UNALGALG
UNALGALG
S. L. Wismath, K. Denecke
v
vi
The second approach to the study of abstract algebras involves the study of
terms and identities. Here we classify algebras according to the identities or
axioms they satisfy. Given any set of identities, of a particular type, we can
form the class of all algebras of that type which satisfy the given identities.
Such a class is called an equational class. This point of view is quite different
from the classical structure theoretical approach, and combines logic, model
theory and abstract algebra.
In the first six chapters of this book we develop these two approaches to
general algebra. We begin in Chapter 1 with a detailed list of examples of
various algebras, arising from a number of areas of Mathematics and Com-
puter Science. The subalgebra, homomorphic image, product and quotient
constructions are described in Chapters 1, 3 and 4. In Chapter 5 we intro-
duce terms and identities, to be used in the second approach. This treatment
culminates in Chapter 6 with a complete proof of Birkhoff’s Theorem, which
relates our two approaches: this theorem says that any variety is an equa-
tional class, and vice versa, so that the two approaches lead us to exactly
the same classes of algebras.
These first six chapters provide a solid foundation in the core material of
universal algebra. With a leisurely pace, careful exposition, and lots of ex-
amples and exercises, this section is designed as an introduction suitable for
beginning graduate students or researchers from other areas.
problem involving one kind of object by using the theory of the second kind
of objects.
The study of clones is another important feature of this work. Clones (closed
sets of functions) occur in universal algebra as clones of term operations or
of polynomial operations of an algebra, in automata theory as combinations
of elementary switching circuits, and in logic as sets of truth-value functions
generated by elementary truth-value functions.
The last three chapters, Chapters 13, 14 and 15, tie together the main themes
of Galois-connections, clones and varieties, and algebraic machines. In Chap-
ter 13 we return to the study of complete lattices of closed sets, obtained
from Galois-connections. We describe several methods for obtaining com-
plete sublattices of such complete lattices, and in Chapter 14 illustrate these
methods on our two main examples, the lattice of all varieties of a given
type and the lattice of all clones of operations on a fixed set. This leads to
complete sublattices of M-solid varieties (using M-hyperidentities to obtain
new closure operators) and G-clones and H-clones. We then return to appli-
cations to theoretical Computer Science, by looking at hypersubstitutions as
tree-recognizers. Hypersubstitutions involve replacing not only the leaves of
a tree by elements but also the nodes by term operations. The main result
here is the proof of the equivalence of this “parallel” replacement with the
linear approach. Hypersubstitutions can also be applied to the tree trans-
formations and tree transducers of Chapter 8, and to the syntactical and
semantical hyperunification problems.
The material of this book is based on lectures given by the authors at the
University of Potsdam (Germany), Chiangmai University and KhonKaen
University (Thailand) and the University of Balgoevgrad (Bulgaria), and in
research seminars with our students. The authors are grateful for the critical
input of a number of students.
The work of Shelly Wismath on this book was supported by a research leave
from the University of Lethbridge (January to July 2001) and by funding
from the Natural Sciences and Engineering Research Council of Canada. The
hospitality of the Institute of Mathematics of the University of Potsdam, and
particularly the General Algebra Research Group, during the year July 2000
to June 2001 is also gratefully acknowledged. Special thanks to Stephen and
Alice for accompanying me to Germany for a year.
Contents
Introduction v
1 Basic Concepts 1
1.1 Algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Subalgebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.4 Congruence Relations and Quotients . . . . . . . . . . . . . . 21
1.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
ix
x
5.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
Bibliography 363
Index 373
Chapter 1
Basic Concepts
In this first chapter we shall introduce the two basic concepts of Universal
Algebra, operations and algebras, and provide a number of examples of al-
gebras. One of these examples, lattices, is useful in a theoretical way as well:
any algebra of any type is accompanied by some lattices, making the theory
of lattices important in the study of all algebras. In Section 3 we examine
the concept of a subalgebra, and the generation of subalgebras. In Section 4
we look at congruence relations and quotient algebras. Congruence relations
on algebras generalize the well-known notion of the congruence modulo n
defined on the ring of all integers, and quotient algebras generalize the con-
struction of quotient or residue rings modulo n.
1.1 Algebras
As a branch of Mathematics, Algebra is the study of algebraic structures, or
sets of objects with operations defined on them. These sets and operations
often arise in other fields of Mathematics and in applications. Let us consider
the following example:
1
2 CHAPTER 1. BASIC CONCEPTS
4 · 3
·
· · · · · · ·
g1
O
·
·
1 2
g2
If we label the corners of the rectangle by 1, 2, 3 and 4, then the four map-
pings of symmetry can be described by the following permutations, in the
usual cyclic notation:
e by (1), s1 by (14)(23),
s2 by (12)(34), z0 by (13)(24).
defined on our set D(Φ). Thus we consider the pair (D(Φ); ◦), consisting of
the set of mappings with this binary operation. This is what is called an
algebra: a base set of objects, together with a set of one or more operations
which are defined on this base set. In this example the set of operations
contains only one element, the binary operation of composition of mappings.
1.1. ALGEBRAS 3
Notice that we may use the same example to produce a different algebra, as
follows. We can use as our base set the set {1, 2, 3, 4} of the four corners of
the rectangle, and as our set of operations the set
which is completely different from the first one but which can also be used
to describe the geometric situation.
2. Definition 1.1.2 can be extended in the following way to the special case
that n = 0, for a nullary operation. We define A0 := {∅}. A nullary operation
is defined as a function f : {∅} → A. This means that a nullary operation
on A is uniquely determined by the element f (∅) ∈ A. For every element
a ∈ A there is exactly one mapping fa : {∅} → A with fa (∅) = a. Therefore
a nullary operation may be thought of as selecting an element from the set
A. If A = ∅ then there are no nullary operations on A.
3. If A is the two element set {0, 1}, operations on A are called Boolean
operations. If we associate the truth-values false with 0 and true with 1, then
the Boolean operations are just the truth-value functions of the Classical
4 CHAPTER 1. BASIC CONCEPTS
¬ ∧ 0 1 ∨ 0 1 ⇒ 0 1 ⇔ 0 1
0 1 0 0 0 0 0 1 0 1 1 0 1 0 ,
1 0 1 0 1 1 1 1 1 0 1 1 0 1
respectively.
Using the concept of an operation we can now define an algebra. As we have
seen, an algebra should be a pair consisting of a base set of objects and a
set of operations defined on the base set. The set of operations is usually
indexed by some index set I, so we will define an indexed algebra here;
the more general setting of non-indexed algebras will be considered later, in
Chapter 10.
1.2 Examples
In this section we illustrate our basic definition of an algebra, by presenting
a number of examples, many of which we will also use later. We note that
an operation f A on a set A will often be denoted simply by f , omitting the
superscript. Also, operations of arities zero, one and two are often said to be
nullary, unary and binary respectively.
1.2. EXAMPLES 5
Example 1.2.1 As we saw in the previous section, we may describe the set
of symmetry mappings of a rectangle by two different algebras:
Example 1.2.3 An algebra (G; ·) of type τ = (2), with one binary oper-
ation, is called a groupoid. Here the single binary operation f is denoted by
· or even just by juxtaposition, so we write x · y or just xy instead of f (x, y).
then the lattice is said to be distributive. We remark that (V3) and (V30 )
follow from (V4) and (V40 ) respectively.
(V6) ∀x , y , z ∈ V (z ≤ x ⇒ x ∧ (y ∨ z) = (x ∧ y) ∨ z).
Lattices are important both as examples of a kind of algebra, and also in the
study of all other kinds of algebras too, since any algebra turns out to have
some lattices associated with it. We shall study such associated lattices in
the coming sections. For now, we show how lattices are connected to partially
ordered sets. A binary relation ≤ on a set A is called a partial order on A
if it is reflexive, anti-symmetric and transitive. There is a close connection
between lattices and partially ordered sets, in the sense that each determines
the other, as the following theorem shows.
Theorem 1.2.10 Let (L; ≤) be a partially ordered set in which for all x, y ∈
L both the infimum {x, y} and the supremum {x, y} exist. Then the binary
V W
Proof: Let (L; ≤) be a partially ordered set, in which for any two elements
x, y ∈ L the infimum {x, y} and the supremum {x, y} exist. If we put
V W
8 CHAPTER 1. BASIC CONCEPTS
x∧y = {x, y} and x ∨ y = {x, y}, then the required identities (V1) -
V W
x ≤ y : ⇔ x ∧ y = x,
and show that this gives a partial order relation on L. Reflexivity follows
from (V30 ), x ∧ x = x ; for antisymmetry we see that
(x ≤ y) ∧ (y ≤ x) ⇔ (x ∧ y = x) ∧ (y ∧ x = y),
x = x ∧ y = y ∧ x = y;
Example 1.2.11 A lattice (or partially ordered set) L in which for all sets
B ⊆ L the infimum B and the supremum B exist is called a complete
V W
These axioms tell us that in the partial order determined by the lattice, 0
acts as the least element and 1 as the greatest element.
(B1) ∀x ∈ B (x ∧ ¬x = 0) and
(B2) ∀x ∈ B (x ∨ ¬x = 1) (complement laws).
Example 1.2.14 For applications in Logic and the theory of switching cir-
cuits, algebras whose carrier sets are sets of operations on a base set play an
important role. We recall from Definition 1.1.2 that for a non-empty base set
A, we denote by O(A) the set of all finitary operations on A (with nullary
operations regarded as constant unary operations).
We can make the set O(A) into an algebra in several ways, by defining vari-
ous operations on O(A). We first define the following composition operation
on O(A). If f A ∈ On (A) and g1A , . . ., gnA ∈ Om (A), we obtain a new opera-
tion f A (g1A , . . . , gnA ) in Om (A), by setting f A (g1A , . . . , gnA )(a1 , . . . , am ) to be
f A (g1A (a1 , . . . , am ), . . . , gnA (a1 , . . . , am )), for all a1 , . . . , am ∈ A. This gives an
operation
On (A) × (Om (A))n → Om (A),
which is called composition or superposition. Clearly, the whole set O(A) is
closed under arbitrary compositions.
The set O(A) also contains certain special elements called the projections.
For each n ≥ 1 and each 1 ≤ j ≤ n, the n-ary function enj defined on A
by enj (a1 , . . . , an ) = aj is called the j-th projection mapping of arity n. Any
subset of O(A) which is closed under composition and contains all these
projection mappings is called a clone on A. Of course O(A) itself has these
properties, and it is called the full clone on A.
There is another way to define an algebraic structure on the set O(A), using
the following operations ∗, ξ, τ and ∆ on O(A):
with
(f ∗ g)(x1 , . . . , xm , xm+1 , . . . , xm+n−1 ) : =
f (g(x1 , . . . , xm ), xm+1 , . . . , xm+n−1 ),
for all x1 , . . . , xm+n−1 ∈ A;
with
τ (f )(x1 , . . . , xn ) := f (x2 , x1 , x3 , . . . , xn );
with
∆(f )(x1 , . . . , xn−1 ) := f (x1 , x1 , x2 , . . . , xn−1 )
for every n > 1 and for all x1 , . . . , xn ∈ A; and
ξ(f ) = τ (f ) = ∆(f ) = f, if n = 1.
