You are on page 1of 41

Boolean Algebras

Per-Anders Svensson
Contents

Preface 2

1 Introduction 3

2 Partial Orderings and Lattices 4


Exercises for Chapter 2 . . . . . . . . . . . . . . . . . . . 12

3 Bounded, Distributive, and Complemented Lattices 16


Exercises for Chapter 3 . . . . . . . . . . . . . . . . . . . 19

4 Boolean Algebras 21
Exercises for Chapter 4 . . . . . . . . . . . . . . . . . . . 29

5 Boolean Algebras and Electronic Circuits 32


Exercises for Chapter 5 . . . . . . . . . . . . . . . . . . . 35

Answers 38

1
Preface

This text is based on Chapter 15 of the author’s book Abstrakt Algebra, Stu-
dentlitteratur, , which was originally written in Swedish. It is intended to
be used on the course Algebraic Structures I, to complete the course literature
whenever this course is taught in English at Linnæus University.
During the translation of the material into English, the errors that through-
out the years have been spotted in the original Swedish text have been corrected
(and hopefully not replaced by other errors!). Some parts of the text have also
been extended, and a couple of exercises have been added.

Växjö, on a Dark December Evening


in the Year of 

Per-Anders Svensson

2
Chapter 1

Introduction

In the middle of the th century, the British logician and mathematician George
Boole (–) tried to construct a mathematical model, aimed to describe
human reasoning. His result became an example of class of algebraical struc-
tures that today go by the name of Boolean algebras. Speaking in short terms,
Boole’s idea was to translate logical derivations of propositions and relations,
into algebraical manipulations of mathematical expressions. In s, one found
out that the theory of Boolean algebras had applications in constructing electri-
cal circuits, in which a number of relays are included. This also applies to the
integrated circuits and computer chips, that are widely used today.
The objective with this text is to get a little bit acquainted to Boolean
algebras. When presenting the theory of this topic, one usually starts with a
more general class of algebraical structures, which is what we will do in the first
few chapters.

3
Chapter 2

Partial Orderings and


Lattices

We recall that an equivalence relation is a relation that is reflexive, symmetric,


and transitive. We will shortly introduce another kind of relation, which we
also wish to be reflexive and transitive, but where the condition of symmetry is
replaced by the following property.

Definition 2.1. Let X be a set and R a relation on X. We say that R is


antisymmetric, if x R y and y R x implies x = y, for all x, y ∈ X.

The word ‘antisymmetric’ perhaps makes one to believe that it is the same thing
as ‘not symmetric’. This is however not the case, since the relation of equality =
defined on i.e. the set R of real numbers, is both symmetric and antisymmetric.

Example 2.1. The usual inequality relation ≤ on the set of real numbers is
antisymmetric, since if a ≤ b and b ≤ a, then a = b. ♦

Definition 2.2. A relation R on a set X is called a partial ordering if it


is reflexive, antisymmetric, and transitive. For a partial ordering, we write
x  y (which is pronounced ‘x is less than or equal to y’) instead of x R y. The
notation x ≺ y (‘x is less than y’) is used when we want to express that x  y
but x 6= y. If  is a partial ordering on X, then the pair (X, ) is called a
partially ordered set or poset, for short.

When refereing to a poset (X, ), we will often be sloppy in our notation and
simply write X. Sometimes a  b and a ≺ b will be written b  a and b ≻ a,
respectively.1 If, for a certain a, b ∈ X, the relation a  b does not hold, we will
write a  b.

Example 2.2. The relations ≤ and ≥ are both partially orderings on the set R
of real numbers. ♦
1 Since we have agreed upon how to pronounce a  b and a ≺ b, it does not take a rocket

scientist to find out how a  b and a ≻ b are pronounced.

4
Example 2.3. The power set P(M ), i.e. the set of all subsets of the set M , is
a poset with respect to the subset relation ⊆. ♦

Example 2.4. Recall that if a and b are integers, then we write a | b, if there is
an integer k such that b = ak. From elementary number theory we recall that
this divisibility relation has among other things the following properties: If a, b,
and c denote positive integers, then it is easy to prove2 that
(i) a | a
(ii) if a | b and b | a, then a = b
(iii) if a | b and b | c, then a | c
From this we conclude that (Z+ , |) is a poset. ♦

Example 2.5. Let H denote the set of all subgroups of a given group. Then
the subgroup relation ≤ is a partial ordering on H, whence (H, ≤) is a poset. ♦

One way to illustrate a poset X graphically (at least when it is finite), is by


means of a so-called Hasse diagram. A Hasse diagram is a graph, in which all
of its nodes represent the different elements of X. If x ≺ y, and if there is no
z ∈ X such that x ≺ z ≺ y, we draw an edge in the graph between the nodes
that represent x and y. Since x ≺ y, i.e. x is the smaller of the two elements x
and y, we will place the node that represents y higher up in the diagram, than
the node that represents x. A Hasse diagram for the poset (P({1, 2, 3}), ⊆) is
shown in Figure 2.1 on page 5.
{1, 2, 3}

{1, 2} {1, 3} {2, 3}

{1} {2} {3}


Figure 2.1

Conversely we can use a Hasse diagram to define a partially ordering on a finite


set X. From the diagram it will easily be seen if the relation x ≺ y holds for two
given elements x, y ∈ X. This is the case, if and only if it is possible to travel
from the node that represent x to the node that represents y, only by moving
upwards in the diagram.

Example 2.6. From the Hasse diagram in Figure 2.2 on the next page we can
see that j ≺ b, since we may travel from j to b in the diagram, only by moving
upwards (for example via the nodes h and e). We can however not say the same
thing about for instance k and d. Thus k  d. ♦
2 And we ask you to do this!

5
a b c

d e f

g h i

j k l

m
Figure 2.2

Since a partially ordering in some way compares the ‘size’ between the elements
in a set, it would be of an interest to know whether the set contains an element
that is the ‘greatest’ or the ‘least’ one.

Definition 2.3. Let (X, ) be a poset.


(i) An element a ∈ X is called a universal upper bound of X, if x  a for
all x ∈ X. Similarly we say that a is a universal lower bound, if a  x
for all x ∈ X.
(ii) By a maximal element of X we mean an element b ∈ X such that
b  x ⇒ b = x for all x ∈ X. If we instead have b  x ⇒ b = x for all
x ∈ X, then b is called a minimal element of X.

An element of a poset is thus a universal upper bound, if ‘all other elements in


the set are smaller’, and it is a maximal element if ‘no other element is larger’.
At a first glance, this may look like being the same thing, and in Exercise 2.14 we
ask you to prove that a universal upper (lower) bound is a maximal (minimal)
element. The converse is however not true, as seen in the following example.

Example 2.7. Consider the partially ordered set X, defined by the Hasse dia-
gram in Figure 2.2 on page 6. The elements a, b, and c are all maximal elements
of X. None of these three elements is however a universal upper bound of X.
(One reason that a is not a universal upper bound, is that f  a.)
There is no universal lower bound in X either, but on the other hand there
are two minimal elements (j and m). ♦

Example 2.8. The poset (R, ≤) contains neither a universal upper bound, nor
a universal lower bound. Therefore there are also no any maximal or minimal
elements in R, with respect to ≤. ♦

Example 2.9. The empty set ∅ is a universal lower bound of the partially
ordered set (P(M ), ⊆), since ∅ is a subset of every set in P(M ). All elements
in P(M ) are further subsets of M , so M is a universal upper bound (with
respect to ⊆) of P(M ). ♦

Example 2.10. The poset (Z+ , |) has no universal upper bound a, since if
such an a existed, it would have to be a multiple of all positive integers, which
is impossible. On the other hand, the integer 1 is a universal lower bound,
because we have 1 | b for all b ∈ Z+ . ♦

6
Example 2.11. If we in the previous example replace the set Z+ of positive
integers by the set N of natural numbers (including zero), then (N, |) will be a
poset in which a universal upper bound exists. We will actually have that 0 is
a universal upper bound, since a | 0 for all natural numbers a. ♦

Definition 2.4. Let (X, ) be a poset and Y a subset of X.

(i) An element s ∈ X is called an upper bound of Y , if y  s for all y ∈ Y .


If s is an upper bound of Y , such that s  t for all other upper bounds t
of Y , then s is called a supremum of Y .

(ii) An element i ∈ X is called a lower bound of Y , if i  y for all y ∈ Y .


If i is a lower bound of Y , such that j  i for all lower bounds of Y , then i
is called an infimum of Y .

In the literature, you sometimes see the concepts ‘least upper bound’ (l.u.b.) for
supremum, and ‘greatest lower bound’ (g.l.b.) for infimum.

Example 2.12. Let Y be the open interval

]0, 1[= {x ∈ R | 0 < x < 1}

of R. With respect to the ordering ≤ on R, we have that 1 is a supremum of Y ,


and that 0 is an infimum. ♦

Example 2.13. Let X = {a, b, c, d, e, f, g, h, i} and suppose  is a partial or-


dering on X, defined by the Hasse diagram in Figure 2.3 on page 7. Consider
the subset Y = {d, f, g}, whose elements for the sake of clarity is written in
boldface in the diagram. The elements a, b, c, and d are all upper bounds of Y .
But there are no lower bounds of Y . (For instance, h is not a lower bound of
Y , since h  f .)
Since there are no lower bounds Y , there can of course not be any infimum
of Y either. Among the four upper bounds of Y , we see that d is the least one,
so d is a supremum of Y . ♦
a

b c

d e

f g

h
Figure 2.3

Theorem 2.1. Let (X, ) be a poset and Y a subset of X. Then X contains


at most one supremum and at most one infimum of Y .

7
Proof. Suppose s1 and s2 are suprema of Y . Regarding s1 as a supremum and s2
simply as an upper bound of Y , we must have s1  s2 . If we instead consider s2
as a supremum and s1 as an upper bound, we obtain s2  s1 . Now s1 = s2
follows, by the antisymmetry of .
The proof of uniqueness of an infimum of Y (if it exists) is similar, and is
left to the reader (see Exercise 2.15).

