a special case of a more general
technique from calculus. If one considers the function Y = h(X) as a transfor-
mation (x) → (y), the monotonicity property of h allows us to uniquely solve
y = h(x). That solution is x = h−1 (y). We can then express the probability
density function of Y as
Let Y = h(X) = X 3 . Find the probability density function for Y without explicitly
finding the cumulative distribution function for Y .
Solution: The support for X is [−2, 2]. So the support for Y is [−8, 8].
1
Since h−1 (y) = y 3 , we can use Theorem 5.2 to get
d −1
fY (y) = fX (h−1 (y)) h (y)
dy
1 d 1
= fX (y 3 ) y3
dy
2
= fX (y 3 )( 13 )y − 3
1
1 1 1
When y < −8 or y > 8 we have y 3 < −2 or y 3 > 2. In those cases, fX (y 3 ) = 0
hence fY (y) is zero.
When −8 ≤ y ≤ 8
1 13 1 − 23
fY (y) = 4 |y |( 3 )y
1
3 √
1 1
= 12 |y | 3 y 2
1 − 13
= 12 |y |
Summarizing 1
|y − 3 |/12 if −8 ≤ y ≤ 8
fY (y) =
0 otherwise
Question: Toss a coin. Let X = 1 if heads, X = 0 if tails. Let h(X) denote your
winnings.
−5 x = 0
h(x) =
10 x = 1
Find E(h(X)), your expected winnings.
Question: Let X , the life length of a transistor, be a continuous random variable
with probability density function
2e−2x x ≥ 0
fX (x) =
0 otherwise
Suppose you earn ln(x) + x2 − 3 dollars if the transistor lasts exactly x hours. What
are your expected earnings?
Theorem 5.4. E(a) = a if a is a constant.
Proof. We will provide proofs for both the discrete and continuous cases.
We are really finding E(h(X)) where h(x) = a for all x.
Suppose X is a discrete random variable with probability mass function
pX (xk ) = P (X = xk ) = pk for k = 1, 2, . . ..
Then
E(h(X)) = h(xk )pk
k
E(a) = apk
k
= a pk = a(1) = a
k
Suppose X is a continuous random variable with probability density function
fX (x). Then
∞
E(h(X)) = h(x)fX (x) dx
−∞
∞
E(a) = afX (x) dx
−∞
∞
= a fX (x) dx = a(1) = a
−∞
pX (xk ) = P (X = xk ) = pk for k = 1, 2, . . ..
Then
E(h(X)) = h(xk )pk
k
E(X + b) = (xk + b)pk = (xk pk + bpk )
k k
= xk pk + bpk
k k
= xk pk + b pk = E(X) + b(1)
k k
Theorem 5.7. E[h1 (X) + h2 (X)] = E[h1 (X)] + E[h2 (X)] for any functions
h1 (X) and h2 (X)
Proof. We are finding E(h(X)) where h(x) = h1 (x) + h2 (x).
Suppose X is a discrete random variable with probability mass function
pX (xk ) = P (X = xk ) = pk for k = 1, 2, . . ..
Then
E(h(X)) = h(xk )pk
k
E[h1 (X) + h2 (X)] = [h1 (xk ) + h2 (xk )]pk
k
= [h1 (xk )pk + h2 (xk )pk ]
k
= h1 (xk )pk + h2 (xk )pk
k k
= E[h1 (X)] + E[h2 (X)]
Also notice that Theorem 5.6 is really only a corollary to Theorem 5.7. Simply use
Theorem 5.7 with h1 (x) = x and h2 (x) = b and you get Theorem 5.6.
.
Note that
Var(X) = (x − E(X))2 pX (x)
all x
for a discrete random variable X , and
+∞
Var(X) = (x − E(X))2 fX (x) dx
−∞
h (μ) 2
E(Y ) ≈ h(μ) + σ
2
h (μ) 2
h(μ) − E(Y ) ≈ − σ
2
Therefore, if you incorrectly state that h(μ) is your expected monthly profit, your
answer will be wrong by approximately
h (μ) 2
Δ=− σ
2
Note that if h is linear, h (x) = 0, and Δ = 0. This result for the linear case is also
provided by Theorems 5.5, 5.6 and 5.7.
Example: (Meyer) Under certain conditions, the surface tension of a liquid
(dyn/cm) is given by the formula S = 2(1 − 0.005T )1.2 , where T is the tem-
perature of the liquid (degrees centigrade).
Suppose T is a continuous random variable with probability density function
3000t−4 if t ≥ 10
fX (x) =
0 otherwise
Therefore,
∞
E(T ) = 3000t−3 dt = 15 (degrees centigrade)
10
Var(T ) = E(T 2 ) − (15)2
∞
= 3000t−2 dt − 225 = 75 (degrees centigrade)2
10
h(15) = 1.82
h (15) = 0.01
h (15) = 0.000012(1 − 0.005(15))−0.8 ≈ 0+
σ2
P (|X − μ| ≥ a) ≤
a2
Proof. We will prove this result when X is a continuous random variable. (The
proof for X discrete is similar using summations rather than integrals, and fX (x) dx
replaced by the probability mass function for X .)
We have for any a > 0
+∞
σ2 = (x − μ)2 fX (x) dx
−∞
μ−a μ+a +∞
2 2
= (x − μ) fX (x) dx + (x − μ) fX (x) dx + (x − μ)2 fX (x) dx
−∞ μ−a μ+a
μ−a +∞
= (x − μ)2 fX (x) dx + something ≥ 0 + (x − μ)2 fX (x) dx
−∞ μ+a
μ−a +∞
2
≥ (x − μ) fX (x) dx + (x − μ)2 fX (x) dx
−∞ μ+a
μ+a
since (x − μ)2 fX (x) dx ≥ 0.
μ−a
Note that (x − μ)2 ≥ a2 for every x ∈ (−∞, μ − a] ∪ [μ + a, +∞). Hence, for all
x over this region of integration, we have (x − μ)2 fX (x) ≥ a2 fX (x). Therefore,
μ−a μ−a
2
(x − μ) fX (x) dx ≥ a2 fX (x) dx and
−∞ −∞
+∞ +∞
(x − μ)2 fX (x) dx ≥ a2 fX (x) dx.
μ+a μ+a
3
boon·docks (bōōn’dŏks’) plural noun. Rural country; the backwoods. Example usage: The
1965 Billy Joe Royal Song: “Down in the Boondocks;” Source: The American HeritageR Dictionary
of the English Language, Fourth Edition, Houghton Mifflin Company.
So, we get
μ−a +∞
σ2 ≥ (x − μ)2 fX (x) dx + (x − μ)2 fX (x) dx
−∞ μ+a
μ−a +∞
≥ a2 fX (x) dx + a2 fX (x) dx
−∞ μ+a
2 2
= a P (X ≤ μ − a) + a (P (X ≥ μ + a)
= a2 P (|X − μ| ≥ a)
σ2
P (|X − μ| < a) ≥ 1 −
a2
Example:4
The number of customers who visit a car dealer’s showroom on a Saturday morning
is a random variable, X . Using extensive prior experience, the manager, a former
EAS305 student, has determined that the mean and variance of X are 18 and 6.25
respectively.
What is the probability that there will be between 8 and 28 customers on any given
Saturday?
Solution We are given E(X) = μ = 18 and Var(X) = σ 2 = 6.25, so
The above result is the weak law of large numbers. The theorem guarantees that,
as n gets large, the probability that the ratio Xn /n is observed to be “close” to p
approaches one.