We also use the nullary operation e21 which picks out the binary projection
on the first coordinate; as defined above, e21 (a1 , a2 ) = a1 , for all a1 , a2 ∈ A.
With these five operations, we obtain an algebra (O(A); ∗, ξ, τ, ∆, e21 ) of
type (2, 1, 1, 1, 0). This algebra is called the full iterative algebra on A, or
sometimes also a clone on A. We shall study clones in more detail in Chapter
10.
(M4) 1(x) = x.
Thus modules and vector spaces are algebras with infinitely many opera-
tions. There is another way to describe modules and vector spaces alge-
braically, which avoids this infinitary type, and which is more often used.
But it involves a modification of our definition of an algebra. We have so far
considered only what are called homogeneous or one-based algebras, where
we have only one base set of objects and all operations are defined on this
set. One can also define what are called heterogeneous or multi-sorted or
multi-based algebras, where there are two or more base sets of objects and
operations are allowed on more than one kind of object. We can use this
approach here, by allowing the set V of all vectors and the universe of the
field R as two different base sets in such a heterogeneous algebra. Then we
have two operations, + : V × V → V for the usual vector addition and
· : R × V → V for the scalar multiplication of a scalar from R times a
vector from V . Then (V, R; , +, ·) is an example of a heterogeneous algebra.
Although we will focus on homogeneous or one-based algebras in this book,
we will consider the multi-based algebras described in Example 1.2.16 below
in more detail in Chapter 8, and we remark that the algebraic theory for
multi-based algebras can be developed in a completely analogous way.
input x ∈ X, the value δ(z, x) is the state which results when the input x
is read in when the machine is in the state z. The element λ(z, x) is the
output element which is produced by input x if the automaton is in state z.
If all the sets Z, X and B are finite, then the automaton is said to be finite;
otherwise it is infinite. When δ and λ are functions, and so have exactly one
image for each state-input pair (z, x), the automaton is called deterministic;
in the non-deterministic case we allow δ and λ to take on more than one
value, or to be undefined, for a given input pair. Sometimes a certain state
z0 ∈ Z is selected as an initial state, and in this case we write (Z, X; δ, z0 ) or
(Z, X, B; δ, λ, z0 ) for the automaton. In the finite case, an automaton may
be fully described by means of tables for δ and λ, as shown below.
δ x1 ... xn λ x1 ... xn
z1 δ(z1 , x1 ) . . . δ(z1 , xn ) z1 λ(z1 , x1 ) . . . λ(z1 , xn )
· · · · · ·
· · · · · ·
· · · · · ·
zk δ(zk , x1 ) . . . δ(zk , xn ) zk λ(zk , x1 ) . . . λ(zk , xn )
z1 x2 ; b1 z2
s x2 ; b1 s
x1 ; b2@ @
I
x1 ; b2
@
@ @
@ x3 ; b2 x2 ; b1
x3 ; b2@@ x3 ; b2
@
@ @
@ @
R s
z3
@
x1 ; b1
δ x1 x2 x3 λ x1 x2 x3
z1 z1 z1 z3 z1 b2 b1 b2
z2 z2 z1 z3 z2 b2 b1 b2
z3 z3 z2 z1 z3 b1 b1 b2
1.3 Subalgebras
The set of all mappings of symmetry of an equilateral triangle Φ,
3·
·
g3 · O· · g2
··
· · ·
· ·
1 · · · 2
g1
Here (12), (13) and (23) correspond to reflections at the lines g1 , g3 , and
g2 , respectively; and (1), (123) and (132) correspond to rotations around
the point O by 0, 120 and 240 degrees, respectively. We notice that D(Φ)
forms a group with respect to the composition of these mappings; that is,
with respect to the multiplication of the corresponding permutations.
Since the composition of two rotations around the point O is again a rotation
around O, our operation preserves the subset of D(Φ) consisting of all rota-
tions around O, and we say that this subset is closed under our operation.
This closure can also be seen if we consider the multiplication table of the
corresponding permutations:
(ii) A ⊆ B;
(i) (A ⊆ B) ∧ (B ⊆ C) ⇒ A ⊆ C;
(ii) (A ⊆ B ⊆ C) ∧ (A ⊆ C) ∧ (B ⊆ C) ⇒ A ⊆ B.
Proof: (i) Clearly, for all i ∈ I we have fiA ⊆ fiB ⊆ fiC for the graphs, and
thus fiA ⊆ fiC .
Now suppose that B = (B; (fiB )i∈I ) is an algebra of type τ and that
{Aj | j ∈ J} is a family of subalgebras of B. We define a set A : =
T
Aj
j∈J
and operations fiA : = fiB | A for all i ∈ I. We will verify that when
the intersection set A is non-empty, the algebra A = (A; (fiA )i∈I ) is indeed
an algebra of type τ , to be called the intersection of the algebras Aj , and
denoted by A : = Aj .
T
J
Proof: We will show that A is closed under all fiB , for all i ∈ I. Consider an
element (a1 , . . . , ani ) ∈ Ani , so that (a1 , . . . , ani ) ∈ Anj i for all j ∈ J. There-
fore we have fiB (a1 , . . . , ani ) ∈ Aj for all j ∈ J and i ∈ I, since Aj ⊆ B for all
j ∈ J. Hence fiB (a1 , . . . , ani ) also belongs to the intersection A =
T
Aj ,
j∈J
and A ⊆ B.
16 CHAPTER 1. BASIC CONCEPTS
hXiB : = ∩ {A | A ⊆ B and X ⊆ A, }
Notice that hXiB is the least (with respect to subset inclusion) subalgebra
of B to contain the set X. In particular, X might generate the whole algebra
B, if hXiB = B.
where the elements of the base set are the equivalence classes of the integers
modulo 6 and +, − denote addition and subtraction of equivalence classes.
This algebra is a group, with [0]6 as an identity element. For each one-element
subset X of Z6 we calculate the subalgebra generated by X:
Proof: These properties follow immediately from the properties of the op-
erator h i; we leave the details to the reader.
Then we inductively define E 0 (X) := X, and E k+1 (X) := E(E k (X)), for all
k ∈ N. With this notation we can describe our subalgebra.
Theorem 1.3.8 For any algebra B of type τ and for any non-empty subset
∞
X ⊆ B, we have hXiB = E k (X).
S
k=0
∞
Proof: We first give a proof by induction on k that hXiB ⊇ E k (X).
S
k=0
For k = 0 it is clear that E 0 (X) = X ⊆ hXiB . For the inductive step,
assume that the proposition hXiB ⊇ E k (X) is true. Let a ∈ E k+1 (X); we
may assume that a 6∈ E k (X). Then there exists an element i ∈ I and ele-
ments a1 , . . . , ani ∈ E k (X) with a = fiB (a1 , . . . , ani ). Since E k (X) ⊆ hXiB
and hXiB is the carrier set of a subalgebra of B, we have a ∈ hXiB too.
Thus E k+1 (X) ⊆ hXiB , and therefore by induction on k we have hXiB ⊇
∞
E k (X).
S
k=0
∞ ∞
Now we show that hXiB ⊆ E k (X), by showing that E k (X) is the
S S
k=0 k=0
∞
carrier set of a subalgebra of B. Let i ∈ I and a1 , . . . , ani ∈ E k (X). Then
S
k=0
for every l ∈ {1, . . . , ni } there is a minimal k(l) ∈ N with al ∈ E k(l) (X).
Take m to be the maximum of these k(l), for 1 ≤ l ≤ ni . Then al ∈ E m (X)
for all l = 1, . . . , ni and thus
∞
[
fiB (a1 , . . . , ani ) ∈ E m+1 (X) ⊆ E k (X).
k=0
∞
E k (X) is closed under fiB for i ∈ I. By Lemma 1.3.3
S
This shows that
k=0
∞
E k (X) is the carrier set of a subalgebra of B which contains X. Since by
S
k=0
18 CHAPTER 1. BASIC CONCEPTS
Now we can consider the structural properties of the set of all subalgebras
of an algebra B of type τ . We denote this set by Sub(B). We know by 1.3.5
that the intersection of two subalgebras of B, if it is non-empty, is again a
subalgebra of B, and we would like to use this to define a binary operation
on Sub(B). That is, we want to define a binary operation ∧ on Sub(B) by
∧ : (A1 , A2 ) 7→ A1 ∧ A2 : = A1 ∩ A2 .
However, this operation is not defined on all of Sub(B), because it does not
deal with the case where A1 ∩ A2 is the empty set. But we can fix this, either
by allowing the empty set to be considered as an algebra (see the remark at
the end of Section 1.1), or by adjoining some new element to the set Sub(B)
and defining A1 ∧ A2 to be this new element, whenever A1 ∩ A2 = φ.
1.3. SUBALGEBRAS 19
∨ : (A1 , A2 ) 7→ A1 ∨ A2 : = hA1 ∪ A2 iB .
Our goal in defining these two operations was of course to make Sub(B) into
a lattice, and this turns out to be the case.
We have been considering the problem of finding, for a given algebra B and a
given subset X of B, the subalgebra of B generated by X. We can also look at
the subalgebra generation question from a different point of view, as follows.
Given an algebra B, can we find a (proper) subset X of B which generates
B? In particular, we are often interested in whether we can find a finite such
generating set X. If we can, then B is said to be finitely generated, with X
as a finite generating system; otherwise B is said to be infinitely generated.
Of course any finite algebra has a finite generating system (the carrier
set itself), so this question is only interesting for infinite algebras. Before
addressing the question of when a finite generating system is possible, we
consider two examples.
Example 1.3.11 Consider the algebra (N; ·) of type (2), where · is the usual
multiplication on the set N of all natural numbers. Is this finitely generated?
Since an arbitrary generating system would have to contain the infinite set
of all prime numbers, the answer in this case is no. But the infinite set of
all prime numbers is a generating system of this algebra, since every natural
number has a representation as a product of powers of prime numbers.
20 CHAPTER 1. BASIC CONCEPTS
Example 1.3.12 Recall from Example 1.2.14 in Section 1.2 that for any
base set A, we can define an algebra of type (2, 1, 1, 1, 0) on the set O(A) of
all finitary operations on A. This is the algebra (O(A); ∗, ξ, τ, ∆, e21 ). It is well
known in Logic that for the special case when A is the two-element set {0, 1},
this algebra is finitely generated. The two-element set X = {∧, ¬} (conjunc-
tion and negation) acts as a generating set for the algebra, since all Boolean
functions can be generated from these two using our five operations. In the
general case, it is an important problem to decide whether a given set X is
a generating system of a subalgebra C of the algebra (O(A); ∗, ξ, τ, ∆, e21 ).
Example 1.3.15 Let A be the two-element set {0, 1}, and let
A ⊆ ··· ⊆ C ⊆ B
(see for instance Gluschkow, Zeitlin and Justschenko, [49]). Using this fact
we have the following general completeness criterion for finitely generated
algebras.