Corollary. A poset X contains at most one universal upper (lower) bound.

Proof. Put Y = X in Theorem 2.1. Then a supremum of Y becomes the same


thing as an universal upper bound of X. Similarly every infimum of Y will be
an universal lower bound of X.
Since there can be at most one supremum and at most one infimum for any
subset Y of a poset, we can without ambiguity talk about the supremum and
the infimum of Y . We also introduce the notations sup Y and inf Y , respectively,
for those elements (in case they exist). If Y = {y1 , y2 , . . . , yn } is a finite set, we
will also write sup(y1 , y2 , . . . , yn ) and inf(y1 , y2 , . . . , yn ).
By the corollary above, we can also talk about the universal upper bound
and the universal lower bound of a poset, since there can be at most one such
element of each kind.

Example 2.14. Let X = {1, 2, 3, 4, 6, 9, 12, 18, 36} be the set of all positive
divisors of 36. The Hasse diagram in Figure 2.4 on page 8 illustrates X, regarded
as a poset with respect to the divisibility relation. Both of the subsets Y =
{2, 12, 18} and Z = {4, 6, 9} has a supremum and an infimum; to be precise,
sup Y = 36, inf Y = 2, sup Z = 36, and inf Z = 1. ♦
36

12 18

4 6 9

2 3

1
Figure 2.4

From now on, we will be interested in posets, in which supremum and infimum
exists for each pair of elements in the set.

Definition 2.5. A non-empty poset L is called a lattice, if sup(a, b) and


inf(a, b) exists for all a, b ∈ L.

Example 2.15. In Example 2.3, we concluded that the power set P(M ) of a
set M is a poset with respect to ⊆. Let A and B be two subsets of M . Then
inf(A, B) exists and equals A ∩ B, since A ∩ B is a subset of both A and B (and
therefore a lower bound of these two elements), and if C ⊆ A and C ⊆ B, i.e.

8
if C is any lower bound of {A, B}, then C ⊆ A∩B. By a similar argument, which
we ask the reader to deal with in Exercise 2.18, we have that sup(A, B) = A∪B.
Hence (P(M ), ⊆) is a lattice. ♦

Example 2.16. By Example 2.4, (Z+ , |) is a poset. We claim that this in fact
is a lattice, since sup(a, b) and inf(a, b) both exist for all a, b ∈ Z+ .
Let us prove this claim: If s = sup(a, b) exists, it has, to begin with, be an
upper limit of {a, b}, which in this case means that a | s and b | s. Thus s must be
a common multiple of a and b. But sup(a, b) is the least upper bound of a and b
(with respect to |). We conclude that sup(a, b) exists and equals lcm(a, b), the
least common multiple of a and b. By a similar argument, inf(a, b) = gcd(a, b),
see Exercise 2.19. ♦

If a poset L is a lattice, then the supremum and infimum exists, and are uniquely
determined, for each pair of elements in L. It is therefore close at hand to define
two binary operations on L.

Definition 2.6. Let L be a lattice. Then the meet a ∧ b and join a ∨ b of the
two elements a, b ∈ L are defined as
a ∧ b = inf(a, b) and a ∨ b = sup(a, b),
respectively.

Example 2.17. As we can see from Example 2.15 the join and meet on the
powerset P(M ) of a set M corresponds to union and intersection of sets, re-
spectively. ♦

Example 2.18. Figure 2.5 on page 9 shows a Hasse diagram for a partial or-
dering in a set L = {a, b, c, d, e, f, g}. By examining the Cayley tables for ∧ and
∨ in Figure 2.6 on the following page, we can see that L is a lattice. ♦
a

b c

d e f

g
Figure 2.5

Example 2.19. The poset described in Example 2.14 is a lattice (verify this!).
Here we for instance have 3 ∨ 4 = 12, 2 ∧ 9 = 1, and (4 ∨ 6) ∧ 9 = 12 ∧ 9 = 3. ♦

Suppose A and B are subsets of a given set M . Then one easily convince oneself
(for instance by drawing a Venn diagram) that the three statements A ⊆ B,
A ∩ B = A, and A ∪ B = B all are equivalent. Thus we can relate the binary
operations ∩ and ∪ on P(M ) to the partial ordering ⊆ on P(M ). In the
following Theorem (which we will refer to very frequently), we relate the binary
operations ∧ and ∨ on a lattice L to the partial ordering  on L, in a similar
manner.

9
∧ a b c d e f g ∨ a b c d e f g
a a b c d e f g a a a a a a a a
b b b e d e g g b a b a b b a b
c c e c g e f g c a a c a c c c
d d d g d g g g d a b a d b a d
e e e e g e g g e a b c b e c e
f f g f g g f g f a a c a c f f
g g g g g g g g g a b c d e f g

Figure 2.6

Theorem 2.2. Let a and b be elements of a lattice L. Then the following


statements are equivalent:

(i) a  b

(ii) a ∧ b = a

(iii) a ∨ b = b.
Especially a ∧ a = a ∨ a = a for all a ∈ L, i.e. all elements in L are idempotent
with respect both ∧ and ∨.

Proof. The theorem follows immediately from the definition of ∧ and ∨.


Next we will state some fundamental properties of the binary operations ∧ and ∨
on a lattice.

Theorem 2.3. Let L be a lattice. Then


(i) ∧ are ∨ commutative

(ii) ∧ are ∨ associative

(iii) a ∧ (a ∨ b) = a and a ∨ (a ∧ b) = a for all a, b ∈ L, i.e. the so-called


absorption laws are fulfilled.

Proof. (i) For all a, b ∈ L we have a ∧ b = inf(a, b) = inf(b, a) = b ∧ a. A similar


argument shows that a ∨ b = b ∨ a.
(ii) Both a ∧ (b ∧ c) and (a ∧ b) ∧ c equals inf(a, b, c), which shows that ∧ is
associative. In the same way, one shows that ∨ is associative.
(iii) Since a ∨ b = sup(a, b), we must have a  a ∨ b. From Theorem 2.2 it
follows that a ∧ (a ∨ b) = a, and the first of the two absorption laws is proved.
The proof of the second law is similar.

As we have seen so far, we can see that if we start with a poset L that is
a lattice, we will end up with an algebraic structure (L, ∧, ∨), where the two
binary operations ∧ and ∨ have the properties stated in Theorem 2.3. We will
now show that is possible to go the other way around.

10
Theorem 2.4. Let (L, ∧, ∨) be an algebraic structure, where the binary opera-
tions ∧ and ∨ fulfills the laws stated in Theorem 2.3. Then the relation  on L,
defined by
a  b ⇐⇒ a ∧ b = a (2.1)
for all a, b ∈ L, is a partial ordering on L, such that sup(a, b) and inf(a, b) exists
for each pair of elements a, b ∈ L and are equal to a ∨ b and a ∧ b, respectively.
In other words, the poset (L, ) is a lattice.

Proof. We start by showing that  is reflexive, antisymmetric, and transitive.


Reflexivity. Pick any a ∈ L. To make the conclusion that  is reflexive, we
must show that a ∧ a = a. By putting b = a in the second absorption law of
Theorem 2.3(iii), we obtain a = a ∨ (a ∧ a), which in turn yields

a ∧ a = a ∧ [a ∨ (a ∧ a)].

By putting b = a ∧ a in the first absorption law of Theorem 2.3(iii), we see the


right-hand side above is equal to a. This shows that  is reflexive.
Antisymmetry. Suppose a, b ∈ L fulfills a  b and b  a. Then a ∧ b = a and
b ∧ a = b, according to Theorem 2.2. But ∧ is commutative, so a ∧ b = b ∧ a,
and we see that a = b. Thus  is antisymmetric.
Transitivity. Let a, b, c ∈ L fulfill a  b and b  c. Then Theorem 2.2 yields
a ∧ b = a and b ∧ c = b. These two equalities, together with the fact that ∧ is
associative, implies

a ∧ c = (a ∧ b) ∧ c = a ∧ (b ∧ c) = a ∧ b = a,

and yet another application of Theorem 2.2 yields a  c. Thus  is transitive.


Thereby we have shown (L, ) is a poset.
We now turn into proving that (L, ) is a lattice, or in other words, to show
that s = a ∨ b = sup(a, b) and i = a ∧ b = inf(a, b) for all a, b ∈ L. By using one
of the absorption laws, we find that

a ∧ s = a ∧ (a ∨ b) = a,

and thereby a  s (Theorem 2.2 once again). In a similar way b  s is obtained.


Thus s is an upper bound of {a, b}. But we want to show that s is the least
upper bound of {a, b}. Therefore, let t be any upper bound of {a, b}. We want
to show that s  t, i.e. s ∧ t = s. Being an upper bound of {a, b}, we know
that t fulfills a ∧ t = a and b ∧ t = b. By using the different laws of Theorem 2.3,
it now follows that

t = t ∨ (t ∧ a) = (a ∧ t) ∨ t = a ∨ t,

and similarly, t = b ∨ t. Putting this together finally yields

s ∧ t = s ∧ (a ∨ t)
= s ∧ [a ∨ (b ∨ t)]
= s ∧ [(a ∨ b) ∨ t]
= s ∧ (s ∨ t)
= s,

11
which is exactly what we wanted to prove. Thus sup(a, b) exists and equals
s = a ∨ b. To show that we also have i = inf(a, b) = a ∧ b is left to the reader
(see Exercise 2.20).

Since ∧ and ∨ both are associative binary operations, expressions in the form
a1 ∧ a2 ∧ · · · ∧ V
an and a1 ∨ a
W2n∨ · · · ∨ an are well-defined. We will sometimes use
n
the notations i=1 ai and i=1 ai for those kinds of expressions.