There is a related result called the strong law of large numbers that makes a similar
(but stronger statement) about the entire sequence of random variables
{X1 , 12 X2 , 13 X3 , 14 X4 , . . .}
The strong law states that, with probability one, this sequence converges to p. In
this case, we often say that the sequence { n1 Xn } converges to p almost surely.
X −μ
Z= ∼ N (0, 1).
σ
h−1 (z) = σz + μ
d −1
h (z) = σ
dz
Using Theorem 5.2, the probability density function for the random variable Z =
h(X) is
d −1
fZ (z) = fX (h−1 (z)) h (z)
dz
1 (σz+μ)−μ 2
1
= √ e− 2 σ σ for −∞ < σz + μ < ∞
σ 2π
1 1 2
= √ e− 2 z for −∞ < z < ∞
2π
Note that we couldn’t use the cumulative distribution function of X to find the
cumulative distribution function of Z since neither cumulative distribution function
exists in closed form.
X −μ
If X ∼ N (μ, σ 2 ), Theorem 5.16 tells us three things about Z =
σ
1. E(Z) = 0
2. Var(Z) = 1, and
3. Z has a normal distribution
Actually, the first two results would be true even if X did not have a normal
distribution:
Lemma 5.17. Suppose X is any random variable with finite mean E(X) = μ
and finite, nonzero variance Var(X) = σ 2 . Define the random variable
X −μ
Z= .
σ
Then E(Z) = 0 and Var(Z) = 1.
Proof. We have
X −μ
E(Z) = E
σ
1
= E (X − μ)
σ
1
= E[X − μ]
σ
1
= (E[X] − μ)
σ
1
= (0) = 0
σ
and
X −μ
Var(Z) = Var
σ
1
= Var (X − μ)
σ
1
= Var[X − μ]
σ2
1
= Var[X] = 1
σ2
Proof. This is just a special case of Theorem 5.2. If a > 0, then the function
Y = h(X) is strictly increasing and if a < 0, the function Y = h(X) is strictly
decreasing.
We know 2
1 − 12 x−μ
fX (x) = √ e σ
σ 2π
Suppose a > 0, then
FY (y) = P (Y ≤ y) = P (aX + b ≤ y)
y−b
= P X≤
a
y−b
= FX
a
Take derivatives on both sides and use the chain rule to obtain
d d y−b
fY (y) = FY (y) = FX
dy dy a
y−b 1
= fX
a a
2
y−b
− 12 a −μ
1 1 σ
= √ e
σ 2π a
1 y−(aμ+b) 2
1
= √ e− 2 aσ
aσ 2π
Hence, Y ∼ N (aμ + b, a2 σ 2 ).
Now suppose a < 0, then
FY (y) = P (Y ≤ y) = P (aX + b ≤ y)
y−b
= P X≥
a
y−b y−b
= P X> +P X =
a a
y−b y−b
= 1 − FX +P X =
a a
= and since X is continuous.
. .
y−b
= 1 − FX +0
a
Take derivatives on both sides and use the chain rule to obtain
d d y−b
fY (y) = FY (y) = 1 − FX
dy dy a
y−b 1
= −fX
a a
and remembering that a < 0 . . .
2
y−b
− 12 a −μ
1 1 σ
= √ e
σ 2π −a
y−(aμ+b) 2
1 1
= √ e− 2 aσ
−aσ 2π
and since a2 = (−a)2 . . .
1 y−(aμ+b) 2
1
= √ e− 2 −aσ
−aσ 2π
So, again, we have Y ∼ N (aμ + b, a2 σ 2 ).
Question: What happens to the probability distribution of Y = aX + b when
a = 0?
Linear transformations rarely preserve the “shape” of the probability distribution.
For example, if X is exponentially distributed, Z would not be exponentially
distributed. If X was a Poisson random variable, Z would not be a Poisson random
variable. But, as provided by Theorem 5.16, if X is normally distributed, then Z
is also normally distributed. (See Question 5.6 to see that the uniform distribution
is preserved under linear transformations.)
Φ(a) = 1 − Φ(−a).
P (X ≤ 1) = P (X − 0 ≤ 1 − 0)
X −0 1−0
= P ≤
2 2
Since X ∼ N (0, 4) we know from Theorem 5.16 that
X −0
Z= ∼ N (0, 1)
2
so, using Table I in Appendix A, we have
1−0
P (X ≤ 1) = P Z ≤
2
1
= P Z≤ 2 = 0.6915
0.5
0.4
0.3
φ(z )
φ(
0.2
0.1
-2.0 0.0
-1.8
-1.6
-1.4
-1.2
-1.1
-0.9
-0.7
-0.5
-0.3
-0.1
0.1
0.3
0.5
0.7
0.9
1.0
1.2
1.4
1.6
1.8
2.0
z
P (0 ≤ X ≤ 3) = P (0 − 2 ≤ X − 2 ≤ 3 − 2)
0−2 X −2 3−2
= P ≤ ≤
3 3 3
Since X ∼ N (2, 9) we know that
X −2
Z= ∼ N (0, 1)
3
Hence,
P (0 ≤ X ≤ 3) = P − 23 ≤ Z ≤ 1
3
1
= P Z≤ 3 − P Z < − 23
0.5
0.4
0.3
φ(z )
φ(
0.2
0.1
0.0
-2.0
-1.8
-1.6
-1.4
-1.2
-1.1
-0.9
-0.7
-0.5
-0.3
-0.1
0.1
0.3
0.5
0.7
0.9
1.0
1.2
1.4
1.6
1.8
2.0
z
The value of P (X 2 ≥ 14 ) is
(A) 0
1
(B) 4
1
(C) 2
(D) 1
(E) none of the above.
S5.2 Let X be a random variable with E(X) = 0 and E(X 2 ) = 2. Then the value
of E[X(1 − X)] is
(A) −2
(B) 0
(C) 1
(D) 2
(E) Var(X)
(F) none of the above.
S5.3 Let X be a random variable with E(X) = 1 and E(X 2 ) = 2. Then Var(X)
equals
(A) −1
(B) 0
(C) 1
(D) 3
(E) none of the above.
1
S5.4 Suppose that X is a continuous random variable with P (X < −2) = 4 and
P (X < 2) = 12 . Then P (X 2 ≤ 4) equals.
1
(A) 4
1
(B) 2
3
(C) 4
(D) 1
(E) none of the above.
S5.5 Let X be a random variable with E(X) = 2. Then E[(X − 2)2 ] equals
(A) 0
(B) 2
(C) Var(X)
(D) E(X 2 )
(E) none of the above.
S5.6 Let X be a random variable with cumulative distribution function
⎧
⎪
⎨ 0 if x < 0
FX (x) = x if 0 ≤ x ≤ 1
⎪
⎩ 1 if x > 1
At the beginning of the day, the store owner bakes two loaves of rye bread at
a cost of 20 cents per loaf. The bread is sold for 50 cents a loaf and all unsold
loaves are thrown away at the end of the day. The expected daily profit from
rye bread is equal to
(A) 0
(B) 20 cents
(C) 30 cents
(D) 40 cents
(E) none of the above.
S5.17 Suppose that an electronic device has a life length X (in units of 1000
hours) that is considered a continuous random variable with probability
density function
e−x if x > 0
fX (x) =
0 otherwise
Suppose that the cost of manufacturing one such item is $2.00. The man-
ufacturer sells the item for $5.00, but guarantees a total refund if X ≤ 1.