For instance, let Z be the set of all integers, let m be a natural number, and
let Z/(m) be the set of all equivalence classes modulo m. Then the kernel
of the function ϕ : Z → Z/(m) which is defined by a 7→ [a]m , that is,
which maps every integer to its class modulo m, is obviously the well-known
equivalence or congruence modulo m :
a ≡ b (m) : ⇔ ∃g ∈ Z (a − b = gm).
(This is also often written as a ≡ b mod m). With respect to addition
and multiplication in the ring (Z; +, ·, −, 0) of all integers, this congruence
modulo m has the following additional property:
a1 ≡ b1 (m) ∧ a2 ≡ b2 (m) ⇒ a1 + a2 ≡ b1 + b2 (m) ∧ a1 · a2 ≡ b1 · b2 (m).
This tells us that in this example the kernel of the function ϕ is compatible
with the operations of the ring. This observation motivates the following
more general definition.
(a1 , b1 ) ∈ θ, . . . , (an , bn ) ∈ θ
implies (f (a1 , . . . , an ), f (b1 , . . . , bn )) ∈ θ.
∆A : = {(a, a) | a ∈ A} and ∇A = A × A
Example 1.4.4 1. Let (Z; +, ·, −, 0) be the ring of all integers and let
m ∈ Z be an integer with m ≥ 0. Then the congruence modulo m defined
on Z is a congruence relation on (Z; +, ·, −, 0). Note that for m = 0 or m
= 1 we get the trivial relations on Z.
a b c d
f (x) b a d c .
{a} ∪ {b} ∪ {c, d}, {a, b} ∪ {c} ∪ {d}, {a, b} ∪ {c, d},
where fiA is ni -ary and a1 , . . ., ai−1 , ai+1 , . . ., ani are fixed elements of A.
We want now to define lattice operations on the set ConA of all congruence
relations of an algebra A, much as we did on the set Sub(A) for subalgebras.
As was the case there, our basic tool is the fact that the intersection of
congruences is again a congruence.
Remark 1.4.7 Theorem 1.4.6 is also satisfied for arbitrary families of con-
gruence relations on A. But in general, the union of two congruence relations
of an algebra A is not a congruence relation, since this does not hold even
for equivalence relations, as the following example shows. Take
A = {1, 2, 3}, θ1 = {(1, 1), (2, 2), (3, 3), (1, 2), (2, 1)},
θ2 = {(1, 1), (2, 2), (3, 3), (2, 3), (3, 2)}.
The relations θ1 and θ2 are equivalence relations, but
θ1 ∪ θ2 = {(1, 1), (2, 2), (3, 3), (1, 2), (2, 1), (2, 3), (3, 2)}
But although the union of two congruence relations θ1 and θ2 need not be
a congruence relation, as in the subalgebra case we can use intersections of
congruences to define a smallest congruence generated by the union. This
motivates the following definition.
It is easy to see that hθiConA has the three important properties of a closure
operator:
θ ⊆ hθiConA (extensivity),
θ1 ⊆ θ2 ⇒ hθ1 iConA ⊆ hθ2 iConA (monotonicity),
hhθiConA iConA = hθiConA (idempotency).
Remark 1.4.9 As we did in Section 1.3 for subalgebras, we can ask how
to derive, from a binary relation θ defined on A, the congruence relation of
the algebra A generated by θ. First we have to enlarge the relation θ to its
26 CHAPTER 1. BASIC CONCEPTS
αT : = ∩ {β ⊆ A2 | β ⊇ α and β is transitive}.
defined by
A/θ
([a1 ]θ , . . . , [ani ]θ ) 7→ fi ([a1 ]θ , . . . , [ani ]θ ) := [fiA (a1 , . . . , ani )]θ .
Theorem 1.4.11 For every algebra A of type τ and every congruence rela-
tion θ ∈ ConA, the previous definition produces an algebra A/θ of type τ ,
which is called the quotient algebra or factor algebra of A by θ.
Proof: Let i ∈ I. We have to show that if [aj ]θ = [bj ]θ for all j = 1, . . . , ni ,
then [fiA (a1 , . . . , ani )]θ = [fiA (b1 , . . . , bni )]θ . But this is exactly what our
compatibility property of a congruence guarantees. If [aj ]θ = [bj ]θ , then
(aj , bj ) ∈ θ for all j = 1, . . . , ni , and since θ is a congruence on the algebra
A, it follows that
and thus
Example 1.4.12 Let m ∈ N, and let Z/(m) be the set of all congruence
classes modulo m. We define two binary operations on these classes, by
This makes (Z/(m); +, ·) an algebra of type (2, 2). In fact it is a ring, called
the ring of all congruence classes modulo m. If m is a prime number, this
ring is also a field.
1.5 Exercises
1.5.1. Determine all subalgebras of the algebra
where the four binary operations are defined as follows: max(x, y) and
min(x, y) are the maximum and the minimum, respectively, with respect
to the order relation defined by 0 ≤ 1 ≤ 2 ≤ 3; x + y(4) is addition modulo
4; and 3 − x(4) denotes subtraction modulo 4.
1.5.3. Is the algebra (N; +) of type τ = (2) finitely generated? If so, give a
finite generating system.
28 CHAPTER 1. BASIC CONCEPTS
1.5.5. Let (N; ) be an algebra of type (1), with a unary operation defined
by
0 : = 0, n = n − 1 for n > 0.
Prove that this algebra has no basis.
x a b c d
f (x) c b a d.
Find all congruence relations θ on A, and for each one determine the quo-
tient algebra A/θ.
1.5.9. Determine all subalgebras and all congruence relations of the algebra
A = ({(0), (x), (2)}; ×) of type τ = (2), with × defined by
1.5.10. Draw the corresponding directed graph for the finite deterministic
automaton ({0, 1}, {0, 1}, {0, 1}; δ, λ), given by
1.5. EXERCISES 29
δ 0 1 λ 0 1
0 0 1 and 0 0 0
1 1 0 1 1 1
• •
• •
• • •
•
• •
L1 L2
31
32 CHAPTER 2. GALOIS CONNECTIONS AND CLOSURES
Subsets of A of the form C(X) are called closed (with respect to the operator
C) and C(X) is said to be the closed set generated by X.
(i) A ∈ H, and
(ii) ∩ B ∈ H for every non-empty subset B ⊆ H.
The elements of a closure system H are called closures.
Example 2.1.4 1. By Corollary 1.3.5, for every algebra A the set Sub(A)
of all subalgebras of A is a closure system on A.
2. By Theorem 1.4.6 and Remark 1.4.7, for every algebra A the set ConA
of all congruences on A is a closure system on A2 .
3. For every set A, the set EqA of all equivalence relations defined on A is
a closure system on A2 .
HC : = {X ⊆ A | C(X) = X}.
(ii) The closed sets with respect to CH are exactly the closures of H.
CH (Y ) = ∩ {H ∈ H | H ⊇ Y } ⊇ CH (X) = ∩ {H ∈ H | H ⊇ X}.
This follows because all sets H which contain Y also contain X, and therefore
the second intersection contains more components than the first one, making
it “smaller.” For the idempotency of CH , we observe that CH (CH (X)) is the
intersection of all closures from H which contain CH (X). But CH (X) itself
is such a closure: it is the intersection of all elements of H which contain X,
and since H is a closure system this intersection is also in H. This means
that CH (CH (X)) = CH (X). Altogether, CH is a closure operator.
CH (X) = ∩ {H ∈ H | H ⊇ X} = X ⇔ X ∈ H.
(iv) Since
X ∈ HCH ⇔ CH (X) = X ⇔ X ∈ H,
we have HCH = H.
C(X) ⊆ C(H) = H.
Corollary 2.1.9 For every algebra A = (A; (fiA )i∈I ), the system Sub(A)
is an inductive closure system and h iA is an inductive closure operator.
Proof: Let G ⊆ Sub(A) be upward directed; we have to prove that ∪G
∈ Sub(A). Let fiA be an ni -ary operation and assume that b1 , . . . , bni ∈ ∪G.
Then there exist sets G1 , . . ., Gni ∈ G with b1 ∈ G1 , . . ., bni ∈ Gni . Since
G is upward directed, there is a set G0 ∈ G with b1 , . . . , bni ∈ G0 . But
G0 ∈ Sub(A) means that
Just as we did for Sub(A), we can define a meet and join operation on any
closure system. If H ⊆ P(A) is a closure system, then for arbitrary sets
B ⊆ H, we use
^ _
B : = ∩ B and B : = ∩ {H ∈ H | H ⊇ ∪B}.
In particular, for two-element sets B, this gives us binary meet and join op-
erations on H. Then every closure system is a lattice, under these operations,
and in fact a complete lattice (see Definition 1.2.11).
Example 2.1.10 Let A be any algebra. Since Sub(A) and ConA are closure
systems, the lattices (Sub(A); ⊆) and (ConA; ⊆) are complete lattices.
2.2. GALOIS CONNECTIONS 37
This shows that all subalgebra lattices are complete lattices. We can then ask
whether any complete lattice occurs as the subalgebra lattice of some algebra,
and if not whether there are some classes of complete lattices which do occur
in this way. To answer these questions, we need the following concept.
Theorem 2.1.12 For every inductive closure system H, the partially or-
dered set (H; ⊆) is an algebraic lattice.
Proof: In an inductive closure system, every closure is the supremum of
finitely generated elements. Therefore inductive closure systems lead to al-
gebraic lattices.
We mention here without proof the following important result; a proof may
be found for instance in G. Grätzer, [51] or P. Cohn, [10].
We start with a fixed base set A, and as one of our sets of objects the set
O(A) of all operations on A. In Definition 1.4.1, we considered an intercon-
nection between operations from O(A) and binary relations θ ⊆ A2 , namely
that an operation could be compatible with, or preserve, a binary relation.
38 CHAPTER 2. GALOIS CONNECTIONS AND CLOSURES
1. Let ρ be the unary relation {0}. Notice that unary relations on a set are
simply subsets of this set. Then P olA {0} is the set of all Boolean functions
which preserve {0}, so
2. Let
α = {(a, b, c, d) ∈ A4 | a + b = c + d},
where + is the addition modulo 2. An n-ary operation f on A is called linear,
if there are elements a1 , . . . , an , c ∈ {0, 1} such that
f (x1 , . . . , xn ) = a1 x1 + · · · + an xn + c
for all x1 , . . . , xn ∈ {0, 1}, where again + is the addition modulo 2. It can be
shown that a Boolean function f is linear iff it preserves α. Thus P olA α is
the set of all linear Boolean functions.
2.2. GALOIS CONNECTIONS 39
These examples show how, for any given relation ρ on A, we can look for
the set of all operations which preserve ρ. In the other direction, for a given
operation f ∈ O(A) one can look for the set of all relations from R(A) which
are preserved by f . Such relations are called invariants of f , and the set of
all such invariants is denoted by InvA f , that is,
We can also extend the maps P olA and InvA to sets of relations or functions,
respectively. If F ⊆ O(A) is a set of operations on A, we define InvA F to
be the set of all relations which are invariant for all f ∈ F , and similarly
for a set Q ⊆ R(A) of relations on A, we denote by P olA Q the set of all
operations which preserve every relation ρ ∈ Q.