Example 2.20. In the poset (P(M ), ⊆), the binary operations ∧ and ∨ cor-
responds to ∩ and ∪, respectively, see Example 2.15. This means that the
algebraic structure (P(M ), ∩, ∪) is a lattice. ♦

Example 2.21. Looking back at Example 2.16 we see that a ∧ b = gcd(a, b)


and a ∨ b = lcm(a, b) in the poset (Z+ , |). The algebraic structure (Z+ , gcd, lcm)
is thus a lattice. ♦

When taking another look at Theorem 2.3, we notice that all the laws of ∧
and ∨ come in pairs; both ∧ and ∨ are commutative and associative, and if one
replace each occurrence of ∧ by ∨ (and vice versa) in the first absorbtion law,
we obtain the second one. This implies that we for every formula, that we can
derive by using the laws of Theorem 2.3, will obtain an extra valid formula “into
the bargain”, so to speak, by shifting the occurrences of ∧ and ∨ in the formula
we have derived. This leads to the following principle.

The Principle of Duality. Each true formula in a lattice (L, ∧, ∨) remains true,
if each occurrence of ∧ in the formula is replaced by ∨, and each occurrence of ∨
by ∧.

When we shift the occurrences of ∧ and ∨ in a formula, we say that we dualize


the formula. The formula we obtain after dualizing, is called the dual of the
original formula. We can of course dualize any expression, not necessarily just
those which are true formulas.

Example 2.22. Let a and b be two elements of a lattice L. We will show that
the dual of a  b is a  b. First, a  b is equivalent to a ∧ b = a (Theorem 2.2).
The dual of this formula is a ∨ b = a, i.e. b ∨ a = a, and by using Theorem 2.2
again, we obtain b  a, that is a  b. ♦

Exercises for Chapter 2


2.1. Which ones of the relations below are antisymmetric?
(a) x R y ⇐⇒ x + y > 10
(b) x R y ⇐⇒ |x − y| = 1
(c) x R y ⇐⇒ x2 = y 2
(d) x R y ⇐⇒ x + y ∈ Z
(e) x R y ⇐⇒ x − y ≥ 0
(f) x R y ⇐⇒ x = 1, y = 2

2.2. Are < and > antisymmetric relations on R?

12
2.3. Prove that (Z+ , |) is a poset (see Example 2.4).

2.4. A poset (X, ) are said to be totally ordered, if for every a, b ∈ X we have
a  b or b  a. Decide which ones of the posets below that are totally ordered.

(a) (R, ≤)
(b) (Z+ , |)
(c) (P(M ), ⊆)

2.5. Let R be a non-trivial ring and I the set of all proper ideals of R. Then (I, ⊆)
is a poset. Are there any maximal or minimal elements in I?

2.6. Let (X1 , 1 ) and (X2 , 2 ) be posets. Define a relation  on the Cartesian
product X1 × X2 , according to

(a1 , a2 )  (b1 , b2 ) ⇐⇒ a 1 c1 and a2 2 b2

for all (a1 , a2 ), (b1 , b2 ) ∈ X1 × X2 . Show that (X1 × X2 , ) is a poset.

2.7. Decide, by using the Hasse diagram in Figure 2.2 on page 6, whether each one
of the expressions below exist. If it exists, compute it!

(a) a ∨ h
(b) a ∧ h
(c) (b ∧ d) ∨ k
(d) (d ∧ j) ∨ i
(e) (j ∨ m) ∧ (g ∨ h)
(f) (a ∧ b) ∨ (j ∨ f )

2.8. Draw a Hasse diagram for the divisibility relation | on the set

X = {2, 3, 6, 9, 10, 14, 18, 20}.

2.9. Let X = {1, 2, . . . , 12} and suppose  is a partial ordering on X, defined by the
Hasse diagram in Figure 2.7 on page 13. Furthermore, let Y = {2, 4, 6, 9}.
(a) Find the maximal and minimal elements of X, with respect to .
(b) Find all upper bounds and lower bounds of Y .
(c) Compute sup Y and inf Y , in case they exist.
(d) What relation is described by the Hasse diagram?

8 12

10 4 6 9

5 2 3 7 11

1
Figure 2.7

2.10. What are a ∧ b and a ∨ b, if a and b are elements in the poset (R, ≤)?

13
2.11. Dualize the following expressions:
(a) a ∨ b = c ∧ d
(b) a ∨ (b ∧ c) = (a ∨ b) ∧ c
(c) a ∧ b  c ∨ d
(d) (a ∨ b) ∧ a  a ∨ (b ∧ c).

2.12. Write down the Cayley tables for ∧ and ∨ for the lattice that is defined by the
Hasse diagram in Figure 2.8 on page 14.
a

c d e

f
Figure 2.8

2.13. The Cayley tables in Figure 2.9 on page 14 define two binary operations ∧ and
∨ on the set L = {a, b, c, d, e, f } in such a way that (L, ∧, ∨) becomes a lattice.
Draw a Hasse diagram for L.

∧ a b c d e f ∨ a b c d e f
a a f c a a f a a e a d e a
b f b c b b f b e b b d e b
c c c c c c c c a b c d e f
d a b c d e f d d d d d d d
e a b c e e f e e e e d e e
f f f c f f f f a b f d e f

Figure 2.9

2.14. Let X be a poset. Show that a universal upper (lower) bound of X is a maximal
(minimal) element of X.

2.15. Show that a subset of a poset have at most one infimum (see Theorem 2.1).

2.16. Let X be a poset. Show that every element in X is both an upper bound and
a lower bound of the subset ∅ of X. What can be said about supremum and
infimum of ∅?

2.17. Let L be a lattice and M a finite non-empty subset of L. Show that sup M and
inf M both exist.

2.18. Let M be a set. Show that sup(A, B) = A ∪ B for all A, B ∈ P(M ) (see
Example 2.15).

2.19. Show that inf(a, b) exists and is equal to gcd(a, b), for each pair of elements a
and b in the poset (Z+ , |) (see Example 2.16).

2.20. Finish the proof of Theorem 2.4, by showing that inf(a, b) exists for all a, b ∈ L
and equals a ∧ b.

14
2.21. Show that we may as well have the relation

a  b ⇐⇒ a ∨ b = a

instead of (2.1) in the statement of Theorem 2.4.

2.22. Let L be a lattice. Show that

[(a ∨ b) ∧ (a ∨ c)] ∨ [(a ∨ b) ∧ (b ∨ c)] = a ∨ b

for all a, b, c ∈ L.

15
Chapter 3

Bounded, Distributive, and


Complemented Lattices

We begin this chapter by recalling some fundamental laws of the binary opera-
tions intersection ∩ and union ∪ of sets.

Theorem 3.1. Let A, B, and C be subsets of a given set M . Then

(i) A ∩ (B ∩ C) = (A ∩ B) ∩ C; A ∪ (B ∪ C) = (A ∪ B) ∪ C

(ii) A ∩ B = B ∩ A; A∪B =B∪A


(iii) A ∩ M = A; A ∪ ∅ = A

(iv) A ∩ ∅ = ∅; A∪M =M

(v) A ∩ (A ∪ B) = A; A ∪ (A ∩ B) = A

(vi) A ∪ ∁A = M ; A ∩ ∁A = ∅

(vii) A ∩ (B ∪ C) = (A ∩ B) ∪ (A ∩ C); A ∪ (B ∩ C) = (A ∪ B) ∩ (A ∪ C).

In the theorem above, we especially recognize the associative laws (i), commuta-
tive laws (ii), and absorption laws (v), that together tells us that (P(M ), ∩, ∪)
is a lattice. But the theorem also states some more properties of ∩ and ∪. In this
chapter we will try to generalize those properties to a general lattice (L, ∧, ∨).
We will begin by generalizing (iii) and (iv).

Definition 3.1. A lattice L is said to be bounded, if it contains both a univer-


sal upper bound and a universal lower bound. We denote the universal upper
bound by 1L , and the universal lower bound by 0L .

Example 3.1. Since Z+ contains no universal upper bound with respect to the
divisibility relation | (see Example 2.10), the lattice (Z+ , |) is not bounded. On
the other hand, the lattice in Example 2.14 is bounded, having 1 and 36 as its
universal lower and upper bound, respectively. ♦

16
A consequence of Theorem 3.1(iv) is that A ∩ ∅ = ∅ and A ∪ M = M for all
A ∈ P(M ). According to Theorem 2.2, this is means that ∅ is the universal
lower bound and M the universal upper bound (with respect to ⊆) in P(M ),
a fact that was also established in Example 2.9.
Thus in the lattice L = P(M ), we have 1L = M and 0L = ∅. In Theo-
rem 3.1(iii) we see that the universal upper bound M in P(M ) is an identity
element of the binary operation ∩, and that the universal lower bound ∅ is an
identity element of the binary operation ∪. This fact holds in any bounded
lattice.

Theorem 3.2. Let (L, ∧, ∨) be a bounded lattice, with universal bounds 1L


(upper) and 0L (lower). Then 1L and 0L are identity elements with respect to ∧
and ∨, respectively.
Conversely, if (L, ∧, ∨) is a lattice such that L contains identity elements 1L
and 0L for ∧ and ∨, respectively, then L is bounded, with 1L and 0L as the
universal upper bound and universal lower bound, respectively.

Proof. Suppose L is bounded. Then the universal lower bound 0L fulfills 0L  a


for all a ∈ L, which according to Theorem 2.2 is equivalent to a ∨ 0L a = a for all
a ∈ L. Thus 0L is an identity element with respect to the binary operation ∨.
Similarly, 1L is proven to be an identity element with respect to ∧.
On the other hand, if ∧ and ∨ possesses identity elements 1L and 0L , respec-
tively, then for all x ∈ L we have x ∧ 1L = x. By Theorem 2.2, this means that
x  1L for all x ∈ L. Thus 1L is a universal upper bound of L. In the same way
one shows that the identity element 0L with respect to ∨ is a universal lower
bound of L. Hence L is bounded.