What is the manufacturer’s expected profit per item?
(A) 5e−1
(B) 2e−1
(C) −2 + 5e−1
(D) 5 − 2e−1
(E) none of the above.
S5.18 Let X be a random variable with E(X) = 2 and E(X 2 ) = 5. Then Var(X)
equals
(A) 1
(B) 3
(C) 7
(D) 9
(E) none of the above.
S5.19 Let X be a random variable with E(X) = 1 and Var(X) = 4 then
E((X − 1)2 )
equals
(A) 0
(B) 1
(C) 4
(D) 16
(E) none of the above.
S5.20 A fair coin is tossed 3 times. You win $2 for each head that appears, and
you lose $1 for each tail that appears. Let Y denote your earnings after 3
tosses. The value of E(Y ) is
(A) 0
(B) 3/2
(C) 1
(D) 3
(E) none of the above.
S5.21 Let X be any random variable with E(X) = μ and Var(X) = σ 2 . Consider
the random variable
X −μ
Y = .
σ
Which of the following are always true:
(A) E(Y ) = 0 and Var(Y ) = 1
(B) Y has a normal distribution
(C) P (Y > 0) = 1/2
(D) all of the above are true.
(E) none of the above are true.
S5.22 Let X be a normally distributed random variable with E(X) = 3 and
Var(X) = 9. Then P (X < 0) equals
(A) 0
(B) 0.1587
(C) 0.5
(D) 0.8413
(E) none of the above
S5.23 If Y is a random variable with P (Y ≥ 0) = p, then P (−2Y < 0) equals
(A) p
(B) 1 − p
(C) 0.5
(D) 0
(E) none of the above.
S5.24 Let X be a random variable with
The restaurant will pay a customer a $10 reward if they wait more than 5
minutes for service. Let Y denote the reward for any one customer. The
value of E(Y ), the expected reward, is
(A) 0
(B) 5
(C) 10
(D) 10e−25
(E) none of the above.
S5.26 Suppose that the random variable X has a normal distribution with mean 1
and variance 4. Then, P (|X| ≤ 1) equals
(A) 0.1528
(B) 0.1826
(C) 0.3174
(D) 0.3413
(E) none of the above.
S5.27 Suppose X is a continuous random variable with cumulative distribution
function ⎧
⎪
⎨ 0 if x < 0
FX (x) = x2 if 0 ≤ x ≤ 1
⎪
⎩ 1 if x > 1
Let ln(x) denote the natural logarithm of x. Define the random variable
Y = ln(X). Then P (Y < ln(0.50)) equals
(A) 0
(B) 0.25
(C) 0.50
√
(D) 0.50
(E) none of the above.
S5.28 The life length X (in years) of a printer costing $200 is a continuous random
variable with probability density function
e−x if x ≥ 0
fX (x) =
0 otherwise
The manufacturer agrees to pay a full refund to the buyer if the printer fails
within the first year. If the manufacturer sells 100 printers, how much should
he expect to pay in refunds?
(A) $2706.71
(B) $7357.59
(C) $12,642.41
(D) $17,293.29
(E) none of the above.
6
RANDOM VECTORS
Basic Concepts
Often, a single random variable cannot adequately provide all of the information
needed about the outcome of an experiment. For example, tomorrow’s weather
is really best described by an array of random variables that includes wind speed,
wind direction, atmospheric pressure, relative humidity and temperature. It would
not be either easy or desirable to attempt to combine all of this information into a
single measurement.
We would like to extend the notion of a random variable to deal with an experiment
that results in several observations each time the experiment is run. For example,
let T be a random variable representing tomorrow’s maximum temperature and
let R be a random variable representing tomorrow’s total rainfall. It would be
reasonable to ask for the probability that tomorrow’s temperature is greater than
70◦ and tomorrow’s total rainfall is less than 0.1 inch. In other words, we wish to
determine the probability of the event
Another question that we might like to have answered is, “What is the probability
that the temperature will be greater than 70◦ regardless of the rainfall?” To answer
this question, we would need to compute the probability of the event
B = {T > 70}.
In this chapter, we will build on our probability model and extend our definition of
a random variable to permit such calculations.
Definition
The first thing we must do is to precisely define what we mean by “an array of
random variables.”
155
y
ω 3
1 (4,1)
1 2 3 4 5
x
Ω
Joint Distributions
Now that we know the definition of a random vector, we can begin to use it to assign
probabilities to events. For any random vector, we can define a joint cumulative
distribution function for all of the components as follows:
Definition 6.2. Let (X1 , X2 , . . . , Xn ) be a random vector. The joint cumulative
distribution function for this random vector is given by
In the two-dimensional case, the joint cumulative distribution function for the
random vector (X, Y ) evaluated at the point (x, y), namely
is the probability that the experiment results in a two-dimensional value within the
shaded region shown in Figure 6.2.
Every joint cumulative distribution function must posses the following properties:
1. lim FX1 ,X2 ,...,Xn (x1 , x2 , . . . , xn ) = 0
all xi →−∞
x2
6
5
4
3
2
1
x1
1 2 3 4 5 6
is a nondecreasing function of xi .
As in the case of one-dimensional random variables, we shall identify two ma-
jor classifications of vector-valued random variables: discrete and continuous.
Although there are many common properties between these two types, we shall
discuss each separately.
Discrete Distributions
A random vector that can only assume at most a countable collection of discrete
values is said to be discrete. As an example, consider once again the example on
page 156 where a die is rolled twice. The possible values for either X1 or X2 are in
the set {1, 2, 3, 4, 5, 6}. Hence, the random vector (X1 , X2 ) can only take on one
of the 36 values shown in Figure 6.3.
If the die is fair, then each of the points can be considered to have a probability
1
mass of 36 . This prompts us to define a joint probability mass function for this type
of random vector, as follows:
Definition 6.3. Let (X1 , X2 , . . . Xn ) be a discrete random vector. Then
is the joint probability mass function for the random vector (X1 , X2 , . . . , Xn ).
Referring again to the example on page 156, we find that the joint probability mass
function for (X1 , X2 ) is given by
1
pX1 ,X2 (x1 , x2 ) = 36 for x1 = 1, 2, . . . , 6 and x2 = 1, 2, . . . , 6
Therefore, if we wished to evaluate FX1 ,X2 (3.0, 4.5) we would sum all of the
probability mass in the shaded region shown in Figure 6.3, and obtain
1
FX1 ,X2 (3.0, 4.5) = 12 36 = 13 .
This is the probability that the first roll is less than or equal to 3 and the second roll
is less than or equal to 4.5.
Every joint probability mass function must have the following properties:
1. pX1 ,X2 ,...,Xn (x1 , x2 , . . . xn ) ≥ 0 for any (x1 , x2 , . . . , xn ).
2. ... pX1 ,X2 ,...,Xn (x1 , x2 , . . . xn ) = 1
all x1 all xn
3. P (E) = pX1 ,X2 ,...,Xn (x1 , x2 , . . . xn ) for any event E .
(x1 ,...,xn )∈E
You should compare these properties with those of probability mass functions for
single-valued discrete random variables given in Chapter 5.
Continuous Distributions
Extending our notion of probability density functions to continuous random vectors
is a bit tricky and the mathematical details of the problem is beyond the scope of
an introductory course. In essence, it is not possible to define the joint density as a
derivative of the cumulative distribution function as we did in the one-dimensional
case.