These maps P olA and InvA , between sets of relations and sets of operations
on A, show the basic idea of a Galois-connection, between two sets of objects.
The following theorem of Pöschel and Kalužnin shows that in this example,
we have the two important properties that will be used shortly to define a
Galois-connection. These properties are that our maps between the two sets
of objects map larger sets to smaller images, and that mapping a set twice
returns us to a set which contains the starting set.
The concept of an invariant, or a set of all objects which are invariant under
certain changes, also plays an important role in other branches of Mathemat-
ics and Science. From Analytic Geometry we know for example that by Felix
Klein’s “Erlanger Programm” ([64]), different geometries may be regarded
as the theories of invariants of different transformation groups. Instead of
sets of the form P olA R, in this setting we have sets of transformations, more
40 CHAPTER 2. GALOIS CONNECTIONS AND CLOSURES
These examples of mappings between two sets, with the two basic properties
from Theorem 2.2.2, form a model for our definition of a Galois-connection.
such that for all X, X 0 ⊆ A and all Y, Y 0 ⊆ B the following conditions are
satisfied:
(i) στ σ = σ and τ στ = τ ;
(iii) The sets closed under τ σ are precisely the sets of the form τ (Y ), for
some Y ⊆ B; the sets closed under στ are precisely the sets of the form
σ(X), for some X ⊆ A.
since σ(X) and σ(X 0 ) are subsets of B; and in the analogous way we get
from Y ⊆ Y 0 the inclusion στ (Y ) ⊆ στ (Y 0 ). Applying σ to the equation τ στ
= τ from part (i) gives us the idempotency of στ , and similarly for τ σ.
A relation between the sets A and B is simply a subset of A×B. Any relation
R between A and B induces a Galois-connection, as follows. We can define
the mappings
σ : P(A) → P(B), τ : P(B) → P(A),
by
σ(X) : = {y ∈ B | ∀x ∈ X ((x, y) ∈ R)},
τ (Y ) : = {x ∈ A | ∀y ∈ Y ((x, y) ∈ R)}.
Then it is easy to verify that the pair (σ, τ ) forms a Galois-connection be-
tween A and B, called the Galois-connection induced by R. In our example
with P ol and Inv, consider the preservation relation R between O(A) and
R(A), defined by
The reader can verify that the Galois-connection induced by this R is in fact
the one we described.
42 CHAPTER 2. GALOIS CONNECTIONS AND CLOSURES
It is clear from this definition that such formal concepts consist of pairs of
closed sets, under the closure operators τ σ and στ .
2.3. CONCEPT ANALYSIS 43
Let us denote by B(G, M, I) the set of all concepts of the context (G, M, I).
We can define a partial order on this set, by
(X1 , Y1 ) ≤ (X2 , Y2 ) :⇔ X1 ⊆ X2 (⇔ Y1 ⊇ Y2 ).
When (X1 , Y1 ) ≤ (X2 , Y2 ), the pair (X1 , Y1 ) is called a subconcept of the pair
(X2 , Y2 ), and conversely (X2 , Y2 ) is called a superconcept of (X1 , Y1 ). It is
easy to verify that (B(G, M, I), ≤) is a complete lattice.
To illustrate how the concepts of this context may be identified, let us choose
an object, say the planet Jupiter. We find the set of all properties or at-
tributes this object has: large, far from the sun and has moons. Now we look
for the set of all objects which has exactly these properties: Jupiter and Sat-
urn. This gives us the concept ({J, S}, {l, f, y}). We can also start with a set
of objects instead of with a single object, and dually we can work from a set
of attributes. For instance, we obtain the concept ({J, S, U, N, P }, {f, y}),
which is a superconcept of the first one.
Me x x x
V x x x
E x x x
Ma x x x
J x x x
S x x x
U x x x
N x x x
P x x x
44 CHAPTER 2. GALOIS CONNECTIONS AND CLOSURES
({M e, V }, {k, c, nm}), ({E, M a}, {k, c, y}), ({J, S}, {l, c, y}),
({U, N }, {m, c, y}), ({P }, {k, c, y}), ({M e, C, E, M a}, {k, c}),
({J, S, U, N, P }, {c, y}), ({E, M a, J, S, U, N, P }, {y}).
Extending this list by taking intersections, we reach the Hasse diagram shown
below for the lattice B(G, M, I); ≤).
The main goal is to find, for a given context, the hierarchy of concepts. The
super- and sub-concept relations also give implications between concepts,
allowing this method to be used to search for new implications. For more
information and examples about the theory of concept analysis we refer the
reader to the book [47] by B. Ganter and R. Wille.
M e, V, E, M a, J, S, U, N, P
s
M e, V, E, M a, P s s E, M a, J, S, U, N, P
E, M a, P
s s s J, S, U, N, P
M e, V, E, M a
s s P s
M e, V E, M a J, S
s s U, N
s
∅
2.4 Exercises
2.4.1. Prove that for any relation R ⊆ A × B, the maps σ and τ defined by
2.4.2. Let R ⊆ A × B be a relation between the sets A and B and let (µ, ι)
be the Galois-connection between A and B induced by R. Prove that for any
families {Ti ⊆ A | i ∈ I} and {Si ⊆ B | i ∈ I}, the following equalities hold:
S T
a) µ( Ti ) = µ(Ti ).
i∈I
S i∈I
T
b) ι( Si ) = ι(Si ).
i∈I i∈I
Homomorphisms and
Isomorphisms
Consider the function which assigns to every real number a ∈ R its absolute
value | a | in the set R+ of non-negative real numbers. This function h : R →
R+ has the property that it is compatible with the multiplicative structure of
R, because | a · b | = | a | · | b |. We get the same result whether we multiply
the numbers first and then use our mapping, or permute these actions, since
calculation with the images proceeds in the same way as calculations with
the originals. A corresponding observation can be made about assigning to
a square matrix A its determinant | A |, or to a permutation s on the set
{1, . . . , n} its sign. In these cases the compatibility of the mapping with
the operation can be described by the equations
the group of all even permutations (the alternating group) of order three
and the cyclic group of order three, respectively. Both algebras have type
47
48 CHAPTER 3. HOMOMORPHISMS AND ISOMORPHISMS
τ = (2), and it is easy to verify that both are groups. From their Cayley
tables
defined by the mapping (1) 7→ [0]3 , (123) 7→ [1]3 and (132) 7→ [2]3 , is com-
patible with the structure. For instance, we have
The following example shows that this compatibility property occurs in very
“practical” cases. An industrial automaton is designed to construct more
complex parts of a machine from certain simpler parts. There are working
phases in which two parts are combined into a new part, other phases in
which three parts are combined into a new one, and so on. These correspond
to binary operations, ternary operations, etc. If the parts are combined in a
different order, represented by a mapping h, then the automaton will con-
tinue to work in the correct way when h has our compatibility property. Thus
the concepts of homomorphism and isomorphism may be said to model math-
ematically the “artificial intelligence” we expect from such an automaton.
h
(different order)
Definition 3.1.1 Let A = (A; (fiA )i∈I ) and B = (B; (fiB )i∈I ) be algebras
of the same type τ . Then a function h : A → B is called a homomorphism
h : A → B of A into B if for all i ∈ I we have
h(fiA (a1 , . . . , ani )) = fiB (h(a1 ), . . . , h(ani )),
for all a1 , . . . , ani ∈ A. In the special case that ni = 0, this equation means
that h(fiA (∅)) = fiB (∅). That is, the element designated by the nullary
operation fiA in A must be mapped to the corresponding element fiB in B.
Example 3.1.2 1. It is easy to prove that for every algebra A, the identical
mapping idA : A → A, defined by idA (x) = x for all x ∈ A, is an automor-
phism of A.
50 CHAPTER 3. HOMOMORPHISMS AND ISOMORPHISMS
Let fiB be an ni -ary operation on B, for i ∈ I, and let (b1 , . . . , bni ) ∈ h(A1 )ni .
Then for each 1 ≤ j ≤ ni , we have bj = h(aj ) for some aj in A. Then
fiB (b1 , . . . , bni ) = fiB (h(a1 ), . . . , h(ani )) = h(fiA (a1 , . . . , ani )),
and the latter is in h(A1 ) since fiA (a1 , . . . , ani ) ∈ A1 . Application of the
subalgebra criterion thus proves that the image B1 is a subalgebra of B.
3.1. THE HOMOMORPHISM THEOREM 51
(ii) Let (a1 , . . . , ani ) ∈ (h−1 (B 0 ))ni and let fiA be ni -ary. We have
Then
We show first that E(h(X)) = h(E(X)) for all X ⊆ A. The set E(h(X))
consists of all elements h(y) with y ∈ X, plus elements of the form
The set h(E(X)) also consists of the elements h(y) with y ∈ X plus the
elements h(fiA (y1 , . . . , yni )), which agree with fiB (h(y1 ), . . . , h(yni )).
fiA (a1 , . . . , ani ) = fiA (h(a1 ), . . . , h(ani )) = h(fiA (a1 , . . . , ani )),
we can choose a subset B of A and look for the set of all automorphisms on A
whose set of fixed points contains B. When B is exactly the set of fixed points
of this set of automorphisms, Lemma 3.1.5 tells us that B is a subalgebra of
A. To explore this connection further, we will need the following definitions.
Definition 3.1.6 Let A and A0 be algebras of the same type τ and let
B be a subalgebra of both A and A0 . An isomorphism h : A → A0 with
h(b) = b for all b ∈ B is called a relative isomorphism between A and
A0 with respect to B. When h is an automorphism on A, it is then called a
relative automorphism of A with respect to the subalgebra B.
We can consider such a relative isomorphism, h : A → A0 with respect
to a common subalgebra B, in two ways. From “above,” we say that the
restriction of h to B is the identity isomorphism on B, while from “below,”
we call h an extension of the identity isomorphism on B.
The careful reader will have noticed that these interconnections amount to
a Galois-connection between the sets A and AutrelB A. We have a basic re-
lation
Definition 3.1.8 Let A and B be algebras of the same type τ and let h :
A → B be a homomorphism. The following binary relation is called the
kernel of the homomorphism h :
We may alternately express this as ker h = h−1 ◦ h, where h−1 is the inverse
relation of h.
and thus
S 6
(=)
S
S
S
S
g S f
S
S
S
SS
w
C
56 CHAPTER 3. HOMOMORPHISMS AND ISOMORPHISMS
since h = f ◦ g is a homomorphism.
(ii) We are assuming that ker g ⊆ ker h, so that f exists. Now we assume
that f is injective, and take (a1 , a2 ) ∈ ker h. So h(a1 ) = h(a2 ), and using
f ◦ g = h gives (f ◦ g)(a1 ) = (f ◦ g)(a2 ) and f (g(a1 )) = f (g(a2 )). Now the
injectivity of f gives g(a1 ) = g(a2 ) and thus (a1 , a2 ) ∈ ker g. This shows
that ker h ⊆ ker g, and hence the two kernels are equal.
Conversely, suppose that ker g = ker h, and let us show that f is injective.