Example 3.2. As we have seen, the lattice (P(M ), ∩, ∪) is bounded, having ∅


and M as its universal bounds. Therefore ∅ and M are identity elements with
respect to ∪ and ∩, respectively, as claimed in Theorem 3.1(iv). This is due to
Theorem 3.2. ♦

The statement a  1L is true for all elements in a bounded lattice L. When


dualizing the expression a  1L , we should replace  by , see Example 2.22.
But if we do this, the result becomes a  1L , for all a ∈ L, a statement that
is false. By the Principle of Duality on page 12, a true statement has a true
statement as its dual. It would be nice of this also was true in bounded lattices.
Therefore we extend the Principle of Duality to bounded lattices by replacing
every occurrence of 0L by 1L and vice versa. Thus the dual of a  1L for all
a ∈ L (which is true), will be a  0L for all a ∈ L (which still is true).

Example 3.3. The dual of the expression a ∧ b = 0L is a ∨ b = 1L . ♦

According to Theorem 3.1(vii), ∩ and ∪ are distributive over one another. We


will now generalize this property to general lattice.

Definition 3.2. A lattice L is said do be distributive, if the distributive laws


a ∨ (b ∧ c) = (a ∨ b) ∧ (a ∨ c)
a ∧ (b ∨ c) = (a ∧ b) ∨ (a ∧ c)
holds, for all a, b, c ∈ L.

17
Actually, it is enough to only require that one of the two distributive laws is
fulfilled, since one can show that they are equivalent (see Exercise 3.4).
The generalized distributive laws
n
! n n
! n
^ ^ _ _
a∨ bi = (a ∨ bi ) and a ∧ bi = (a ∧ bi ) (3.1)
i=1 i=1 i=1 i=1

are easily shown (by induction) to be valid in a distributive lattice, as well as


the formulae !  n 
m
^ ^ ^m ^n
ai ∨  bj  = (ai ∨ bj ) (3.2)
i=1 j=1 i=1 j=1

and !  
m
_ n
_ m _
_ n
ai ∧ bj  = (ai ∧ bj ). (3.3)
i=1 j=1 i=1 j=1

Example 3.4. We have already noted (cf. Theorem 3.1(vii)) that P(M ) is a
distributive lattice. In Exercise 3.3 we ask you to prove that the same holds for
the lattice (Z+ , gcd, lcm). ♦

Example 3.5. Consider the two lattices L and M , defined by the Hasse dia-
grams in Figure 3.1 on page 18. For the lattice L (defined by the Hasse diagram
to the left in the figure) we have

a ∧ (b ∨ c) = a ∧ 1L = a,

while on the other hand

(a ∧ b) ∨ (a ∧ c) = 0L ∨ 0L = 0L .

Hence L is not distributive. We leave as an exercise for the reader (Exercise 3.1)
to show that the gitter M (defined by the right diagram) on the other hand is
distributive. ♦
1L 1M

a b c a b

0L 0M
Figure 3.1

Theorem 3.3 (Cancellation Law). Let a, b, and c be elements in a distributive


lattice L. Suppose that a ∧ b = a ∧ c and a ∨ b = a ∨ c. Then b = c.

Proof. By using the absorption laws and the distributive laws, we obtain

b = b ∧ (a ∨ b) = b ∧ (a ∨ c) = (b ∧ a) ∨ (b ∧ c)
= (c ∧ a) ∨ (b ∧ c) = c ∧ (a ∨ b) = c ∧ (a ∨ c) = c,

as claimed.

18
When playing around with sets in P(M ), we can, besides forming intersections
and unions of two sets, also take the complement of a single set. The complement
of a set A in P(M ) is given by A′ = M \ A, and we have A ∩ A′ = ∅ and
A ∪ A′ = M (cf. Theorem 3.1(vi)). This inspire us to formulate the following
definition.

Definition 3.3. Let L be a bounded lattice and a an element of L. An element


b ∈ L is called a complement of a, if a ∧ b = 0L and a ∨ b = 1L . If every
element in L has a complement, then L is said to be a complemented lattice.

Example 3.6. By the discussion that foregoes the above definition, P(M ) is
a complemented lattice. ♦

Example 3.7. In every bounded lattice L, the universal lower bound 0L is a


complement of the universal upper bound 1L , and vice versa. This is due to
0L ∧ 1L = 0L and 0L ∨ 1L = 1L . ♦

Example 3.8. Consider the bounded lattice defined by the Hasse diagram in
Figure 2.4 on page 8 (where the universal upper bound is 36 and the universal
lower bound is 1). Here we have 2 ∧ 3 = 2 ∧ 9 = 1, so 3 and 9 might be
complements of 2. However, none of these elements will do, since neither 2 ∨ 3
nor 2 ∨ 9 equals 36. In this lattice, there are only four elements that have a
complement, namely 1, 4, 9, and 36. ♦

Example 3.9. A complement of an element is not necessarily uniquely deter-


mined, if it exists. An example which demonstrates this is the lattice L, defined
by the left Hasse diagram in Figure 3.1 on the preceding page. Here we have
a ∧ b = a ∧ c = 0L and a ∨ b = a ∨ c = 1L , which means that both b and c are
complements of a. ♦

With the previous example in mind, the natural question to ask is what (if
anything) a lattice has to fulfill, for each one of its elements to have a uniquely
determined complement (whenever such a complement exists). It turns out that
distributivity is a sufficient condition.

Theorem 3.4. In a bounded distributive lattice, each element has at most one
complement.

Proof. Let L be a bounded distributive lattice. Suppose a ∈ L has two com-


plements b and c. Then a ∧ b = a ∧ c = 0L and a ∨ b = a ∨ c = 1L . But in
a distributive lattice, the cancellation law holds (see Theorem 3.3). From this
b = c follows.

Exercises for Chapter 3


3.1. Verify that Hasse diagram to the right in Figure 3.1 on the previous page defines
a distributive lattice.

3.2. The Hasse diagram in Figure 3.2 on the following page defines a lattice L. Decide
whether or not this is distributive and/or complementary.

19
1L

0L
Figure 3.2

3.3. Show that (Z+ , gcd, lcm) is a distributive lattice.

3.4. Let L be a lattice. Show that the distributive laws (see Definition 3.2) are
equivalent.
(N.B. This is not a consequence of the Principle of Duality. Why not?)

3.5. Show that the formulae (3.1), (3.2), and (3.3) are valid in a distributive lattice.

3.6. In Figure 3.3 on page 20 we find a Hasse diagram for a bounded lattice L. Which
elements have a complement? Find these complements.
1L

a b

c d e

f g

0L
Figure 3.3

3.7. Show that every finite lattice is bounded.

3.8. Let (L, ∧, ∨) be a lattice. A subset I ⊆ L is called an ideal of L, if


• for all r, s ∈ I, we have r ∨ s ∈ I
• for all r ∈ I and s ∈ L, we have r ∧ s ∈ I.
Suppose now that L is a distributive lattice, and let a, b ∈ L.

(a) Show that


I(a, b) = {x ∈ L | a ∧ x = b ∧ x}
is an ideal of L.
(b) Let M = {1, 2, 3, 4}, A = {1, 2} and B = {2, 4}. Find the ideal I(A, B) of
the lattice P(M ).

20
Chapter 4

Boolean Algebras

We are now finally ready to introduce the class of lattices that are referred to
as Boolean algebras.

Definition 4.1. By a Boolean algebra we mean a lattice that is both dis-


tributive and complementary.

We note a couple of immediate consequences of the definition.

• Any Boolean algebra is a bounded lattice. This is because a Boolean


algebra is complementary, and by definition a complementary lattice has
to be bounded (see Definition 3.3),

• The cancellation law holds in a Boolean algebra, since a Boolean algebra


is a distributive lattice (see Theorem 3.3),

• Each element in a Boolean algebra has exactly one complement (due to


Theorem 3.4).

The notation (B, ∧, ∨, ′ , 0B , 1B ) will henceforth stand for a general Boolean


algebra.1 Here B is a set, ∧ and ∨ are the two binary operations (join and
meet) on a lattice, ′ denotes the mapping B ∋ a 7→ a′ ∈ B that maps a ∈ B
onto its uniquely determined complement a′ , while finally 0B and 1B denote the
universal lower bound and upper bound, respectively, of B.

Example 4.1. If we on P(M ) define ′ as the mapping A 7→ M \ A, then


(P(M ), ∩, ∪, ′ , ∅, M ) becomes a Boolean algebra. Later on in this chapter,
we will see that the power set in some sense acts as a prototype for Boolean
algebras. ♦

Above, we have defined a Boolean algebra as a certain kind of lattice, and


thereby as a certain kind of partially ordered set. But it also possible to define
a Boolean algebra as a purely algebraic structure. In Theorem 2.4 we did this
for a general lattice.
There are several different ways to define a Boolean algebra as an algebraic
structure, one of which we present in the following Theorem.
1 But for the sake of convenience, we will usually only write B.

21
Theorem 4.1. Let B be a set on which two binary operations ∧ and ∨ are
defined, such that
(i) ∧ and ∨ are commutative
(ii) ∧ and ∨ are associative
(iii) ∧ and ∨ are distributive over one another
(iv) B contains an identity element 1B with respect to ∧ and an identity ele-
ment 0B with respect to ∨
(v) for each a ∈ B there is an element a′ ∈ B such that a ∧ a′ = 0B and
a ∨ a′ = 1B .
Then the relation  on B, defined as

a  b ⇐⇒ a ∧ b = a for all a, b ∈ B,

is a partial ordering on B, and the poset (B, ) is a Boolean algebra.

Before we can prove this theorem, we need a lemma.

Lemma 4.1. With the same prerequisites as in Theorem 4.1, every a ∈ B is


idempotent with respect to both ∧ and ∨, i.e. a ∧ a = a ∨ a = a.