Let Rn denote n-dimensional Euclidean space. We sidestep the problem by defining
the density of an n-dimensional random vector to be a function that when integrated
over the set
{(x1 , . . . , xn ) ∈ Rn : x1 ≤ b1 , x2 ≤ b2 , . . . , xn ≤ bn }
will yield the value for the cumulative distribution function evaluated at
(b1 , b2 , . . . , bn ).
for all (b1 , . . . , bn ) is called the joint probability density function for the random
vector (X1 , . . . , Xn ).
Now, instead of obtaining a derived relationship between the density and the
cumulative distribution function by using integrals as anti-derivatives, we have
enforced such a relationship by the above definition.
Every joint probability density function must have the following properties:
1. fX1 ,X2 ,...,Xn (x1 , x2 , . . . xn ) ≥ 0 for any (x1 , x2 , . . . , xn ).
+∞ +∞
2. ··· fX1 ,...,Xn (x1 , . . . , xn ) dx1 · · · dxn = 1
−∞ −∞
3. P (E) = ··· fX1 ,...,Xn (x1 , . . . , xn ) dx1 · · · dxn for any event E .
E
You should compare these properties with those of probability density functions
for single-valued continuous random variables given in Chapter 5.
x2
b2
a2
x1
a1 b1
This worked for any type of probability distribution. The situation in the multi-
dimensional case is a little more complicated, with a comparable formula given
by
You should be able to verify this formula for yourself by accounting for all of the
probability masses in the regions shown in Figure 6.4.
Example: Let (X, Y ) be a two-dimensional random variable with the following
joint probability density function (see Figure 6.5):
2−y if 0 ≤ x ≤ 2 and 1 ≤ y ≤ 2
fX,Y (x, y) =
0 otherwise
Note that 2 2
(2 − y) dx dy = 1.
1 0
fXY(x,y)
1 2
y
Marginal Distributions
Given the probability distribution for a vector-valued random variable (X1 , . . . Xn ),
we might ask the question, “Can we determine the distribution of X1 , disregarding
the other components?” The answer is yes, and the solution requires the careful
use of English rather than mathematics.
For example, in the two-dimensional case, we may be given a random vector (X, Y )
with joint cumulative distribution function FX,Y (x, y). Suppose we would like to
find the cumulative distribution function for X alone, i.e., FX (x)? We know that
FX,Y (x, y) = P (X ≤ x, Y ≤ y)
and we are asking for
(1) FX (x) = P (X ≤ x).
But in terms of both X and Y , expression 1 can be read: “the probability that X
takes on a value less than or equal to x and Y takes on any value.” Therefore, it
would make sense to say
FX (x) = P (X ≤ x)
= P (X ≤ x, Y ≤ ∞)
= lim FX,Y (x, y).
y→∞
Using this idea, we shall define what we will call the marginal cumulative distri-
bution function:
Definition 6.5. Let (X1 , . . . , Xn ) be a random vector with joint cumulative dis-
tribution function FX1 ,...,Xn (x1 , . . . , xn ). The marginal cumulative distribution
function for X1 is given by
Notice that we can renumber the components of the random vector and call any one
of them X1 . So we can use the above definition to find the marginal cumulative
distribution function for any of the Xi ’s.
Although Definition 6.5 is a nice definition, it is more useful to examine marginal
probability mass functions and marginal probability density functions. For exam-
ple, suppose we have a discrete random vector (X, Y ) with joint probability mass
function pX,Y (x, y). To find pX (x), we ask “What is the probability that X = x
regardless of the value that Y takes on? This can be written as
6
pX1 (2) = P (X1 = 2) = pX1 ,X2 (2, k) = 16 .
k=1
Table 6.1: Joint pmf for daily production
Y
pX,Y (x, y) 1 2 3 4 5
1 .05 0 0 0 0
2 .15 .10 0 0 0
X 3 .05 .05 .10 0 0
4 .05 .025 .025 0 0
5 .10 .10 .10 .10 0
This is hardly a surprising result, but it brings some comfort to know we can get it
from all of the mathematical machinery we’ve developed thus far.
Example: Let X be the total number of items produced in a day’s work at a factory,
and let Y be the number of defective items produced. Suppose that the probability
mass function for (X, Y ) is given by Table 6.1. Using this joint distribution, we
can see that the probability of producing 2 items with exactly 1 of those items being
defective is
pX,Y (2, 1) = 0.15.
To find the marginal probability mass function for the total daily production, X ,
we sum the probabilities over all possible values of Y for each fixed x:
But notice that in these computations, we are simply adding the entries all columns
for each row of Table 6.1. Doing this for Y as well as X we can obtain Table 6.2
So, for example, P (Y = 2) = pY (2) = 0.275. We simply look for the result in
the margin1 for the entry Y = 2.
The procedure is similar for obtaining marginal probability density functions. Re-
call that a density, fX (x), itself is not a probability measure, but fX (x)dx, is. So
1
Would you believe that this is why they are called marginal distributions?
Table 6.2: Marginal pmf’s for daily production
Y
pX,Y (x, y) 1 2 3 4 5 pX (x)
1 .05 0 0 0 0 .05
2 .15 .10 0 0 0 .25
X 3 .05 .05 .10 0 0 .20
4 .05 .025 .025 0 0 .10
5 .10 .10 .10 .10 0 .40
pY (y) 0.4 .275 .225 .10 0
Notice that in both the discrete and continuous cases, we sum (or integrate) over
all possible values of the unwanted random variable components.
y
x
0 a 1
Figure 6.6: The support for the random vector (X, Y ) in the example
The trick in such problems is to insure that your limits of integration are correct.
Drawing a picture of the region where there is positive probability mass (the support
of the distribution) often helps.
For the above example, the picture of the support would be as shown in Figure 6.6.
If the dotted line in Figure 6.6 indicates a particular value for x (call it a), by
integrating over all values of y , we are actually determining how much probability
mass has been placed along the line x = a. The integration process assigns all of
that mass to x = a in one-dimension. Repeating this process for each x yields the
desired probability density function.
Example: Let (X, Y ) be a two-dimensional continuous random variable with joint
probability density function
x+y 0 ≤ x ≤ 1 and 0 ≤ y ≤ 1
fX,Y (x, y) =
0 otherwise
We will leave it to you to check that fX (x) is, in fact, a probability density
! +∞
functionby making sure that fX (x) ≥ 0 for all x and that −∞ fX (x) dx = 1.
Example: Let (X, Y ) be the two-dimensional random variable with the following
joint probability density function (see Figure 6.5 on page 162)
2−y if 0 ≤ x ≤ 2 and 1 ≤ y ≤ 2
fX,Y (x, y) =
0 otherwise
Find the marginal probability density function for X and the marginal probability
density function for Y .
Solution: Let’s first find the marginal probability density function for X . Consider
the case where we fix x so that 0 ≤ x ≤ 2. We compute
+∞
fX (x) = fX,Y (x, y) dy
−∞
2
= (2 − y) dy
1
y=2
y2
= 2y −
2
y=1
1
=
2
If x is outside the interval [0, 2], we have fX (x) = 0. Therefore,
1
0≤x≤2
fX (x) = 2
0 otherwise
To find the marginal probability density function for Y , consider the case where
we fix y so that 1 ≤ y ≤ 2. We compute
+∞
fY (y) = fX,Y (x, y) dx
−∞
2
= (2 − y) dx
0
= (2 − y)x|x=2
x=0
= 4 − 2y
If y is outside the interval [1, 2], we have fX,Y (x, y) = 0 giving us fY (y) = 0.