Let f (c1 ) = f (c2 ) for c1 , c2 ∈ C. Since g is surjective, we can represent cj
in the form g(aj ), for some aj ∈ A, for j = 1, 2. Now we have f (g(a1 )) =
f (g(a2 )), and hence h(a1 ) = h(a2 ). This put (a1 , a2 ) in ker h, and by our
assumption also in ker g. But this means c1 = g(a1 ) = g(a2 ) = c2 , as re-
quired.
S 6
(=)
S
S
S
S
nat(ker h) S f
S
S
S
SS
w
A/ker h
A x
H H
HH h HH
6 HH 6 HH
HH
j HH
j
H
B h(x)
6 6
(=)
A∗1 PP ∗ given x P
h1 P
6
PP
PP by 6
PP
PP
q
P P
q
P
h(x)
h(A1 )
(=) >
>
h1
A1 x
is an isomorphism.
Proof: It is clear from the definition that θ2 /θ1 is a relation and in fact
A/θ
an equivalence relation on A/θ1 . Let fi 1 , for i ∈ I, be a fundamental
operation of the algebra A/θ1 and let
([a1 ]θ1 , [b1 ]θ1 ) ∈ θ2 /θ1 , . . . , ([ani ]θ1 , [bni ]θ1 ) ∈ θ2 /θ1 .
and thus
(fiA (a1 , . . . , ani ), fiA (b1 , . . . , bni )) ∈ θ2 .
This makes
and hence
A/θ1 A/θ1
(fi ([a1 ]θ1 , . . . , [ani ]θ1 ), fi ([b1 ]θ1 , . . . , [bni ]θ1 )) ∈ θ2 /θ1 .
3.3. EXERCISES 61
because ker(nat(θ2 /θ1 )) = θ2 /θ1 . Therefore part (i) of the General Ho-
momorphism Theorem tells us that our surjective homomorphism is also
injective, and so ϕ is an isomorphism.
3.3 Exercises
3.3.1. Let A be a set, let Θ be an equivalence relation on A and let f : A → A
be a function. Prove that f is compatible with Θ iff there is a mapping
g : A → A with
3.3.3. Let G = ({0, 1, 2, 3}; +, 0), where + is the operation of addition mod-
ulo 4, and let A = ({e, a}; ·, e) with a = e · a = a · e, e · e = a · a = e, both
algebras of type (2, 0). Let h : G → A be the mapping defined by 0 7→ e,
1 7→ e, 2 7→ a, and 3 7→ a. Is h a homomorphism?
In the previous chapters, we have seen three ways to construct new alge-
bras from given algebras: by formation of subalgebras, quotient algebras,
and homomorphic images. In this chapter we examine another important
construction, the formation of product algebras. One useful feature of this
new construction involves the cardinalities of the algebras obtained. The for-
mation of subalgebras or of homomorphic images of a given algebra leads
to algebras with cardinality no larger than the cardinality of the given alge-
bra. The formation of products, however, can lead to algebras with bigger
cardinalities than those we started with. There are several ways to define a
product of given algebras; we shall examine two products, called the direct
product and the subdirect product.
Y
P : = Aj : = {(xj )j∈J | ∀j ∈ J (xj ∈ Aj )}
j∈J
63
64 CHAPTER 4. DIRECT AND SUBDIRECT PRODUCTS
It is easy to check that the projections of the direct product are in fact sur-
jective homomorphisms.
fj
A - Aj a - fj (a)
@ 6 S 6
@ S
@ S
f @ (=) pj given by S
@ S
@ S
@ S
@ w
S
RQ
@
Aj (fj (a))j∈J
j∈J
Example 4.1.3 Let us consider the direct product of the two permutation
groups S2 and A3 . Here
−1 −1
S2 = ({τ0 , τ1 }; ◦, , (1)) and A3 = ({τ0 , α1 , α2 }; ◦, , (1)),
4.1. DIRECT PRODUCTS 65
with
We now consider a direct product of two factors. In this case we have two
projection mappings, p1 and p2 , each of which has a kernel which is a congru-
ence relation on the product. We will show that these two kernels have some
special properties. We recall first the definition of the product (composition)
θ1 ◦ θ2 of two binary relations θ1 , θ2 on any set A:
Proof: (i) Since ker p1 and ker p2 are equivalence relations on A1 × A2 , the
relation ker p1 ∧ ker p2 is also an equivalence relation on A1 × A2 , with
∆A1 ×A2 ⊆ ker p1 ∧ ker p2 . Conversely, let (x, y) ∈ ker p1 ∧ ker p2 , with x =
(a1 , b1 ) and y = (a2 , b2 ). From (x, y) ∈ ker p1 we have
a1 = p1 ((a1 , b1 )) = p1 ((a2 , b2 )) = a2 .
b1 = p2 ((a1 , b1 )) = p2 ((a2 , b2 )) = b2 .
we always have
((a1 , b1 ), (a2 , b2 )) ∈ ker p2 ◦ ker p1 ,
giving
(A1 × A2 )2 ⊆ ker p2 ◦ ker p1 .
The converse inclusion is true by definition, so we have
from (ii) to show that ker p1 ∨ ker p2 = (A1 × A2 )2 . For this we need the
well-known fact that for any two equivalence relations θ1 and θ2 , the equation
θ 1 ∨ θ2 = θ1 ◦ θ2
(see for instance Th. Ihringer, [58]), so when ker p1 ◦ker p2 = ker p2 ◦ker p1
we obtain
Thus any direct product of two factors produces two congruences with the
three special properties of Lemma 4.1.4. Conversely, the next theorem shows
that if we have two congruences on an algebra with these properties, we can
use them to write the algebra as a direct product of two factors.
(i) θ1 ∧ θ2 = ∆A ;
(ii) θ1 ∨ θ2 = A2 ;
(iii) θ1 ◦ θ2 = θ2 ◦ θ1 .
ϕ : A → A/θ1 × A/θ2
given by:
ϕ(a) = ([a]θ1 , [a]θ2 ), a ∈ A.
Proof: The given mapping ϕ is defined using the two natural homomor-
phisms, and is the unique map determined by them, as in Remark 4.1.2.
This makes ϕ a homomorphism, and we will show that it is also a bijection.
First, ϕ is injective: if ϕ(a) = ϕ(b), then [a]θ1 = [b]θ1 and [a]θ2 = [b]θ2 , so it
68 CHAPTER 4. DIRECT AND SUBDIRECT PRODUCTS
follows that (a, b) ∈ θ1 ∧ θ2 and a = b by (i). To see that the map ϕ is also
surjective, let (a, b) be any pair in A2 . Conditions (ii) and (iii) mean that
there exists an element c ∈ A with (a, c) ∈ θ1 and (c, b) ∈ θ2 , and therefore
Assume now that conversely (∆A , A×A) is the only pair with the properties
(i) - (iii), and let A ∼
= A1 × A2 . Then (∆A1 ×A2 , (A1 × A2 )2 ) is also the only
pair of congruence relations on A1 × A2 to satisfy conditions (i) - (iii). But
by Lemma 4.1.4, the kernels of the projection mappings p1 and p2 do satisfy
the three conditions. Therefore one of ker p1 or ker p2 must equal ∆A1 ×A2 ,
and thus one of A1 or A2 must have cardinality one.
pk (B) = Ak .
3. Consider the two lattices C2 and C3 , chains on two and three elements,
respectively, shown below:
•3
•2
•b
•1
•a
•(b, 3)
•(a, 1)
.
• (b, 3)
(b, 2) • • (a, 3)
• (a, 2)
• (a, 1)
Theorem 4.2.3 Let B be a subdirect product of the family (Aj )j∈J of alge-
Aj → Ak satisfy the
Q
bras of type τ . Then the projection mappings pk :
j∈J
ker(pj | B) = ∆B .
T
equation
j∈J
Proof: Let (a, b) ∈ ker(pj |B). This implies that pk (a) = pk (b) for all
T
j∈J
k ∈ J, so that every component of a agrees with the corresponding compo-
nent of b. This means that a = b and thus (a, b) ∈ ∆B . Conversely, it is
clear that ∆B ⊆ ker(pk |B) for all pk .
As was the case for direct products, it turns out that this property of the
kernels of the projection mappings can be used to characterize subdirect
products, in the sense that any set of congruences on an algebra with these
properties can be used to express the algebra as a subdirect product.
ϕ(a) = ϕ(b) implies [a]θj = [b]θj and thus (a, b) ∈ θj for all j ∈ J. There-
fore (a, b) ∈ θj = ∆A and so a = b. This proves the isomorphism of A
T
j∈J
(A/θj ) → A/θk denotes the k-th projection
Q
and ϕ(A). Moreover, if pk :
j∈J
mapping, then by the definition of ϕ we have pk (ϕ(A)) = A/θk for all
k ∈ J. Therefore ϕ(A) is a subdirect product of the algebras A/θj .
• A2
• {ConA \ {∆A }}
T
• ∆A
4.3 Exercises
4.3.1. Prove that if A is a finite algebra, then A is isomorphic to a direct
product of directly irreducible algebras.
4.3.3. Let A and B be algebras of the same type. Show that if A has a
one-element subalgebra, the direct product A × B has a subalgebra which is
4.3. EXERCISES 73
isomorphic to B.
4.3.4. Prove that for any algebra A, the diagonal relation ∆A on the set A
is the universe of a subalgebra of the algebra A × A.
4.3.5. We can define an operator P on the class Alg(τ ) of all algebras of type
τ , as follows. For any class K ⊆ Alg(τ ), let P(K) be the class of all algebras
of type τ which are products of one or more algebras in K. Prove that this
operator P is extensive and monotone, but not idempotent, and hence is not
a closure operator on Alg(τ ). (This operator will be used again in Chapter
6.)
4.3.7. Prove the claim made in the remark following Theorem 4.2.4.
74
Chapter 5
75
76 CHAPTER 5. TERMS, TREES, AND POLYNOMIALS
Now we proceed to define this formal language in the general setting. Let
n ≥ 1 be a natural number. Let Xn = {x1 , . . . , xn } be an n-element set. The
set Xn is called an alphabet and its elements are called variables. We also
need a set {fi |i ∈ I} of operation symbols, indexed by the set I. The sets Xn
and {fi |i ∈ I} have to be disjoint. To every operation symbol fi we assign
a natural number ni ≥ 1, called the arity of fi . As in the definition of an
algebra, the sequence τ = (ni )i∈I of all the arities is called the type of the
language. With this notation for operation symbols and variables, we can
define the terms of our type τ language.
Definition 5.1.1 Let n ≥ 1. The n-ary terms of type τ are defined in the
following inductive way:
(ii) If t1 , . . . , tni are n-ary terms and fi is an ni -ary operation symbol, then
fi (t1 , . . . , tni ) is an n-ary term.
(iii) The set Wτ (Xn ) = Wτ (x1 , . . . , xn ) of all n-ary terms is the smallest
set which contains x1 , . . ., xn and is closed under finite application of
(ii).