Proof. Let a ∈ B. By (v) in Theorem 4.1 there is an a′ ∈ B such that a∧a′ = 0B


and a ∨ a′ = 1B . Moreover (iv) tells us that 0B and 1B are identity elements
with respect to ∨ and ∧, respectively. Using this and the distributive laws (iii),
we obtain

a = a ∧ 1B = a ∧ (a ∨ a′ ) = (a ∧ a) ∨ (a ∧ a′ ) = (a ∧ a) ∨ 0B = a ∧ a.

The proof of a ∨ a = a is left as an exercise for the reader.


Proof of Theorem 4.1. We will first show that the absorption laws are fulfilled,
even though they are not explicitly mentioned in the statement of the theo-
rem. The reason for us wanting to show these laws, is that we want to apply
Theorem 2.4 later on in the proof.
We will only prove that a ∧ (a ∨ b) = a for all a, b ∈ B, while the reader is
asked to deal with the proof of the other absorption law. Since 0B is the identity
element of ∨, we have

a ∧ (a ∨ b) = (a ∨ 0B ) ∧ (a ∨ b) = a ∨ (0B ∧ b). (4.1)

In the last equality above, we have used the fact that we know that ∨ is dis-
tributive over ∧. We now choose b′ ∈ B so that b ∧ b′ = 0B (which is possible
according to (v) in the statement of the theorem). Then

0B ∧ b = (b ∧ b′ ) ∧ b = (b′ ∧ b) ∧ b = b′ ∧ (b ∧ b) = b′ ∧ b = 0B , (4.2)

where we among other things have used that b is idempotent with respect to ∧
(Lemma 4.1). By combining (4.1) and (4.2), and again using the fact that 0B
is the identity element of ∨, we obtain

a ∧ (a ∨ b) = a ∨ (0B ∧ b) = a ∨ 0B = a,

22
and the absorption law is thereby proved.
Thus we know for the time of being, that ∧ and ∨ are commutative and as-
sociative (due to (i) and (ii)), and that they fulfill the absorption laws. Thereby
Theorem 2.4 applies, and we can conclude that

a  b ⇐⇒ a ∧ b = a for all a, b ∈ B

defines a partial ordering on B, such that for each a, b ∈ B both sup(a, b) and
inf(a, b) exists and are equal to a ∨ b and a ∧ b, respectively. In other words, B
is a lattice. This lattice B is distributive, since we demand (iii) to be true,
It remains to show that B is bounded and complementary. Theorem 3.2
and (iv) together implies that 1B is a universal upper bound and that 0B is a
universal lower bound of B, whence B is bounded. The requirement (v) finally
gives that B is complementary.

In the next theorem we will state an algebraic identity for Boolean algebras,
that we recognize from elementary set theory.

Theorem 4.2 (De Morgan’s Laws). Let a and b be any elements of a Boolean
algebra B. Then

(a ∧ b)′ = a′ ∨ b′ and (a ∨ b)′ = a′ ∧ b′ .

Proof. We noted in the beginning of this chapter that a complement of an


element in a Boolean algebra is uniquely determined. Therefore, if we can show
that a′ ∨ b′ is a complement of a ∧ b, the left one of the two laws above will
follow. Now

(a ∧ b) ∧ (a′ ∨ b′ ) = (a ∧ b ∧ a′ ) ∨ (a ∧ b ∧ b′ )
= (b ∧ 0B ) ∨ (a ∧ 0B )
= 0B ∨ 0B
= 0B ,

and similarly (a ∧ b) ∨ (a′ ∨ b′ ) = 1B . This proves one of the two laws. The
second one now follows from the Principle of Duality.

Corollary. Let a1 , a2 , . . . , an ∈ B, where B is a Boolean algebra. Then


n
!′ n n
!′ n
^ _ _ ^

ai = ai och ai = a′i .
i=1 i=1 i=1 i=1

Proof. Exercise 4.7.

By a finite Boolean algebra we of course mean a Boolean algebra that con-


tains a finite number of elements. We will shortly prove that there are essentially
only kind of finite Boolean algebra, namely the power set of some finite set. In
other words: Given a finite Boolean algebra B, there is a finite set M , such
that B is isomorphic to (P(M ), ∩, ∪, ′ , ∅, M ). But to be able to prove this we
must first define the equivalence for Boolean algebras, of group isomorphisms
and ring isomorphisms.

23
Definition 4.2. A mapping φ : B → C, where B and C are Boolean algebras,
is called a (Boolean) homomorphism, if

φ(a ∧ b) = φ(a) ∧ φ(b)

φ(a ∨ b) = φ(a) ∨ φ(b) (4.3)
 ′ ′
φ(a ) = φ(a)

for all a, b ∈ B. If φ in addition is bijective, we call φ a (Boolean) isomorphism.


If there is a Boolean isomorphism φ : B → C, then B and C are said to be
isomorphic, and we write B ≃ C.

Example 4.2. The left Hasse diagram in Figure 4.1 on page 24 illustrates the
Boolean algebra P(M ), where M = {1, 2}, and the right Hasse diagram of the
same figure illustrates the Boolean algebra (B, gcd, lcm), where B = {1, 3, 5, 15}.
As we can see, the Hasse diagrams are similar, which indicates that these two
Boolean algebras are isomorphic. It is also easy to verify that φ : P(M ) → L,
defined as φ(∅) = 1, φ({1}) = 3, φ({2}) = 5, and φ(M ) = 15, is an isomorphism.
For instance

φ({1} ∪ {2}) = φ(M ) = 15 = lcm(3, 5) = lcm[φ({1}), φ({2})]

and similarly φ({1} ∩ {2}) = gcd[φ({1}), φ({2})]. Furthermore {2}, the comple-
ment of {1} in P(M ), is mapped onto 5, which is the complement of 3 = φ({1})
in B.
By the way, there is one more possible Boolean isomorphism from P(M )
to B. Which one? ♦
M 15

{1} {2} 3 5

∅ 1
Figure 4.1

Theorem 4.3. Let φ : B → C be a Boolean homomorphism.


(i) If a, b ∈ B fulfills a  b in B, then φ(a)  φ(b) in C.
(ii) φ(0B ) = 0B and φ(1B ) = 1C .

Proof. (i) Since a  b is equivalent to a ∧ b = a (or, if you wish, a ∨ b = b), the


relation φ(a)  φ(b) follows from φ(a) = φ(a ∧ b) = φ(a) ∧ φ(b).
(ii) We have φ(0B ) = φ(a ∧ a′ ) = φ(a) ∧ φ(a′ ) = φ(a) ∧ φ(a)′ = 0C , and in a
similar way φ(1B ) = 1C .
Just as the case are for groups and rings, the Boolean isomorphism relation ≃
is an equivalence relation on any set of Boolean algebras, which we leave as an
exercise for the reader to prove.
We are now almost ready to prove that any finite Boolean algebra B is
isomorphic to P(M ) for some finite set M . But we have not yet any idea on

24
how a Boolean isomorphism φ : B → P(M ) would look like. The get such an
idea, we will take a closer look at P(M ).
A property that any set A ∈ P(M ) has, is that it is uniquely determined of
its elements that is contained in the set. This means that any non-empty finite
set {x1 , x2 , . . . , xn } can be written as a union
n
[
{x1 , x2 , . . . , xn } = {x1 } ∪ {x2 } ∪ · · · ∪ {xn } = {xi }
i=1

of uniquely determined singletons {xi }. Notice that for each i, there is no set
A ∈ P(M ) such that ∅ ⊂ A ⊂ {xi }.
Could we establish something similar for a general finite Boolean algebra B?
Yes, in fact we can! In Theorem 4.4, a few pages ahead, we will show that any
a 6= 0B in B has a unique representation of the form
n
_
a = a1 ∨ a2 ∨ · · · ∨ an = ai ,
i=1

for some n ∈ Z+ , where each ai 6= 0B is so ‘small’, that no element b ∈ B can be


found, such that 0B ≺ b ≺ ai . By a ‘unique representation’ we here mean that
the relative positions of the elements ai have no importance. This is of course
due to ∨ being commutative.

Definition 4.3. Let B be a Boolean algebra (not necessarily a finite one), and
let a ∈ B. We say that a is an atom of B, if a 6= 0B , and if there is no b ∈ B
such that 0B ≺ b ≺ a.

Example 4.3. The atoms of P(M ) are exactly the singletons. So for example,
if M = {1, 2, 3} then the atoms of P(M ) are {1}, {2}, and {3}. ♦

We will now approach Theorem 4.4 via a few lemmas. Before each lemma, we
will illustrate it in the special case of the Boolean algebra P(M ).
The first lemma is a generalization of the following fact about P(M ): Every
non-empty set A ∈ P(M ) has a singleton {x} ∈ M as one of its subsets.

Lemma 4.2. Let B be a finite Boolean algebra. Then for each b 6= 0B in B


there is an atom a ∈ B such that a  b.

Proof. Let A1 = {a ∈ B | 0B ≺ a ≺ b}. If A1 happens to be empty, then b itself


is an atom and we are done. Otherwise, we let a1 be any element of A1 . We then
have 0B ≺ a1 ≺ b. If the set A2 = {a ∈ B | 0B ≺ a ≺ a1 } is empty, then a1 is an
atom that fulfills a1  b, and we are done. If not, we pick any a2 ∈ A2 (we will
then have 0B ≺ a2 ≺ a1 ≺ b), and consider the set A3 = {a ∈ B | 0B ≺ a ≺ a2 }.
If A3 = ∅ then a2 is an atom such that a2  b, but if A3 is non-empty we repeat
the above process. Now we cannot repeat this procedure forever, since B is
finite. Hence there are finitely many elements a1 , a2 , . . . , an ∈ B such that

0B ≺ an ≺ an−1 ≺ · · · ≺ a1 ≺ b,

where an is an atom. Thereby the lemma is proved.