Therefore,
4 − 2y 1≤y≤2
fY (x) =
0 otherwise
You should also double check that fX and fY are both probability density functions.
{X1 = 1, X2 = 3},
{X1 = 2, X2 = 2},
or
{X1 = 3, X2 = 1}
occurs. So,
3
pZ (4) = pX1 ,X2 (1, 3) + pX1 ,X2 (2, 2) + pX1 ,X2 (3, 1) = 36 .
Continuing in this manner, you should be able to verify that
1 2 3
pZ (2) = 36 ; pZ (3) = 36 ; pZ (4) = 36 ;
4 5 6
pZ (5) = 36 ; pZ (6) = 36 ; pZ (7) = 36 ;
5 4 3
pZ (8) = 36 ; pZ (9) = 36 ; pZ (10) = 36 ;
2 1
pZ (11) = 36 ; pZ (12) = 36 .
Example: Let (X, Y ) be the two-dimensional random variable given in the exam-
ple on page 166. Find the cumulative distribution function for the random variable
Z =X +Y.
Solution: The support for the random variable Z is the interval [0, 2], so
⎧
⎪
⎨ 0 if z < 0
FZ (z) = ? if 0 ≤ z ≤ 2
⎪
⎩ 1 if z > 2
FZ (z) = P (Z ≤ z) = P (X + Y ≤ z).
In other words, we are computing the probability mass assigned to the shaded set
in Figure 6.7 as z varies from 0 to 2.
In establishing limits of integration, we notice that there are two cases to worry
about as shown in Figure 6.8:
Case I (z ≤ 1):
z z−y
FZ (z) = (x + y) dx dy
0 0
1 3
= 3z .
= z 2 − 13 z 3 − 13 .
y
1
z x+
y=
z
x
0 z 1
y y
z
1 1
z
x x
0 z 1 0 1 z
Case I Case II
Theorem 6.1. If X and Y are independent random variables then any event A
involving X alone is independent of any event B involving Y alone.
Case I: Discrete
y
pX,Y (x, y) 0 1
0 0.1 0.3
x 1 0.2 0.1
2 0.2 0.1
2e−2y y ≥ 0
fY (y) =
0 otherwise
We must check the probability density functions for (X, Y ), X and Y for all values
of (x, y).
For x ≥ 0 and y ≥ 0:
We will prove the following results for the case when (X, Y ) is a continuous random
vector. The proofs for the discrete case are similar using summations rather than
integrals, probability mass functions rather than probability density functions.
Theorem 6.3. E(X + Y ) = E(X) + E(Y )
Proof. Using Theorem 6.2:
∞ ∞
E(X + Y ) = (x + y)fX,Y (x, y) dx dy
−∞ −∞
∞ ∞ ∞ ∞
= xfX,Y (x, y) dx dy + yfX,Y (x, y) dx dy
−∞ −∞ −∞ −∞
∞ ∞ ∞ ∞
= xfX,Y (x, y) dy dx + yfX,Y (x, y) dx dy
−∞ −∞ −∞ −∞
∞ ∞ ∞ ∞
= x fX,Y (x, y) dy dx + y fX,Y (x, y) dx dy
−∞ −∞ −∞ −∞
∞ ∞
= x fX (x) dx + y fY (y) dy
−∞ −∞
= E(X) + E(Y )
E[h(X)g(Y )] = E[h(X)]E[g(Y )]
Proof. Using Theorem 6.2:
∞ ∞
E[h(X)g(Y )] = h(x)g(y)fX,Y (x, y) dx dy
−∞ −∞
E[h(X)g(Y )] = E(h(X))E(g(Y ))
E[XY ] = E(X)E(Y )
Now let W = X + Y . . .
Definition 6.11. The correlation coefficient for two random variables X and Y
is given by
Cov(X, Y )
ρ(X, Y ) ≡ " .
Var(X)Var(Y )
Theorems 6.10, 6.11 and 6.12 are stated without proof. Proofs of these results may
be found in the book by Meyer.2
Theorem 6.10. For any random variables X and Y ,
|ρ(X, Y )| ≤ 1.
y
pX,Y (x, y) 0 1
0 0.1 0.3
x 1 0.2 0.1
2 0.2 0.1
Given that no grade B compressors were produced on a given day, what is the
probability that 2 grade A compressors were produced?
Solution:
P (X = 2, Y = 0)
P (X = 2 | Y = 0) =
P (Y = 0)
0.2 2
= =
0.5 5
Given that 2 compressors were produced on a given day, what is the probability
that one of them is a grade B compressor?
Solution:
P (Y = 1, X + Y = 2)
P (Y = 1 | X + Y = 2) =
P (X + Y = 2)
P (X = 1, Y = 1)
=
P (X + Y = 2)
0.1 1
= =
0.3 3
Given that the total life time for the two transistors is less than two hours, what is
the probability that the first transistor lasted more than one hour?
Solution:
P (X > 1, X + Y ≤ 2)
P (X > 1 | X + Y ≤ 2) =
P (X + Y ≤ 2)
We then compute
2 2−x
P (X > 1, X + Y ≤ 2) = 4e−2(x+y) dy dx
1 0
−2
= e − 3e−4
and
2 2−x
P (X + Y ≤ 2) = 4e−2(x+y) dy dx
0 0
= 1 − 5e−4
to get
e−2 − 3e−4
P (X > 1 | X + Y ≤ 2) = = 0.0885
1 − 5e−4
Conditional distributions
Case I: Discrete
Let X and Y be random variables with joint probability mass function pX,Y (x, y)
and let pY (y) be the marginal probability mass function for Y .
We define the conditional probability mass function of X given Y as
pX,Y (x, y)
pX|Y (x | y) =
pY (y)
whenever pY (y) > 0.
Let X and Y be random variables with joint probability density function fX,Y (x, y)
and let fY (y) be the marginal probability density function for Y .
We define the conditional probability density function of X given Y as
fX,Y (x, y)
fX|Y (x | y) =
fY (y)
whenever fY (y) > 0.
Law of total probability
Case I: Discrete
pX (x) = pX|Y (x | y)pY (y)
y
(D) 1
(E) none of the above.
S6.2 Let (X, Y ) be a discrete random vector with joint probability mass function
given by
Then P (Y = 1) equals
(A) 1/3
(B) 1/4
(C) 1/2
(D) 3/4
(E) none of the above.
S6.3 Let (X, Y ) be a random vector with joint probability mass function given by
1 1 1
pX,Y (1, −1) = 4 pX,Y (1, 1) = 2 pX,Y (0, 0) = 4
Let pX (x) denote the marginal probability mass function for X . The value
of pX (1) is
(A) 0
1
(B) 4
1
(C) 2
(D) 1
(E) none of the above.
S6.4 Let (X, Y ) be a random vector with joint probability mass function given by
1 1 1
pX,Y (1, −1) = 4 pX,Y (1, 1) = 2 pX,Y (0, 0) = 4
(D) 1
(E) none of the above.
S6.5 Let (X, Y ) be a continuous random vector with joint probability density
function
1/2 if −1 ≤ x ≤ 1 and 0 ≤ y ≤ 1
fX,Y (x, y) =
0 otherwise
Let
W = X(Y + Z).
Then E(W ) equals
(A) 0
(B) E(X 2 )
(C) 1
(D) 2
(E) none of the above
S6.17 Toss a fair die twice. Let the random variable X represent the outcome of
the first toss and the random variable Y represent the outcome of the second
toss. What is the probability that X is odd and Y is even.