5.1. TERMS AND TREES 77
2. Our definition does not allow nullary terms. This could be changed by
adding a fourth condition to the inductive definition, stipulating that every
nullary operation symbol of our type is an n-ary term. We could also extend
our language to include a third set of symbols, to be used as constants or
nullary terms; we shall explore this approach later, in Section 5.3.
Example 5.1.3 Let τ = (2), with one binary operation symbol f . Let X2 =
{x1 , x2 }. Then f (f (x1 , x2 ), x2 ), f (x2 , x1 ), x1 , x2 and f (f (f (x1 , x2 ), x1 ), x2 )
are binary terms. The expression f (f (x3 , f (x1 , x2 )), x4 ) is a quaternary or
4-ary term, but f (f (x1 , x2 ), x3 is not a term (one bracket is missing).
Example 5.1.4 Let τ = (1), with one unary operation symbol f . Let X1 =
{x1 }. Then the unary terms of this type are x1 , f (x1 ), f (f (x1 )), f (f (f (x1 ))),
and so on. Note that W(1) (X1 ) is infinite. In a specific application such as
group theory, we might denote our unary operation by −1 instead of f ,
writing our terms as x1 , x−1 −1 −1
1 , (x1 ) , etc. In the group theory case we
might want to consider the terms x1 and (x−1 1 )
−1 as equal. But such an
equality depends on a specific application, and does not hold in the most
general sense that we are defining here. Thus our terms are often called
the “absolutely free” terms, in the sense that we make no restrictions or
assumptions about the properties of our operation symbols, beyond their
arity as specified in the type.
There are various methods used to measure the complexity of a term, be-
sides the number of operation symbols which occur in it. Another common
measure is what is called the depth of the term, defined by the following
steps:
78 CHAPTER 5. TERMS, TREES, AND POLYNOMIALS
(i) If t = xi , then the semantic tree of t consists only of one vertex which
is labelled with xi , and this vertex is called the root of the tree.
(ii) If t = fi (t1 , . . . , tni ) then the semantic tree of t has as its root a vertex
labelled with fi , and has ni edges which are incident with the vertex
fi ; each of these edges is incident with the root of the corresponding
term t1 , . . . , tni (ordered by 1 ≤ 2 ≤ · · · ≤ ni , starting from the left).
Consider for example the type τ = (2, 1) with a binary operation symbol f2
and a unary operation symbol f1 , and variable set X3 = {x1 , x2 , x3 }. Then
the term t = f2 (f1 (f2 (f2 (x1 , x2 ), f1 (x3 ))), f1 (f2 (f1 (x1 ), f1 (f1 (x2 ))))) corre-
sponds to the semantic tree shown below.
x1 x2 x3 x1 x2
I 6
s s f
1
f2
s
f1 s f1
s s
f2 f1
s
cs f2
f1 s
f1
s
f2
Notice that semantic trees are ordered in such a way that we start with
the first variable on the left hand side which occurs in a term t. We can
also describe, for any node or vertex in a tree, the path from the root of
the tree to that node or vertex. That is, to any node or vertex of a term
t = fi (t1 , . . . , tni ), we can assign a sequence (or word) over the set N+ of
positive integers, as follows. Suppose the outermost term is ni -ary. The root
5.1. TERMS AND TREES 79
of the tree is labelled by the symbol e, which we call the empty sequence.
The vertices on the second level up are labelled by 1, . . . , ni , from the left
to the right. Continuing in this way, we assign to each branch of the tree a
sequence on N+ . For instance, the tree from our example above is labelled
as shown below.
I 6
s s
112
111
s 2121
s 211 s s 212
11
s
cs 21
1 s
2
s
e
Now we want to use this set Wτ (X) as the universe of some algebra, of the
same type τ . What operations can we perform on these terms? In fact, for
80 CHAPTER 5. TERMS, TREES, AND POLYNOMIALS
Note the distinction here between the concrete operation f¯i being defined
on the set of all terms, and the formal operation symbol fi , used in the for-
mation of terms. The second step of Definition 5.1.1 shows that the element
fi (t1 , . . . , tni ) in our definition belongs to Wτ (X), and so our operation f¯i
is well defined. In this way we make Wτ (X) into the universe of an algebra
of type τ = (ni )i∈I , since for every operation symbol fi we have a concrete
operation f¯i on Wτ (X).
Definition 5.1.5 The algebra Fτ (X) := (Wτ (X); (f¯i )i∈I ) is called the term
algebra, or the absolutely free algebra, of type τ over the set X.
Lemma 5.1.6 For any type τ , the term algebra Fτ (X) is generated by the
set X.
Proof: Definition 5.1.1 (i) shows that X ⊆ Wτ (X), and 5.1.1 (ii) gives
hXiFτ (X) = Fτ (X).
f
-
X A
6
ϕ fˆ
~
Fτ (X)
We have so far defined term algebras on the finite sets Xn and the countably
infinite set X, using variable symbols x1 , x2 , x3 , . . . . However, it should
be clear that we could start with any non-empty set Y of symbols, of any
cardinality, and carry out the same process. We define terms on the set
Y inductively just as in Definition 5.1.1, with all the variables in Y being
terms and then any result of applying the operation symbols fi to terms
giving terms. Then we can form the set Wτ (Y ) of all terms of type τ over Y ,
and make it into an algebra Fτ (Y ) generated by Y which has the analogous
freeness property of Theorem 5.1.7. Thus we have a free algebra over any set
of symbols of any cardinality, although we shall see in Chapter 6 that in some
sense the sets of finite or countably infinite cardinality are sufficient for our
purposes. Moreover, the next theorem shows that for a fixed cardinality, one
may use any choice of variable symbols. For example, the reader may have
noticed that in our formal language we have variables xi for i ≥ 1, while in
our example with the associative law at the beginning of this section we used
variables x, y and z. That it is justified to make such a change of variable
symbols, where convenient, follows from the following theorem:
82 CHAPTER 5. TERMS, TREES, AND POLYNOMIALS
Theorem 5.1.8 Let Y and Z be alphabets with the same cardinality. Then
the term algebras Fτ (Y ) and Fτ (Z) are isomorphic.
Proof: When |Y | = |Z| there exists a bijection ϕ : Y → Z. Since
Fτ (Z) ∈ Alg (τ ), by Theorem 5.1.7 we can extend ϕ to a homomor-
phism ϕ̂ : Fτ (Y ) → Fτ (Z). Now using Theorem 5.1.7 again on the
mapping ϕ−1 : Z → Y gives a homomorphism (ϕ−1 )ˆ : Fτ (Z) →
Fτ (Y ). We will show by induction on the complexity of the term t
that (ϕ−1 )ˆ ◦ ϕ̂ = idWτ (Y ) and ϕ̂ ◦ (ϕ−1 )ˆ = idWτ (Z) . This will prove
that ϕ̂ is an isomorphism, with (ϕ−1 )ˆ as its inverse. The claim is clear
for the base case that t = x is a variable. Now assume that t =
fi (t1 , . . . , tni ), and that the claim is true for the terms t1 , . . ., tni . Then
F (Z)
we have ((ϕ−1 )ˆ ◦ ϕ̂)(t) = (ϕ−1 )ˆ(ϕ̂(t)) = (ϕ−1 )ˆ(fi τ (ϕ̂(t1 ), . . . , ϕ̂(tni ))
F (Y )
= (ϕ−1 )ˆ(fi (ϕ̂(t1 ), . . . , (ϕ̂(tni ))) = fi τ (((ϕ−1 )ˆ ◦ ϕ)(t1 ), . . . , ((ϕ−1 )ˆ ◦
ϕ)(tni ))) = fi (t1 , . . . , tni ) = t. The proof for ϕ̂ ◦ (ϕ−1 )ˆ is similar.
fiA (tA A
1 , . . . , tni ).
In part (ii) of this definition, the right hand side of the equation refers to
the composition or superposition of operations, so that
5.2. TERM OPERATIONS 83
for all a1 , . . . , an ∈ A.
There is another way to obtain the set Wτ (X)A of all term operations on A,
using clone operations. Using A as our base set, we consider the set O(A) of
all finitary operations on A. Recall from Example 1.2.14 that the set O(A)
is closed under a composition operation
and contains all the projection operations on A. This makes O(A) a clone
on the set A, called the full clone on A; any subset of O(A) which contains
the projections and is also closed under composition is called a subclone of
O(A), or a clone on A.
Theorem 5.2.3 Let A = (A; (fiA )i∈I ) be an algebra of type τ , and let
Wτ (X) be the set of all terms of type τ over X. Then Wτ (X)A is a clone
84 CHAPTER 5. TERMS, TREES, AND POLYNOMIALS
We point out that Theorem 5.2.3 is no longer true in the case of partial
algebras, that is algebras in which the fundamental operations fi are not
totally defined on A.
Theorem 5.2.4 Let A be an algebra of type τ and let tA be the n-ary term
operation on A induced by the n-ary term t ∈ Wτ (X).
ϕ(tA (a1 , . . . , an ))
= ϕ(fiA (tA A
1 , . . . , tni )(a1 , . . . , an ))
= ϕ(fi (t1 (a1 , . . . , an ), . . . , tA
A A
ni (a1 , . . . , an )))
= fi (ϕ(t1 (a1 , . . . , an )), . . . , ϕ(tA
B A
ni (a1 , . . . , an )))
= fiB (tB 1 (ϕ(a 1 ), . . . , ϕ(an )), . . . , tB
ni (ϕ(a1 ), . . . , ϕ(an )))
B B B
= fi (t1 , . . . , tni )(ϕ(a1 ), . . . , ϕ(an ))
= tB (ϕ(a1 ), . . . , ϕ(an )).
There is one subtlety here that the reader should notice. We are defining
terms and polynomials in this chapter in a formal or general way, based only
on a type, and not on any specific algebra. Thus we should have one set of
constants, to be used in the formation of all polynomials of the given type.
However, when we consider the induced polynomial operations on a given
algebra A, we usually want the constants in our polynomials to represent
specific elements of our base set A. There are several ways to deal with this
obstacle. We will proceed by fixing one set A of constant symbols to be used
for all polynomials. Another approach is to associate to every algebra A of
type τ a corresponding set A of constants, which has the same cardinality as
the universe set A. This approach was used by Denecke and Leeratanavalee
in [30] and [31].
Let A be our set of constant symbols, pairwise disjoint from both the set X of
variables and the set {fi | i ∈ I} of operation symbols. We define polynomials
of type τ over A (for short, polynomials) via the following inductive steps:
(iv) The set Pτ (X, A) of all polynomials of type τ over A is the smallest
set which contains X ∪ A and is closed under finite application of (iii).
Much of the work we did in Section 5.1 for terms can now be carried out
for polynomials. We can define the polynomial algebra Pτ (X, A) of type τ
over A, generated by X ∪ A, and prove results similar to 5.1.6 and 5.1.7. We
leave the verification as an exercise for the reader.