25
Let {x} be a singleton, i.e. an atom in the Boolean algebra P(M ). If A is
any subset of M , then {x} ∩ A could either be equal to {x} (which happens
if x ∈ A), or to ∅ (which happens if x ∈
/ A). This can be generalized to any
Boolean algebra.

Lemma 4.3. Let a be an atom of the Boolean algebra B. Then for all b ∈ B
we either have a ∧ b = 0B or a ∧ b = a.

Proof. Since a ∧ b = inf(a, b) is the greatest lower bound of a and b, we have


a ∧ b  a for all b ∈ B. But since a is an atom, we can only have a ∧ b = 0B or
a ∧ b = a, as claimed.
Looking at P(M ) again, we see that if we pick any two different elements x
and y of M , then the corresponding singletons {x} and {y} must be disjoint, i.e.
{x} ∩ {y} = ∅. In a general Boolean algebra, we have the following situation:

Lemma 4.4. If a1 and a2 are two different atoms of a Boolean algebra B, then
a1 ∧ a2 = 0B .

Proof. Suppose, on the contrary, that a1 ∧ a2 6= 0B . Then since both a1 and a2


are atoms, Lemma 4.3 implies that a1 ∧ a2 = a1 and a1 ∧ a2 = a2 . But a1 6= a2 ,
so we have a contradiction.
Let A be any subset of a set M . Suppose that A ∩ {x} = ∅ for all x ∈ M , i.e.
that the intersection of A and any singleton is empty. Then A of course is the
empty set. The next lemma, which is the final lemma we need before we can
state and prove Theorem 4.4, generalizes this fact to a general Boolean algebra.

Lemma 4.5. Let b be an element of a finite Boolean algebra B. Suppose


a ∧ b = 0B for all atoms a of B. Then b = 0B .

Proof. Suppose b 6= 0B . Then by Lemma 4.2, there is an atom a such that


a  b, which is equivalent to a ∧ b = a (Theorem 2.2). But by the premises
of the lemma we are proving, a ∧ b = 0B whenever a is an atom. But since
a ∧ b = a, we then must have a = 0B . This is a contradiction, since an atom is
by definition different from 0B . The assumption b 6= 0B must therefore be false;
whence b = 0B .

Theorem 4.4. Let B be a finite Boolean algebra and b 6= 0B an element of B.


Then there are uniquely determined atoms a1 , a2 , . . . , an ∈ B such that
n
_
b= ai . (4.4)
i=1

Proof. Since b 6= 0B , we have a ∧ b 6= 0B for at least one atom a, because of


Lemma 4.5. This means that the set

A = {a ∈ B | a is an atom, a ∧ b 6= 0B } = {a1 , a2 , . . . , an }

is non-empty. Here a1 , a2 , . . . , an are exactly those atoms a of B that fulfills


a ∧ bW6= 0B . If a is not one of these atoms, then a ∧ b = 0B . We claim that
n
b = i=1 ai is the representation of b that we want to prove exists and is unique.

26
Wn
But c = i=1 ai . Then we wish to show that b = c. This wish will be
fulfilled, if we first can show that b ∧ c = c and b ∧ c′ = 0B holds, since we then
will obtain

b = b ∧ 1B = b ∧ (c ∨ c′ ) = (b ∧ c) ∨ (b ∧ c′ ) = c ∨ 0B = c.

Let us therefore in turn show that

b ∧ c = c and b ∧ c′ = 0B . (4.5)

We know that for all ai ∈ A have that ai ∧ b 6= 0B . Since these ai all are atoms,
Lemma 4.3 implies that all ai have the property b ∧ ai = ai . Hence
n
! n n
_ _ _
b∧c=b∧ ai = (b ∧ ai ) = ai = c,
i=1 i=1 i=1

and so the left formula of (4.5) is proved.


Let a be any atom of B. Then either a ∈ A or a ∈ / A. If a ∈ A, then a = aj
for some j ∈ {1, 2, . . . , n}. Since the succession of the ai does not matter, we will
not loose in generality, if we assume that j = 1. By the Corollary of De Morgan’s
Laws (see Theorem 4.2),
n
!
^
′ ′
a ∧ (b ∧ c ) = a1 ∧ b ∧ ai
i=1
n
! (4.6)
^
= (a1 ∧ a′1 ) ∧b∧ a′i = 0B ,
i=2

where the last equality is due to a1 ∧ a′1 = 0B . On the other hand, if a ∈


/ A,
then a ∧ b = 0B and therefore

a ∧ (b ∧ c′ ) = 0B . (4.7)

Putting (4.6) and (4.7) together, we obtain a ∧ (b ∧ c′ ) = 0B for all atoms a ∈ B.


By Lemma 4.5 we therefore must have

b ∧ c′ = 0B .

Thus the right formula of (4.5) is also verified.


We have now shown that a representation of b in the form (4.4) exists. It
remains to show that it also is unique. In other words, we want to show if we
also have
m
_
b= bi ,
i=1

where b1 , b2 , . . . , bm are atoms of B, then {b1 , b2 , . . . , bm } = {a1 , a2 , . . . , an }.


By Lemma 4.4, br ∧ bs = 0B if r 6= s, yielding
m
! m
_ _
bj ∧ b = bj ∧ bi = (bj ∧ bi ) = bj
i=1 i=1

27
for all j = 1, 2, . . . , m. But we also have
n
! n
_ _
bj ∧ b = bj ∧ ai = (bj ∧ ai ),
i=1 i=1

and so by Lemma 4.4 we must have bj ∈ A = {a1 , a2 , . . . , an }, since otherwise


bj ∧ ai = 0B for all i, which would lead to the contradiction bj = 0B . Hence we
have
{b1 , b2 , . . . , bm } ⊆ {a1 , a2 , . . . , an },
and by a similar argument the reversed inclusion is also true. The proof are
thereby finished.

Theorem 4.5. Let B be a finite Boolean algebra. Then B ≃ P(M ) for some
finite set M .

Proof. Let M be the set of all atoms of B. Then M of course is finite. If


a ∈ B, then according to Theorem 4.4 there is a uniquely determined Wn subset
{a1 , a2 , . . . , an } of M (i.e. an element of P(M )) such that a = i=1 ai , if a 6= 0B .
We make the advanced(?) guess that the mapping φ : B → P(M ) defined as
( Wn
φ(a) = φ ( i=1 ai ) = {a1 , a2 , . . . , an }, if a 6= 0B
φ(0B ) = ∅

is a Boolean isomorphism.
To begin with, it is easy to see Wnthat φ is bijective: If {a1 , a2 , . . . , an } are
any non-empty set of atoms, then i=1 ai is mapped onto {a1 , a2 , . .W . , an } by φ,
m
implying
Wn that φ is surjective. Moreover, if φ(a) = φ(b), where a = i=1 ai and
b = j=1 bj , then the sets {a1 , a2 , . . . , am } are {b1 , b2 , . . . , bn } equal. Hence they
contain exactly the same elements, and therefore a and b are the same element
in B, by the uniqueness of a representation of an element in a Boolean algebra,
as a finite join of atoms, see Theorem 4.4. Thus φ is injective as well.
It remains to prove that φ(a ∧ b) = φ(a) ∩ φ(b), φ(a ∨ b) = φ(a) ∪ φ(b), and
φ(a′ ) = M \ φ(a) = ∁φ(a). We leave the proof of φ(a ∨ b) = φ(a) ∪ φ(b) to the
reader (see Exercise 4.12). To show that φ(a ∧ b) = φ(a) ∩ φ(b), we first note
that !  n 
_m _ m _
_ n
a∧b= ai ∧  bj  = (ai ∧ bj ). (4.8)
i=1 j=1 i=1 j=1

By Lemma 4.4 we know that ai ∧ bj = 0B , whenever ai 6= bj . Thus those i and j


for which ai 6= bj , do not give any contribution to the right-hand side of (4.8).
Only those aiW∧ bj where ai = bj contributes to the right-hand side. This means
r
that a ∧ b = i=1 ci , where the elements ci are given by

{c1 , c2 , . . . , cr } = {a1 , a2 , . . . , am } ∩ {b1 , b2 , . . . , bn }.

This implies that

φ(a ∧ b) = {c1 , c2 , . . . , cr }
= {a1 , a2 , . . . , am } ∩ {b1 , b2 , . . . , bn }
= φ(a) ∩ φ(b).

28
We now turn to the proof of φ(a′ ) = ∁φ(a). From WnTheorem 4.4 we know
Wm that a
and a′ both have unique representations a = i=1 ai and a′ = j=1 cj , re-
spectively, where all ai and cj are atoms of B. Let A = {a1 , a2 , . . . , an } and
C = {c1 , c2 , . . . , cm }. Then φ(a) = A and φ(a′ ) = C, and we want to show that
C = ∁A, i.e. that A ∪ C = M and A ∩ C = ∅.
Suppose A ∪ C 6= M . Then there is an atom d of B such that d ∈ / A and
d∈/ C. Hence d is different from all the atoms ai and cj . Then by Lemma 4.4,
d ∧ ai = d ∧ cj = 0B for all i, j, which implies
n
! n n
_ _ _
d∧a=d∧ ai = (d ∧ ai ) = 0B = 0B ,
i=1 i=1 i=1

and by the same reasoning d ∧ a′ = 0B . From this it follows that

d = d ∧ 1B = d ∧ (a ∨ a′ ) = (d ∧ a) ∨ (d ∧ a′ ) = 0B ∨ 0B = 0B .

But being an atom, d 6= 0B . Hence A ∪ C = M .


Suppose A ∩ C 6= ∅. Then we can find i and j such that the atoms ai
and cj are equal. We can assume that i = j = 1, whence a1 = c1 = b, say. By
Lemma 4.4, b ∧ ai = b ∧ cj = 0B for all i = 2, 3, . . . , n and j = 2, 3, . . . , m, and
therefore !
n
_ _n
b∧a=b∧ ai = (b ∧ ai ) = b ∧ a1 = a1 ,
i=1 i=1

and similarly b ∧ a = c1 . But a1 = c1 , and therefore


a1 = a1 ∧ a1 = a1 ∧ c1 = (b ∧ a) ∧ (b ∧ a′ ) = (b ∧ b) ∧ (a ∧ a′ ) = b ∧ 0B = 0B ;

a contradiction. Thus we must have A ∩ C = ∅, which completes the proof of


the theorem.