(A) 1/4
(B) 1/3
(C) 1/2
(D) 1
(E) none of the above
S6.18 Assume that (X, Y ) is random vector with joint probability density function
k for 0 ≤ x ≤ 1 and 0 ≤ y ≤ 1 and y ≤ x
fX,Y (x, y) =
0 otherwise
Then P (X + Y ≤ 1) equals
!1!1
(A) 0 0 2e−(x+2y) dx dy
! 1 ! 1−y
(B) 0 0 2e−(x+2y) dx dy
!1!1
(C) 0 y 2e−(x+2y) dx dy
7
TRANSFORMS AND SUMS
Transforms
Notation
Definition 7.2. Let X be a discrete random variable with support on the nonneg-
ative integers. Then the probability generating function for X is given by
199
Example: Let X have a Poisson distribution with parameter α > 0 Then
e−α αk
pX (k) = for k = 0, 1, 2, . . .
k!
∞
Recalling that ex = k
k=0 (x /k !), we get
∞
X e−α αk
gX (z) = E(z ) = zk
k=0
k!
∞
(αz)k
= e−α
k=0
k!
−α αz
= e e
α(z−1)
= e
Proof. (Informal proof) Let g (k) (z) denote the k th derivative of g with respect to
z . We have
gX (z) = E(z X )
∞
= z k P (X = k)
k=0
gX (z) = z 0 P (X = 0) + z 1 P (X = 1) + z 2 P (X = 2) + z 3 P (X = 3) + . . .
gX (0) = (1)P (X = 0)
(1)
gX (z) = P (X = 1) + 2zP (X = 2) + 3z 2 P (X = 3) + 4x3 P (X = 4) + . . .
(1)
gX (0) = P (X = 1)
(3)
gX (z) = 3(2)P (X = 3) + 4(3)(2)zP (X = 4) + 5(4)(3)z 2 P (X = 5) + . . .
(3)
gX (0) = 3(2)P (X = 3)
And, in general
(n)
gX (0) = n!P (X = n)
E(h(X)g(Y )) = E(h(X))E(g(Y ))
Therefore,
gY (z) = z a gX (z).
Proof. We have,
We offer Theorem 7.5 without proof. Except for some technical details, it is a
direct result of Theorem 7.1.
Theorem 7.5. The probability generating function of a random variable uniquely
determines its probability distribution.
Example: Let T have a Bernoulli distribution with parameter p.
P (T = 1) = p
P (T = 0) = 1 − p = q
gT (z) = z 0 q + z 1 p = q + zp
Example: Now suppose, T1 , T2 , . . . , Tn are n independent Bernoulli random vari-
ables each with
P (Ti = 1) = p
P (Ti = 0) = 1 − p = q
One could obtain the probability generating function for X is given by computing
n
n k (n−k)
X k
gX (z) = E(z ) = z p q
k=0
k
Definition 7.3. The kth moment about the origin of a random variable X is
defined as E(X k ).
Definition 7.4. Let X be (almost) any random variable. Then the moment
generating function for X is given by
MX (t) ≡ E(etX ).
Example: Let X have a uniform distribution on the interval (a, b). Then
⎧
⎨ 1
if a ≤ x ≤ b
fX (x) = b − a
⎩
0 otherwise
Therefore,
∞
tX
MX (t) = E(e ) = etx fX (x) dx
−∞
b
1
= etx dx
a b−a
b
1
= tetx dx
t(b − a) a
1 b
= etx
t(b − a) a
etb − eta
= for t = 0
t(b − a)
It’s important to remember that this relationship will only hold for random vari-
ables with support {0, 1, 2, . . .} (that is, those random variables with probability
generating functions).
Example: Let X have an exponential distribution with parameter α > 0. Then
αe−αx if x ≥ 0
fX (x) =
0 otherwise
Therefore,
∞
tX
MX (t) = E(e ) = etx αe−αx dx
0
∞
= αex(t−α) dx
0
= e(tμ+σ
2 t2 /2)
.
(k)
Proof. (Informal proof) Let MX (t) denote the k th derivative of MX with respect
to t. We have
MX (t) = E(etX )
∞
(tX)k
= E
k=0
k!
Theorem 7.8. If MX (t) is the moment generating function of X , and a and b are
constants, then the moment generating function of the random variable Y = aX +b
is
MY (t) = ebt MX (at).
MY (t) = et
2 /2
= et
2 /2
which is the moment generating function of a normal random variable with expected
value equal to zero and variance equal to one. Therefore, we have shown that if
X ∼ N (μ, σ 2 ) then Z = (X − μ)/σ ∼ N (0, 1).
Other transforms of random variables
Laplace transform
Definition 7.5. Let X be (almost) any random variable. Then the Laplace
transform for X is given by
LX (t) ≡ E(e−tX ).
Note that the moment generating function for a continuous random variable X
with probability density function fX (x) is
∞
MX (t) = etx fX (x) dx
−∞
Characteristic function
Definition 7.6. Let X be any random variable. Then the characteristic function
for X is given by
ψX (t) ≡ E(eitX )
√
where i ≡ −1.
where
n
μY = b + ai μi
i=1
and
n
σY2 = a2i σi2 .
i=1
Example: The weights of snails are normally distributed with mean 6 gm. and
variance 4 gm2 . Therefore, it is reasonable to assume that the weights of n snails
chosen at random from the population of all snails are n independent identically
distributed random variables, X1 , X2 , . . . , Xn , with each Xi ∼ N (6, 4).
3 Suppose that 4 snails are selected at random. and we would like to find the
probability that the total weight, T = X1 + X2 + X3 + X4 , will exceed 28
gm.?
Since X1 , X2 , . . . , Xn , are iid with each Xi ∼ N (6, 4), we have T =
4
i=1 Xi ∼ N (24, 16). Therefore,
T − 24 28 − 24
P (T > 28) = P >
4 4
= P (Z > 1) where Z ∼ N (0, 1)
= 1 − P (Z ≤ 1) = 1 − 0.8413 = 0.1587
3 Once again, suppose that 4 snails are selected at random. Let X = 14 (X1 +
X2 + X3 + X4 ) denote the average weight of the 4 snails. Let’s find the
probability that the observed average weight deviates from the expected
weight by more than 2 gm. That is, we want to find P (|X − 6| > 2).
Since X1 , X2 , . . . , Xn , are iid with each Xi ∼ N (6, 4), we have X =
1 4
4 i=1 Xi ∼ N (6, 1). Therefore,
X ∼ N (μ, σ 2 /n).
Proof. First note that E(Sn ) = nμ and Var(Sn ) = nσ 2 . We will get the result by
showing that the moment generating function for Zn must approach the moment
generating function of a N (0, 1) random variable. Since each Xi has the same
moment generating function for all i, we can let M (t) denote the common moment
generating function of each X1 , X2 , . . . , Xn . Then
Since MaY +b (t) = etb MY (at) for any random variable Y and any constants a and
b, we get √ √
(1) MZn (t) = e−μ nt/σ MSn (t/(σ n))
Since each Xi has the same moment generating function for all i, we can let
M (t) denote the common moment generating function of each X1 , X2 , . . . , Xn .
Therefore, MSn (t) = [M (t)]n , and
√ √ n
MZn (t) = e−μ nt/σ
M (t/(σ n))
Now, take the natural log of both sides of equation (1) and use fact (A):
√
−μ nt t
ln MZn (t) = + n ln M √
σ σ n
√ ∞ k
−μ nt E(X k ) t
= + n ln 1 + √ .