The next step is to make polynomial operations over an algebra A out of
our formal polynomials. We proceed as for terms, with the addition that
we interpret the constant symbols from A by elements selected from A as
nullary operations. In this case we assume that |A| ≥ |A|, and consider a
subset A1 ⊆ A with |A1 | = |A|. Then just as for terms we obtain for every
5.3. POLYNOMIALS AND POLYNOMIAL OPERATIONS 87
Theorem 5.3.1 Pτ (X, A)A is a clone, and is generated by the set {fiA |i ∈
I} ∪ {ca |a ∈ A}, where ca is the nullary operation which selects a ∈ A. We
write Pτ (X, A)A = h{fiA |i ∈ I} ∪ {ca |a ∈ A}i.
We leave it as an exercise for the reader to prove that every polynomial
operation of an algebra A is compatible with any congruence relation on A.
In Theorem 1.4.5 we saw that an equivalence relation on an algebra A is a
congruence iff it is compatible with all translations on the algebra. Since such
translations are in fact just unary polynomial operations on the algebra, we
have the following useful result.
u = pA1 (a1 ),
pA
i (b A
i = pi+1 (ai+1 ),
) for 1 ≤ i < k,
pA
k (bk ) = v.
5.4 Exercises
L L L L
5.4.1. Let L = ({a, b, c, d}; ∧, ∨) be a lattice, with operations ∧ and ∨ given
by the following Cayley tables:
L L
∧ a b c d ∨ a b c d
a a a a a a a a c d
b a b a b b b b d d
c a a c c c c d c d
d a b c d d d d d d
Let h be the binary operation on the set L given by the Cayley table:
h a b c d
a a b a b
b b b b b
c a b a b
d b b b b
L L L
Fτ (Y ). Calculate fˆ(t) for the terms t = x ∧ y and t = (x ∨ y) ∧ z.
5.4.3. Determine all the term operations and all the polynomial operations
of the algebra (N; ¬), where ¬x := x + 1.
5.4.4. Show that if Y and Z are non-empty sets with |Y | ≤ |Z|, then the
algebra Fτ (Y ) can be embedded in Fτ (Z) in a natural way. (One algebra
can be embedded in another if the second contains an isomorphic copy of
the first.)
5.4.6. Prove that the polynomial algebra of type τ over A, from Theorem
5.3.1, satisfies properties similar to those of Lemma 5.1.6 and Theorem 5.1.7.
Our motivation for defining terms and polynomials was to use them to de-
fine equations and identities. An equation is a statement of the form t1 ≈ t2 ,
where t1 and t2 are terms. We will define what it means for such an equa-
tion to be satisfied, or to be an identity, in an algebra A. The relation of
satisfaction, of an equation by an algebra, will give us a Galois-connection
between sets of equations and classes of algebras, and allow us to consider
classes of algebras which are defined by sets of equations. Finally, we show
that such equational classes, or model classes, are precisely the same classes
of algebras as those we are interested in from the algebraic approach of the
first four chapters.
91
92 CHAPTER 6. IDENTITIES AND VARIETIES
We now consider the class Alg(τ ) of all algebras of type τ , and the class
Wτ (X) × Wτ (X) of all equations of type τ . Satisfaction of an equation by an
algebra gives us a fundamental relation between these two sets. Formally, we
have the relation |= of all pairs (A, s ≈ t) for which A |= s ≈ t. As discussed
in Section 2.2, this relation induces a Galois-connection between Alg(τ ) and
Wτ (X) × Wτ (X). We will use the names Id and M od for the two associated
mappings. That is, for any subset Σ ⊆ Wτ (X) × Wτ (X) and any subclass
K ⊆ Alg(τ ) we define
(i) For all subsets Σ and Σ0 of Wτ (X) × Wτ (X), and for all subclasses K
and K 0 of Alg(τ ), we have
0 0
Σ ⊆ Σ ⇒ M odΣ ⊇ M odΣ and K ⊆ K 0 ⇒ IdK ⊇ IdK 0 ;
(ii) For all subsets Σ of Wτ (X) × Wτ (X) and all subclasses K of Alg(τ ),
we have Σ ⊆ IdM odΣ and K ⊆ M odIdK;
(iii) The maps IdM od and M odId are closure operators on Wτ (X)×Wτ (X)
and on Alg(τ ), respectively.
6.1. THE GALOIS CONNECTION (ID, MOD) 93
(iv) The sets closed under M odId are exactly the sets of the form M odΣ,
for some Σ ⊆ Wτ (X) × Wτ (X), and the sets closed under IdM od are
exactly the sets of the form IdK, for some K ⊆ Alg(τ ).
From Theorem 6.1.3, part (iv), we see that the equational classes are ex-
actly the closed sets, or fixed points, with respect to the closure operator
M odId, and, dually, the equational theories are exactly the closed sets, or
fixed points, with respect to the closure operator IdM od. As we mentioned
in Chapter 2, the collections of such closed sets form complete lattices.
Proof: As remarked above, the fact that these collections are complete lat-
tices follows from the general results on Galois-connections and closure op-
erators in Chapter 2. For arbitrary subclasses K of L(τ ), the infimum of
K is the set-theoretical intersection, ∧K = ∩K, and the supremum is the
variety which is generated by the set-theoretical union, so K = {K 0 ∈
W T
Next we show that our map ϕ has the property claimed on meets. We have
ϕ(K1 ∧ K2 ) = Id(K1 ∩ K2 ), by definition. Since the operator Id reverses
inclusions, our set Id(K1 ∩ K2 ) contains both IdK1 and IdK2 . Since it is
an equational theory, it also contains their join, IdK1 ∨ IdK2 , which equals
ϕ(K1 ) ∨ ϕ(K2 ). This gives us the inclusion ϕ(K1 ∧ K2 ) ⊇ ϕ(K1 ) ∨ ϕ(K2 ).
For the opposite inclusion, we start with the fact that for each i = 1, 2,
we have IdKi ⊆ IdK1 ∨ IdK2 . Applying the operator M od to this, and
using the fact that Ki = M odIdKi for equational classes, we have Ki ⊇
M od(IdK1 ∨ IdK2 ). Thus we have
M od(IdK1 ∨ IdK2 ) ⊆ K1 ∩ K2 .
Thus we have our inclusion ϕ(K1 ∧ K2 ) ⊆ ϕ(K1 ) ∨ ϕ(K2 ), and hence the
equality we needed.
Finally, we verify the claim for joins. Since ϕ reverses inclusions, the inclu-
sion Ki ⊆ K1 ∨K2 , for i = 1, 2, implies ϕ(K1 ∨K2 ) ⊆ ϕ(Ki ), for i = 1, 2. This
gives us one direction, namely that ϕ(K1 ∨K2 ) ⊆ ϕ(K1 )∧ ϕ(K2 ). Conversely,
we know that IdK1 ∧ IdK2 = IdK1 ∩ IdK2 ⊆ IdKi , for i = 1, 2; so applying
M od gives M od(IdK1 ∧ IdK2 ) ⊇ M odIdKi , for i = 1, 2. Since K1 and K2
are closed under M odId, we get M od(IdK1 ∧ IdK2 ) ⊇ K1 ∨ K2 . Applying Id
once more gives IdM od(IdK1 ∧IdK2 ) ⊆ Id(K1 ∨K2 ). But IdK1 ∧IdK2 is an
equational theory and closed under IdM od; so finally we have IdK1 ∧ IdK2
⊆ Id(K1 ∨ K2 ). This amounts to ϕ(K1 )∧ ϕ(K2 ) ⊆ ϕ(K1 ∨ K2 ). This com-
pletes our proof of the equality ϕ(K1 ) ∩ ϕ(K2 ) = ϕ(K1 ∨ K2 ).
We point out that in fact the claims of Theorem 6.1.5 are true for the lat-
tices of closed sets of any Galois-connection. A careful reading of the previous
proof will show that we used only properties of the Galois connection, and
not any properties of the particular connection Id − M od.
6.2. FULLY INVARIANT CONGRUENCE RELATIONS 95
means that sA A A A
1 = t1 , . . . , sni = tni for every algebra A from K. Let fi be an
ni -ary operation symbol. Then our assumption means that fiA (sA A
1 , . . . , sni ) =
A A
fi (t1 , . . . , tni ). By the inductive step of the definition of a term operation in-
duced by a term, this means that [fi (s1 , . . . , sni )]A = [fi (t1 , . . . , tni )]A , and
the definition of satisfaction gives fi (s1 , . . . , sni ) ≈ fi (t1 , . . . , tni ) ∈ IdA.
Thus fi (s1 , . . . , sni ) ≈ fi (t1 , . . . , tni ) ∈ IdK, as required for a congruence.
For the fully invariant property, we have to show that IdK is preserved by
an arbitrary endomorphism ϕ of Fτ (X). So we take (s, t) ∈ IdK = Σ, and
show that (ϕ(s), ϕ(t)) is also in Σ. For this we use the property from Re-
mark 6.1.2, that Σ = IdK is equal to the intersections of the kernels of the
homomorphisms fˆ, for all maps f : X → A and all algebras A in K. But
for any such A and map f , the map fˆ ◦ ϕ is also a homomorphism from
Fτ (X) to A, and is the extension of some map g from X to A. Thus our pair
(s, t) from Σ must also be in the kernel of this new homomorphism fˆ ◦ ϕ.
This means precisely that the pair (ϕ(s), ϕ(t)) must be in ker fˆ. Since this
is true for all algebras A and maps f : X → A, we have (ϕ(s), ϕ(t)) in Σ.
This shows that Σ is a fully invariant congruence.
For the opposite inclusion, we use Remark 6.1.2 again; it will suffice to show
that for any map f : X → Fτ (X)/θ, we have θ ⊆ ker fˆ, where as usual fˆ is
the unique extension of f . To show this, we first define a map g : X → Fτ (X)
by x 7→ s in such a way that we assign to x a representative s ∈ fˆ(x) = [x]θ .
Then we get the commutative diagram shown below, with ϕ the natural
embedding. Combining ĝ ◦ ϕ = g and natθ ◦ g = fˆ ◦ ϕ gives us natθ ◦ ĝ ◦ ϕ
= fˆ ◦ ϕ, and since ϕ is an embedding we have natθ ◦ ĝ = fˆ. Now for any
(s, t) ∈ θ, the full invariance of θ means that (ĝ(s), ĝ(t)) ∈ θ. Therefore
fˆ(s) = [ĝ(s)]θ = [ĝ(t)]θ = fˆ(t), and (s, t) ∈ kerfˆ.
6.3. THE ALGEBRAIC CONSEQUENCE RELATION 97
6
(=) (=)
ĝ
ϕ fˆ
w /
Fτ (X)
As a consequence of this theorem, we see that the set of all fully invari-
ant congruences on the free algebra Fτ (X) forms a complete lattice, called
Conf i Fτ (X), and that this lattice is a sublattice of the congruence lattice
ConFτ (X).
(1) ∅ ` s ≈ s,
(2) {s ≈ t} ` t ≈ s,
(3) {t1 ≈ t2 , t2 ≈ t3 } ` t1 ≈ t3 ,
0 0 0
(4) {tj ≈ tj : 1 ≤ j ≤ ni } ` fi (t1 , . . . , tni ) ≈ fi (t1 , . . . , tni ), for every
operation symbol fi (i ∈ I) (the replacement rule),
(5) Let s, t, r ∈ Wτ (X) and let s̃, t̃ be the terms obtained from s, t by
replacing every occurrence of a given variable x ∈ X by r. Then s ≈
t ` s̃ ≈ t̃. (This is called the substitution rule.)