Corollary. If B is a finite Boolean algebra, then |B| = 2n for some n ∈ N.2

Proof. By Theorem 4.4 we know that |B| = |P(M )| for some finite set M . As-
sume that M contains n elements, and suppose we want to construct a subset A
of M . For each one of the n elements of M we have two choices; either we
include the element in A or we do not. This means that there are 2n possible
ways to construct a member A of P(M ), whence |P(M )| = 2n .

Example 4.4. In Example 3.8 we concluded that the lattice L of Example 2.14
is not complementary. For this reason, L cannot be a Boolean algebra. We can
make the same conclusion by referring to the Corollary of Theorem 4.4 above:
If L is a Boolean algebra, then |L| must be a power of 2. This is not the case;
we have |L| = 9. ♦

Exercises for Chapter 4


4.1. Does any of the two Hasse diagrams in Figure 3.1 on page 18 define a Boolean
algebra?
2 Here we use the convention that N = {0, 1, 2, . . . }.

29
4.2. Show that (a′ )′ = a for all elements a in a Boolean algebra.

4.3. Let a and b be elements in a Boolean algebra B. Show that

a ∧ b′ = 0R ⇐⇒ a ∧ b = a

and
a ∨ b′ = 1R ⇐⇒ a ∨ b = a.

4.4. Let a, b ∈ B, where B is a Boolean algebra. Show that a  b, if and only if


a′  b ′ .

4.5. Given the premises of Theorem 4.1, show that


1. a ∨ a = a for all a ∈ B
2. a ∨ (a ∧ b) = a for all a, b ∈ B.

4.6. Let B be a Boolean algebra. Show that

a ∨ (a′ ∧ b) = a ∨ b and a ∧ (a′ ∨ b) = a ∧ b

for all a, b ∈ B.

4.7. Prove the Corollary of Theorem 4.2.

4.8. Suppose M = {1, 2, 3, 4} and that B is the set of all positive divisors of 330.
Let φ be a mapping from P(M ) to B, such that

φ({1}) = 11, φ({2}) = 2, φ({2, 3}) = 6, and φ({1, 4}) = 55.

If φ is supposed to be a Boolean isomorphism, then onto what element in B


must the following elements of P(M ) be mapped by φ?
(a) {1, 2}
(b) {4}
(c) {1, 3, 4}

4.9. Let B and C be Boolean algebras and φ : B → C a homomorphism. Show that


n
! n n
! n
^ ^ _ _
φ ai = φ(ai ) and φ ai = φ(ai )
i=1 i=1 i=1 i=1

for all a1 , a2 , . . . , an ∈ B.

4.10. Show that Boolean isomorphism ≃ is an equivalence relation on each set of


Boolean algebras.

4.11. Let φ : B → C be a Boolean isomorphism. Show that if a is an atom of B, then


φ(a) is an atom of C.

4.12. Complete the proof of Theorem 4.5, by showing that φ(a ∨ b) = φ(a) ∪ φ(b).

4.13. Let B = {1, 2, 3, 5, 6, 10, 15, 30} be the set of all positive integers in 30. Show
that (B, gcd, lcm) is a Boolean algebra. Derive a formula for how to compute
the complement a′ of an element a ∈ B.

4.14. Let n be a positive integer and B the set of all positive divisors of n. Show that
(B, gcd, lcm) is a Boolean algebra, if and only if n has no divisor in the form p2 ,
where p is a prime.

30
4.15. Let n and B be as in the previous exercise. What are the atoms of B?

4.16. If a and b are elements of a Boolean algebra B, such that (a ∨ b′ ) ∧ (a′ ∨ b) = 1B ,


show that a = b.

4.17. Let a, b, and c be elements of a Boolean algebra. Show att

(a ∧ b) ∨ (a′ ∧ c) ∨ (b ∧ c) = (a ∧ b) ∨ (a′ ∧ c).

4.18. Let (B, ∧, ∨) be a Boolean algebra, where |B| ≥ 2. Define two new binary
operations + and · on B, according to

a + b = (a ∧ b′ ) ∨ (a′ ∧ b) and a · b = a ∧ b,

respectively. Show that (B, +, ·) is a commutative ring with unity, in which


a2 = a holds for every a ∈ B. This is a so-called Boolean ring.
Conversely, let (R, +, ·) be a Boolean ring. Define two new binary operations ∧
and ∨ on R, according to

a ∧ b = ab and a ∨ b = a + b + ab,

respectively. Show that (R, ∧, ∨) is a Boolean algebra.

31
Chapter 5

Boolean Algebras and


Electronic Circuits

We will close this text by a concise description of how the theory of Boolean
algebras can be applied, when it comes to constructing electronic circuits. In
this context one works with a certain Boolean algebra, which we now will define.
We start from the set B = {0, 1} and define the binary operations ∧ and ∨
on B, according to the Cayley tables in Figure 5.1 on page 32. Moreover, taking
complements of the elements in B are defined by the rules 0′ = 1 and 1′ = 0.
It is easy to see that (B, ∧, ∨, ′ , 0, 1) actually is a Boolean algebra. For each

∧ 0 1 ∨ 0 1
0 0 0 0 0 1
1 0 1 1 1 1

Figure 5.1

positive integer n we let B n denote the cartesian product of n copies of B, i.e.

B n = {(b1 , b2 , . . . , bn ) | bi ∈ {0, 1}, i = 1, 2, . . . , n}.

By a Boolean mapping of n variables we will mean a mapping f : B n → B.


We will sometimes write f = f (x1 , x2 , . . . , xn ) when we refer to such a mapping.
Here each xi is a variable that may attain any of the elements of B, i.e. 0 or 1.
This means among other things that

xi ∧ xi = xi ∨ xi = xi

for all i = 1, 2, . . . , n. Such a variable will be called a Boolean variable.


Two Boolean mappings f and g of n variables are said to be equal, if the
usual definition of equality of mappings is fulfilled. In this case, it means that
we should have
f (b1 , b2 , . . . , bn ) = g(b1 , b2 , . . . , bn )
for all (b1 , b2 , . . . , bn ) ∈ B n . The set of all Boolean mappings of n variables will
be denoted Pn . The Boolean mapping that is constantly equal to 0, will be

32
denoted O, while we will let I denote the Boolean mapping that is constantly
equal to 1. Thus we have

O(b1 , b2 , . . . , bn ) = 0 and I(b1 , b2 , . . . , bn ) = 1 (5.1)

for all (b1 , b2 , . . . , bn ) ∈ B n .

Example 5.1. Consider the two Boolean mappings f (x, y) = (x ∧ y) ∨ (x′ ∧ y ′ )


and g(x, y) = (x ∨ y ′ ) ∧ (x′ ∨ y). To compute f (a, b) or g(a, b), where (a, b) ∈ B 2 ,
we may use the Cayley tables in Figure 5.1 on the previous page, along with
the rules 1′ = 0 and 0′ = 1. For example we have

g(1, 1) = (1 ∨ 1′ ) ∧ (1′ ∨ 1) = (1 ∨ 0) ∧ (0 ∨ 1) = 1 ∧ 1 = 1.

The table in Figure 5.2 on page 33 shows that f = g as Boolean mappings. ♦

x y f (x, y) g(x, y)
0 0 1 1
0 1 0 0
1 0 0 0
1 1 1 1

Figure 5.2

Definition 5.1. Given two Boolean mappings f, g ∈ Pn , we define the meet


f ∧ g ∈ Pn of f and g, as the mapping that fulfills

(f ∧ g)(b1 , b2 , . . . , bn ) = f (b1 , b2 , . . . , bn ) ∧ g(b1 , b2 , . . . , bn )

for all (b1 , b2 , . . . , bn ) ∈ B n . In the same spirit we define the join f ∨ g ∈ Pn


of f and g, and the complement f ′ ∈ Pn of f , by

(f ∨ g)(b1 , b2 , . . . , bn ) = f (b1 , b2 , . . . , bn ) ∨ g(b1 , b2 , . . . , bn )

and
f ′ (b1 , b2 , . . . , bn ) = f (b1 , b2 , . . . , bn )′ ,
respectively, for all (b1 , b2 , . . . , bn ).

Example 5.2. If f (x, y) = x ∨ y ′ and g(x, y) = x ∧ y ′ , then the meet f ∧ g and


the join f ∨ g of f and g are

(f ∧ g)(x, y) = (x ∨ y ′ ) ∧ (x ∧ y ′ )

and
(f ∨ g)(x, y) = (x ∨ y ′ ) ∨ (x ∧ y ′ ),
respectively. The complements of f and g are f ′ (x, y) = (x ∨ y ′ )′ = x′ ∧ y and
g ′ (x, y) = (x ∧ y ′ )′ = x′ ∨ y, respectively. Here we have applied De Morgan’s
Laws (Theorem 4.2). ♦

Now that we have defined join, meet, and complement on Pn , the following
theorem would not come as a surprise.

33
Theorem 5.1. The set Pn , provided with the operations ∧, ∨, and ′ from
Definition 5.1, is a Boolean algebra. The universal upper and lower bounds in
this Boolean algebra are the Boolean mappings I and O, defined by (5.1).