σ k=1
k! σ n
Note that the natural log term looks like fact (B) with
∞
k
E(X k ) t
ξ= √ .
k=1
k! σ n
t2
lim ln MZn (t) = .
n→∞ 2
Therefore,
lim MZn (t) = et
2 /2
n→∞
which is the moment generating function for N (0, 1). So limn→∞ P (Zn ≤ a) =
Φ(a).
A proof of the Central Limit Theorem was provided by Liapounoff in 1902. But
DeMoivre (1733) and Laplace (1812) proved a similar result for the following
special case:
Let T1 , T2 , . . . , Tn be a random sample from a Bernoulli distribution, that is let
P (Ti = 1) = p
P (Ti = 0) = 1 − p.
n
Then E(Ti ) = p and Var(Ti ) = p(1 − p). We already know that Sn = i=1 Ti
has a Binomial distribution with parameters n and p, that is
n k
P (Sn = k) = p (1 − p)n−k for k = 0, 1, . . . , n.
k
Simply using the rules for expected value and variance, we get
E(X) = μ
Var(X) = σ 2 /n
z + z5 + z6
g(z) =
3
The value of E(X) equals
(A) 0
1
(B) 3
(C) 1
(D) 4
(E) none of the above.
S7.2 Suppose X ∼ N (2, 9) and Y ∼ N (2, 16) are independent random variables.
Then P (X + Y < 9) equals (approximately)
(A) 0.1587
(B) 0.5199
(C) 0.8413
(D) 0.9772
(E) none of the above.
S7.3 Suppose X is a random variable with moment generating function
Let X1 and X2 be two independent random variables, each with the same
probability distribution as X . Then the moment generating function for
Y = X1 + X2 is
2
(A) e0.5t
2
(B) 2e0.5t
(C) et
2
4
(D) e0.5t
(E) none of the above.
S7.4 An automated assembly line produces incandescent light bulbs. Suppose
that a batch of n bulbs are selected at random. Let T1 , T2 , . . . , Tn be n
independent identically distributed random variables with {Ti = 1} if light
bulb i is good, and {Ti = 0} if light bulb i is bad. Then S = ni=1 Ti is the
total number of good light bulbs in the batch. If p is the probability that any
one light bulb is good, then each Ti has probability mass function
1−p if k = 0
pTi (k) =
p if k = 1
(D) 1
(E) none of the above.
S7.5 If X and Y are independent, normally distributed random variables each
with mean 0 and variance 1, then 3X − Y
(A) has a normal distribution with mean 2 and variance 1
(B) has a normal distribution with mean 0 and variance 10
(C) has a normal distribution with mean 0 and variance 2
(D) has a normal distribution with mean 2 and variance 4
(E) none of the above.
S7.6 Let X and Y be independent random variables with X ∼ N (0, 1) and
Y ∼ N (0, 2). The value of P (X > Y ) is
(A) 0
(B) 0.05
(C) 0.50
(D) 0.95
(E) none of the above.
S7.7 Let X be a random variable with probability generating function
z(1 + z)
g(z) =
2
The value of P (X = 0) is
(A) 0
(B) 1/4
(C) 1/2
(D) 1
(E) none of the above.
S7.8 Let X1 , X2 , X3 , X4 be independent random variables, each with E(Xi ) = 0
and Var(Xi ) = 1 for i = 1, 2, 3, 4. Let Y = X1 + X2 + X3 + X4 . Then
Var(Y ) equals
(A) 1/16
(B) 1/4
(C) 4
(D) 16
(E) none of the above.
S7.9 Let X and Y be independent, normally distributed random variables each
with mean 0 and variance 1. Then the random variables
W1 = X + Y
W2 = X − Y
et (1 + e2t )
MX (t) =
2
The value of E(X) is
(A) 0
(B) 1
(C) 2
(D) 3
(E) none of the above.
S7.11 Let X be a random variable with moment generating function MX (t) =
(1 − t)−1 . Let Y be a random variable with moment generating function
MY (s) = (1 − s)−1 . If X and Y are independent, then X + Y + 1 has a
moment generating function equal to
(A) et (1 − t)−2
(B) (1 − t)−2
(C) 2et (1 − t)−1
(D) es (1 − t)(1 − s)
(E) none of the above.
S7.12 Jack is attempting to fill a 100 gallon water tank using a one gallon bucket.
Jack repeatedly goes to a well at the top of a hill, fills his one gallon bucket
with water and brings that water down to the 100 gallon tank. The amount
of water he carries on any one trip from the well to the tank is normally
distributed with mean 1 gallon and standard deviation 0.1 gallon. The amount
of water carried on any trip is independent of all other trips. The probability
that Jack will need to make more than 100 trips to the well is
(A) 0.05
(B) 0.0636
(C) 0.2580
(D) 0.5
(E) none of the above.
S7.13 Let X and Y be independent random variables with X ∼ N (0, 1) and
Y ∼ N (0, 2). The value of P (X − Y < 1) equals
(A) 0
(B) 0.5
(C) 1
(D) P (X + Y < 1)
(E) none of the above.
S7.14 Let X and Y be independent random variables with X ∼ N (1, 2) and
Y ∼ N (1, 2). Then P (|X − Y | ≤ 1) equals
(A) 0
(B) 0.3830
(C) 0.6915
(D) 1.0
(E) none of the above.
S7.15 Let X be a discrete random variable with probability mass function given
by
P (X = 0) = 13 P (X = 1) = 23
Then the moment generating function for X is
(A) e2t/3
(B) 1
3 (1 + 2et )
(C) 1
2 (1 + et )
1 t
(D) 2 e (1 + e2t )
(E) none of the above.
S7.16 Suppose X1 , X2 , . . . , Xn are independent random variables with each Xi ∼
N (2, 4) for i = 1, 2, . . . , n. Let W = ni=1 Xi and define the random
variable
W − 2n
Z=
c
Then Z ∼ N (0, 1) if and only if
(A) c = 1
(B) c = n
√
(C) c = 2 n
√
(D) c = 2/ n
(E) none of the above.
S7.17 Let X be a random variable with moment generating function
et (1 + et )
MX (t) =
2
The value of E(X) is
(A) 0
2
(B) 3
(C) 1
3
(D) 2
P (Xi = 0) = 1 − p P (Xi = 1) = p
Good and bad luck are often mistaken for good and bad jusment
– ANONYMOUS, Salada TM tea bag tag
Basic Concepts
Random samples
fX (x; θ).
Statistics
227
If a statistic is being used to establish statistical inferences about a particular
unknown parameter θ, then it is often written as θ̂(X1 , X2 , . . . , Xn ).
Note: Any function of a random sample can be called a statistic.
Example: Let X1 , X2 , . . . , Xn be a random sample. You can think of the random
variables X1 , X2 , . . . , Xn as representing the future data values that will be obtained
by making n independent observations of our experiment.
Then the following are statistics:
n
1
1. θ̂(X1 , X2 , . . . , Xn ) = Xi
n i=1
Unbiased estimators
Finding Estimators
Method of maximum likelihood
Maximum likelihood estimators are obtained by finding that value of θ that max-
imizes L for a given set of observations x1 , x2 , . . . , xn . Since the value of θ that
does this will usually vary with x1 , x2 , . . . , xn , θ can be thought of as a function of
x1 , x2 , . . . , xn , namely θ̂(x1 , x2 , . . . , xn ).