Proof. We leave as an exercise for the reader to show that (i)-(v) of Theorem 4.1
are fulfilled. This will prove the theorem.
The Boolean algebra Pn can be used to describe in which way some simple
electronic circuits are constructed. We picture ourselves that a number of relays
occurs here and there in such a circuit. These relays allow us to control the way
in which the electric current flows throughout the circuit. A relay can be in two
different states. It can be switched on, so that the current may pass through
the relay, or it can be switched off, in which case the current is unable to pass
through the relay. A relay is assumed to be in exactly one of these states. In
other words, it cannot attend both states simultaneously, and there are no other
states “in between”.
Assume that a given circuit contains finitely many relays r1 , r2 , . . . , rn . As
we have seen, any relay can be switched off or on, in similar way that a Boolean
variable may attain any of the values 0 or 1. It lies near at hand to let a
Boolean variable xi describe the state of the relay ri . We agree to let xi = 1
if the relay ri is switched on, and to let xi = 0 if ri is switched off. The
ability (or inability) for the circuit to have current flowing through it, from one
side to the other, depends on which states its relays are in. Since each relay
is described by a Boolean variable, we will in a natural way obtain a Boolean
mapping f = f (x1 , x2 , . . . , xn ) that describes the behavior of the whole circuit.
If it turns out that f (b1 , b2 , . . . , bn ) = 1 for some b1 , b2 , . . . , bn ∈ B, then it is
possible for the current to flow through the circuit, if its relays r1 , r2 , . . . , rn are
in the states that corresponds to the assignments x1 = b1 , x2 = b2 , . . . , xn = bn
of the Boolean variables. On the other hand, if f (b1 , b2 , . . . , bn ) = 0, the opposite
situation is of course at hand, i.e. the circuit does not let the current through.
In what follows, we will use the same notation on a relay, as we do for its
corresponding Boolean variable.

Example 5.3. There are two fundamental ways to connect relays. Both these
ways are pictured in Figure 5.3 on page 34. To the left in this figure, there is a
series connection of two relays x1 and x2 . The current may pass through a
series connection of two relays, if and only if both relays are switched on. This
means that if f = f (x1 , x2 ) is the Boolean mapping that describes the behavior
of the circuit, then f (x1 , x2 ) = 1, if and only if x1 = x2 = 1. Hence the circuit
can be represented by the Boolean mapping f (x1 , x2 ) = x1 ∧ x2 .
To the right in Figure 5.3 on page 34 we have a parallel coupling of the
relays x1 and x2 . In this case, the current can not flow through the circuit,
if and only if both x1 and x2 are switched off. Therefore the circuit can be
represented by the Boolean mapping g(x1 , x2 ) = x1 ∨ x2 . ♦
x1
x1 x2
x2
Figure 5.3

34
If x is a relay, then x′ will signify a relay that is switched on, if and only if x is
switched off, or in other words,

x′ = 1 ⇐⇒ x = 0.

Regarding x and x′ as Boolean variables, we have x′ ∧ x = 0 and x′ ∨ x = 1,


which means that current cannot pass through a series connection of the relays x
and x′ , but that it always will manage to this, in a parallel coupling of these
relays.
In a circuit, we could also have a relay y that behaves exactly in the same
way as another relay x, i.e. that y = 1, if and only if x = 1. We will then use
the same notation for both these relays.
Two circuits are said to be equivalent, if their corresponding Boolean map-
pings are equal. It would be advantageous (to decrease the cost, for instance) if
one could replace a given circuit by another, equivalent circuit that has fewer re-
lays. This can be accomplished if you translate the given circuit into a Boolean
mapping, simplify the expression of this mapping (by using the laws of arith-
metics in Boolean algebra), and finally interpret this simplified expression as a
circuit.

Example 5.4. The two circuits of Figure 5.4 on page 35 are equivalent, since
they are represented by f (x1 , x2 ) = (x1 ∧ x2 ) ∨ (x′1 ∧ x′2 ) and g(x1 , x2 ) = (x1 ∨
x′2 ) ∧ (x′1 ∨ x2 ), respectively, and as we saw in Example 5.1, these two Boolean
mappings are equal. ♦
x1 x2 x1 x′1

x′1 x′2 x′2 x2


Figure 5.4

Example 5.5. The left circuit of Figure 5.5 on the following page can be rep-
resented by the Boolean mapping

f (x1 , x2 , x3 ) = (x1 ∧ x2 ) ∨ [(x′1 ∨ x′2 ) ∧ x2 ] ∨ x3 .

But since

f (x1 , x2 , x3 ) = (x1 ∧ x2 ) ∨ [(x1 ∧ x2 )′ ∧ x2 ] ∨ x3



= [(x1 ∧ x2 ) ∨ (x1 ∧ x2 )′ ] ∧ [(x1 ∧ x2 ) ∨ x2 ] ∨ x3
= (1 ∧ x2 ) ∨ x3
= x2 ∨ x3 ,

this circuit is equivalent to a circuit, consisting of a single parallel coupling of


the relays x2 and x3 ; the right circuit of the same figure. ♦

Exercises for Chapter 5


5.1. Show that (B, ∧, ∨), where B = {0, 1}, and where ∧ and ∨ is defined by the
Cayley tables in Figure 5.1 on page 32, is a Boolean algebra.

35
x1 x2

x′1 x2
x2
x′2 x3
x3

Figure 5.5

5.2. Prove Theorem 5.1.


5.3. Consider the Boolean mapping f (x1 , x2 , x3 ) = (x1 ∨ x2 ) ∧ (x′1 ∨ x3 ). Compute
(a) f (1, 0, 1)
(b) f (0, 0, 1)
(c) f (1, 1, 1) ∧ f (0, 0, 0)
(d) f (1, 1, 0) ∨ f (0, 0, 1)
(e) f ′ (0, 1, 0)
(f) f ′ (0, 1, 1) ∨ f (0, 1, 1)
5.4. Find |Pn |, i.e. the number of Boolean mappings in n variables.
5.5. Find a Boolean mapping f that represents the circuit in Figure 5.6 on page 36.
x1 x4 x8

x3 x6

x2 x5 x7
Figure 5.6

5.6. Draw a circuit for each one of the Boolean mappings


f (x1 , x2 , x3 ) = (x1 ∧ x2 ) ∨ x3 and g(x1 , x2 , x3 ) = x1 ∧ (x2 ∨ x3 ).

5.7. Show that the circuit in Figure 5.7 on page 36 always lets current through, no
matter what state the relays x1 and x2 are in.
x1 x2

x′1 x′2
x1 x′2

x′1 x2
Figure 5.7

5.8. Consider the two Boolean mappings


f (x1 , x2 , x3 ) = (x1 ∧ x3 ) ∨ (x′1 ∧ x2 )

and

g(x1 , x2 , x3 ) = (x1 ∨ x2 ) ∧ (x′1 ∨ x3 ) ∧ (x2 ∨ x3 ).


Show that f = g.

36
5.9. Consider a circuit that can be represented by any of the Boolean mappings f
or g from the previous exercise. (Since f = g, such a circuit will be unique up
to equivalence.) Find all states of x1 , x2 , x3 such that current can pass through
the circuit.

5.10. Which ones (if any) of the circuits in Figure 5.8 on page 37 are equivalent?
x2 x1 x2 x3
x1
x3 x2 x3 x1
(a) (b)
x1 x2
x1 x2
x2 x3

x3 x1 x3 x1
(c) (d)

Figure 5.8

5.11. Construct a circuit consisting of three relays only, that is equivalent to the circuit
in Figure 5.9 on page 37.
x3 x2 x′1 x2 x′2 x′1

x′2 x3 x2 x3 x2 x′1
Figure 5.9

37
Answers

2.1. (e) and (f)

2.2. Yes!

2.4. (a) Totally ordered


(b) Not totally ordered
(c) Not totally ordered (unless M = ∅)

2.5. Yes, all maximal ideals are maximal elements; the trivial ideal is the only mini-
mal element.

2.7. (a) Do not exist


(b) j
(c) Do not exist
(d) f
(e) h
(f) Do not exist

2.8. 20 18

14 10 6 9

2 3

2.9. (a) 7, 8, 9, 10, 11, and 12 are maximal elements; 1 is a minimal element
(b) There are no upper bounds; 1 is a lower bound
(c) sup Y does not exist; inf Y = 1.
(d) The divisibility relation

2.10. a ∧ b = min(a, b), a ∨ b = max(a, b).

2.11. (a) a ∧ b = c ∨ d
(b) a ∧ (b ∨ c) = (a ∧ b) ∨ c
(c) a ∨ b  c ∧ d
(d) (a ∧ b) ∨ a  a ∧ (b ∨ c)

38
2.12. ∧ a b c d e f ∨ a b c d e f
a a b c d e f a a a a a a a
b b b c d e f b a b b b b b
c c c c f f f c a b c b b c
d d d f d f f d a b b d b d
e e e f f e f e a b b b e e
f f f f f f f f a b c d e f

2.13. d

a b

2.16. If X has an universal upper (lower) bound, this will be the infimum (supremum)
of ∅.
3.2. The lattice is complementary but not distributive.

3.6. 1L has the complement 0L ; b has the complements c, d, and f ; c has the com-
plement b; d has the complement b; f has the complement b; 0L has the com-
plement 1L .

3.8. (b) I(A, B) = {∅, {2}, {3}, {2, 3}}


4.1. Yes, the right one.

4.8. (a) 22
(b) 5
(c) 165

4.13. a′ = 30/a

4.15. The prime factors of n


5.3. (a) 1
(b) 0
(c) 0
(d) 0
(e) 0
(f) 1
n
5.4. 22

5.5. f (x1 , x2 , . . . , x8 ) = (x1 ∧ x4 ∧ x8 ) ∨ (x3 ∧ x6 ) ∨ (x2 ∧ x5 ∧ x7 )

5.6. f : x1 x2 g: x2
x1
x3 x3

39
5.9. (x1 , x2 , x3 ) ∈ {(0, 1, 0), (0, 1, 1), (1, 0, 1), (1, 1, 1)}

5.10. Yes: (a) and (d) are equivalent, as well as (b) and (c).

5.11. A series connection of the relays x′1 , x2 , and x3 .

40

You might also like