To evaluate the properties of θ̂, we can look at its performance prior to actually
making the observations x1 , x2 , . . . , xn . That is we can substitute Xi for xi in the
specification for θ̂ and look at its properties as the statistic
θ̂(X1 , X2 , . . . , Xn ).
For example, one of the properties that we might like to check for is whether θ̂ is
an unbiased estimator for θ (i.e., check to see if E(θ̂) = θ).
pT (0) = 1 − p
pT (1) = p
Note: X denotes the total number of successes from n independent and identically
distributed Bernoulli trials.
Theorem 9.1. If X is a binomial random variable with parameters n and p, then
n k
pX (k) = p (1 − p)n−k k = 0, 1, 2, . . . , n
k
Support: 0, 1, 2, . . . , n
n k
Probability mass function: pX (k) = p (1 − p)n−k
k
k = 0, 1, 2, . . . , n
233
Probability generating function: g(z) = (zp + (1 − p))n
Moment generating function: MX (t) = (et p + (1 − p))n
Moments: E(X) = np Var(X) = np(1 − p)
Notation
Support: 1, 2, . . .
Probability mass function: pX (k) = p(1 − p)k−1 k = 1, 2, . . .
zp 1
Probability generating function: g(z) = for |z| <
1 − z(1 − p) 1−p
et p
Moment generating function: MX (t) =
1 − et (1 − p)
Moments: E(X) = 1/p Var(X) = (1 − p)/p2
Note
Support: 0, 1, 2, . . .
e−α αk
Probability mass function: pX (k) = k = 0, 1, 2, . . .
k!
Probability generating function: g(z) = eα(z−1) for |z| < 1
t −1)
Moment generating function: MX (t) = eα(e
Moments: E(X) = α Var(X) = α
n
1
The expected number of particles per interval of length one is equal to α. Suppose
the interval is divided into n equal segments with at most one particle placed in
any segment. What is the probability that there are exactly k particles in the unit
interval?
Let X denote the number of particles in the unit interval. X is a binomial random
variable with parameters n and p = α/n. Therefore,
n−k
k
n α α
P (X = k) = 1−
k n n
−k n
α n n−1 n−k+1 αk α
= 1− ··· 1−
n n n n k! n
for k = 0, 1, 2, . . . , n.
Taking the limit as n gets large (the segment widths decrease) we obtain
αk −α
lim P (X = k) = 1(1)(1) · · · (1) e
n→∞ k!
e−α αk
=
k!
for k = 0, 1, 2, . . ..
This limiting particle process is called the Poisson Process with
αk e−α
pX (k) = k = 0, 1, 2, . . .
k!
Let Yi denote the interarrival time between occurrences i − 1 and i (see Figure 9.1).
Each of the Yi are independent and identically distributed random variables with
probability density function given by
αe−αy if y ≥ 0
fYi (y) =
0 otherwise
t
i-1 i
W
t
t0
0.40
0.35
Φ(a) 0.30
0.25
ϕ(z) a a
z
0.20
1 2
Φ(a) = ϕ(z) dz = √1 e− 2 z
0.15
0.10 a 2π
dz
0.05 −∞ −∞
0.00 0.1
0.3
0.5
0.8
1.0
1.2
1.5
1.7
1.9
-2.0
-1.8
-1.5
-1.3
-1.1
-0.9
-0.6
-0.4
-0.2
zz
a 0 1 2 3 4 5 6 7 8 9
0.0 0.5000 0.5040 0.5080 0.5120 0.5160 0.5199 0.5239 0.5279 0.5319 0.5359
0.1 0.5398 0.5438 0.5478 0.5517 0.5557 0.5596 0.5636 0.5675 0.5714 0.5753
0.2 0.5793 0.5832 0.5871 0.5910 0.5948 0.5987 0.6026 0.6064 0.6103 0.6141
0.3 0.6179 0.6217 0.6255 0.6293 0.6331 0.6368 0.6406 0.6443 0.6480 0.6517
0.4 0.6554 0.6591 0.6628 0.6664 0.6700 0.6736 0.6772 0.6808 0.6844 0.6879
0.5 0.6915 0.6950 0.6985 0.7019 0.7054 0.7088 0.7123 0.7157 0.7190 0.7224
0.6 0.7257 0.7291 0.7324 0.7357 0.7389 0.7422 0.7454 0.7486 0.7517 0.7549
0.7 0.7580 0.7611 0.7642 0.7673 0.7704 0.7734 0.7764 0.7794 0.7823 0.7852
0.8 0.7881 0.7910 0.7939 0.7967 0.7995 0.8023 0.8051 0.8078 0.8106 0.8133
0.9 0.8159 0.8186 0.8212 0.8238 0.8264 0.8289 0.8315 0.8340 0.8365 0.8389
1.0 0.8413 0.8438 0.8461 0.8485 0.8508 0.8531 0.8554 0.8577 0.8599 0.8621
1.1 0.8643 0.8665 0.8686 0.8708 0.8729 0.8749 0.8770 0.8790 0.8810 0.8830
1.2 0.8849 0.8869 0.8888 0.8907 0.8925 0.8944 0.8962 0.8980 0.8997 0.9015
1.3 0.9032 0.9049 0.9066 0.9082 0.9099 0.9115 0.9131 0.9147 0.9162 0.9177
1.4 0.9192 0.9207 0.9222 0.9236 0.9251 0.9265 0.9279 0.9292 0.9306 0.9319
1.5 0.9332 0.9345 0.9357 0.9370 0.9382 0.9394 0.9406 0.9418 0.9429 0.9441
1.6 0.9452 0.9463 0.9474 0.9484 0.9495 0.9505 0.9515 0.9525 0.9535 0.9545
1.7 0.9554 0.9564 0.9573 0.9582 0.9591 0.9599 0.9608 0.9616 0.9625 0.9633
1.8 0.9641 0.9649 0.9656 0.9664 0.9671 0.9678 0.9686 0.9693 0.9699 0.9706
1.9 0.9713 0.9719 0.9726 0.9732 0.9738 0.9744 0.9750 0.9756 0.9761 0.9767
2.0 0.9772 0.9778 0.9783 0.9788 0.9793 0.9798 0.9803 0.9808 0.9812 0.9817
2.1 0.9821 0.9826 0.9830 0.9834 0.9838 0.9842 0.9846 0.9850 0.9854 0.9857
2.2 0.9861 0.9864 0.9868 0.9871 0.9875 0.9878 0.9881 0.9884 0.9887 0.9890
2.3 0.9893 0.9896 0.9898 0.9901 0.9904 0.9906 0.9909 0.9911 0.9913 0.9916
2.4 0.9918 0.9920 0.9922 0.9925 0.9927 0.9929 0.9931 0.9932 0.9934 0.9936
2.5 0.9938 0.9940 0.9941 0.9943 0.9945 0.9946 0.9948 0.9949 0.9951 0.9952
2.6 0.9953 0.9955 0.9956 0.9957 0.9959 0.9960 0.9961 0.9962 0.9963 0.9964
2.7 0.9965 0.9966 0.9967 0.9968 0.9969 0.9970 0.9971 0.9972 0.9973 0.9974
2.8 0.9974 0.9975 0.9976 0.9977 0.9977 0.9978 0.9979 0.9979 0.9980 0.9981
2.9 0.9981 0.9982 0.9982 0.9983 0.9984 0.9984 0.9985 0.9985 0.9986 0.9986
3.0 0.9987 0.9987 0.9987 0.9988 0.9988 0.9989 0.9989 0.9989 0.9990 0.9990