98 CHAPTER 6. IDENTITIES AND VARIETIES
It should be clear that these rules (1) - (5) reflect the properties of a fully
invariant congruence relation: the first three are the properties of an equiva-
lence relation, the fourth describes the congruence property and the fifth the
fully invariant property. Thus by Theorem 6.2.2 equational theories Σ are
precisely sets of equations which are closed with respect to finite application
of the rules (1) - (5). This equational approach will be used in Chapter 7,
when we study term-rewriting systems.
6 0, then FK (Y ) ∼
3. If |Y | = |Z| = = FK (Z) under an isomorphism mapping
Y to Z. The proof of this fact is similar to the proof of Theorem 5.1.8. This
means that (up to isomorphism) only the cardinality of the generating set
Y is important, and not the particular choice of variable symbols.
Theorem 6.4.3 Let Y be any non-empty set of variables. For every algebra
A ∈ K ⊆ Alg(τ ) and every mapping f : Y → A, there exists a unique
homomorphism fˆ : FK (Y ) → A which extends f .
f
-
Y A
KA
A
A
A
ϕ f A fˆ
A
A
A
U A
A
natIdK
Fτ (Y ) - FK (Y )
Theorem 6.4.4 means that free algebras relative to K ⊆ Alg(τ ) can be char-
acterized and therefore defined by the freeness property from Theorem 6.4.3.
6.5. VARIETIES 101
Example 6.4.5 Let τ be the type (2), so that we have one binary operation
symbol which we will denote by f . Let K be the class of all semigroups, that
is, of type (2) algebras which satisfy the associative identity, and let Y be any
non-empty set of variables. We consider the K-free algebra FK (Y ) over Y . It
is customary in this case to indicate the binary operation f by juxtaposition,
writing xy for the term f (x, y). Moreover, since x(yz) ≈ (xy)z is an identity
of K, we see that any two terms in Wτ (Y ) which have the same variables
occurring in the same order are equivalent under the congruence IdK. This
means that we can write any term in a normal form in which we omit the
brackets. For example, the term f (f (f (x, y), f (y, x)), f (z, y)) can be written
as xyyxzy. We refer to terms in this normal form as words on the alphabet
Y . It is easy to verify that the set of all such words forms a semigroup under
the operation of concatenation of words, called the free semigroup on Y , and
usually denoted by Y + . By adjoining an empty word e to Y + to act as an
identity element, we can also form the free monoid Y ∗ on the alphabet Y .
6.5 Varieties
In this section we link together our two approaches to classes of algebras, the
equational approach from the preceding sections and the algebraic approach
from Chapters 1,3 and 4. We introduce operators H, S and P on classes
of algebras, corresponding to the algebraic constructions of homomorphic
images, subalgebras and product algebras studied earlier. A class of algebras
which is closed under these operators is called a variety. Our main theorem
in this section will show that in fact varieties are equivalent to equational
classes.
Definition 6.5.1 We define the following operators on the set Alg(τ ) of all
algebras of a fixed type τ . For any class K ⊆ Alg(τ ),
I(K) is the class of all algebras which are isomorphic to algebras from K,
PS (K) is the class of all subdirect products of families of algebras from K.
We can also combine these operators to produce new ones; we write IP for
instance for the composition of I and P. Recall from Chapter 2 that an
operator is a closure operator if it is extensive, monotone and idempotent.
We first verify that some of our operators are in fact closure operators.
Lemma 6.5.2 The operators H, S and IP are closure operators on the set
Alg(τ ).
Proof: We will give a proof only for H; the others are quite similar. It is clear
from the definition that for any subclasses K and L of Alg(τ ), the inclusion
K ⊆ L implies H(K) ⊆ H(L), and since any algebra is a homomorphic im-
age of itself under the identity homomorphism, we always have K ⊆ H(K).
For the idempotency of H, we note that by the extensivity and monotonicity
we have H(K) ⊆ H(H(K)). Conversely, let A be in H(H(K)). Then there
exists an algebra B ∈ H(K) and a surjective homomorphism ϕ : B → A. For
B ∈ H(K) there exists an algebra C ∈ K and a surjective homomorphism
ψ : C → B. Then the composition ϕ ◦ ψ : C → A is also a surjective homo-
morphism, and thus A ∈ H(K).
Note however that the operator P is not a closure operator, since it is not
idempotent: A1 × (A2 × A3 ) is not equal to (A1 × A2 ) × A3 , although they
are isomorphic.
Theorem 6.5.5 For any class K of algebras of type τ , the class HSP(K)
is the least (with respect to set inclusion) variety which contains K.
Proof: We show first that HSP(K) is indeed a variety, that is, that it is
closed under application of H, S and P. We have H(HSP(K)) = HSP(K)
by the idempotence of H, and S(HSP(K)) ⊆ H(SSP(K)) = HSP(K) by
Lemma 6.5.4 (i) and the idempotence of S. For P, we have P(HSP(K)) ⊆
(HPSP(K)) ⊆ HSPP(K) ⊆ HSIPIP(K) = HSIP(K) ⊆ HSHP(K) ⊆
HHSP(K) = HSP(K), using properties from 6.5.2 and 6.5.4.
For any class K of algebras of the same type, the variety HSP(K) from
Theorem 6.5.5 is called the variety generated by K. It is often denoted by
V (K). When K consists of a single algebra A, we usually write V (A) for the
variety generated by K.
Our goal is to prove that equational classes and varieties are the same thing.
The next Lemma gives one direction of this equivalence, that any equational
class of algebras is a variety.
Before we can show the other direction of our equivalence, we need one more
fact. We show that for any class K of algebras of type τ , and for any set Y
of variables, the free algebra FK (Y ) with respect to K belongs to ISP(K).
Theorem 6.5.9 For every class K ⊆ Alg(τ ) and every non-empty set Y of
variables, the relatively free algebra FK (Y ) is in ISP(K).
Proof: We want to use Theorem 4.2.4 to write our algebra FK (Y ) as a sub-
direct product. To do this, we first need to verify that the intersection of all
the congruences on FK (Y ) is the identity relation. First note that FK (Y ) is
the quotient Fτ (Y )/IdK. Any congruence on this algebra is the kernel of a
homomorphism onto some algebra A in K, and any such homomorphism is
the extension fˆ of some mapping f : Y → A. Thus it is enough to show that
the intersection of the kernels of all such fˆ on Fτ (Y ) is the identity relation
on Fτ (Y )/IdK. But this holds, by Remark 6.1.2 and the definition of IdK.
Since varieties are also closed under the operators I, S and P, it follows from
this theorem that every variety K contains all the relatively free algebras
FK (Y ), for each non-empty set Y .
106 CHAPTER 6. IDENTITIES AND VARIETIES
K |= s ≈ t ⇔ FK (X) |= s ≈ t.
Proof: Since K is a variety, we know by the remark just above that the
relatively free algebra FK (X) is in K. This means that any identity of K
must in particular hold in FK (X), giving us one direction of the claim.
If conversely FK (X) satisfies s ≈ t, then sFK (X) = tFK (X) . Since FK (X) is
the quotient of Fτ (X) by IdK, this forces [s]IdK = [t]IdK ; and from this, we
have (s, t) ∈ ker natIdK = IdK and so K satisfies s ≈ t.
With these results, we are ready to prove our main theorem, sometimes
referred to as Birkhoff’s Theorem.
Note that in the proof of Theorem 6.5.11, we needed the existence of rel-
atively free algebras over non-empty variable sets of arbitrary cardinality.
However, as we commented in Chapter 5, it is in some sense sufficient to have
absolutely or relatively free algebras over variable sets of finite or countably
infinite cardinality. The following Lemma explains this more precisely.
finite or countably infinite set of variables. Thus the two generating sets are
contained in K. For the converse, let A be any algebra in K. If we choose
a set Y whose cardinality is greater than the cardinality of A, we can make
a surjective homomorphism from FK (Y ) onto A. Then using Lemma 6.5.12
we can express FK (Y ) as a subdirect product of relatively free algebras with
respect to K on sets E of finite cardinality. Thus A is a homomorphic image
of a subdirect product of the algebras in our generating set. This shows that
K is contained in HSP({FK (n) | n ∈ N}), and hence we have equality.
Theorem 6.5.15 A variety K is locally finite iff the relatively free algebra
FK (Y ) is finite for every non-empty finite set Y .
For the converse, let A be a finitely generated algebra from K, with a finite
set B ⊆ A of generators. Now we choose an alphabet Y in such a way that
there exists a bijection α : Y → B. This bijection can be extended to a
homomorphism α̂ : FK (Y ) → A. The image α̂(FK (Y )) is then a subalgebra
of A containing B, and hence must be equal to A. Therefore α̂ is surjective,
and as FK (Y ) is finite so is A.
Theorem 6.5.16 Let K ⊆ Alg(τ ) be a finite set of finite algebras. Then the
variety V (K) generated by K is a locally finite variety.
6.6. THE LATTICE OF ALL VARIETIES 109
Proof: We will verify first that the class P(K) is locally finite. We define an
equivalence relation ∼ on Wτ ({x1 , . . . , xn }) by p ∼ q iff the term operations
corresponding to p and q are the same for each member of K. The finiteness
condition shows that ∼ has finitely many equivalence classes. Subalgebras of
P(K) are also finite, since only finite sets can be produced from finite sets
using finitary operations. Since every finitely generated member of V (K) =
HSP(K) is a homomorphic image of a finitely generated member of SP(K),
we see that V (K) is locally finite.
It is known that the variety of all Boolean algebras and the variety of all dis-
tributive lattices are minimal. A variety of groups is minimal if and only if it
is abelian of some prime exponent (that is, consists of all abelian groups sat-
isfying the identity xp ≈ e where p is a fixed prime number). For semigroup
varieties of type (2), using the convention of replacing the binary operation
symbol by juxtaposition, we have the following minimal varieties:
of type (2) found by V. L. Murskij in [83]; this is the algebra with base set
A = {0, 1, 2} and binary operation given by
0 1 2
0 0 0 0
1 0 0 1
2 0 2 2
Although the set of all identities of an algebra A may not be finitely axiom-
atizable, we can show that the set of all identities which use only a finite
number of variables is finitely based.
Σ = {x ≈ y | x, y ∈ Xm and (x, y) ∈ θ} ∪
θ ∧ ψ = θ ∧ ϕ =⇒ θ ∧ ψ = θ ∧ (ψ ∨ ϕ).
R. McKenzie proved in [77] that a locally finite variety V having only finitely
many subdirectly irreducible elements and having an additional property
called definable principal congruences is finitely axiomatizable. McKenzie
also proved in [81] that there are only countably many values possible for
6.8. EXERCISES 113
the residual bound of a finite algebra. This residual bound must be either
∞ or one of the following cardinals:
0, 3, 4, . . . , ω, ω1 , (2ω )+ ,
where ω is the cardinal number of the set of natural numbers (that is, the
first infinite cardinal), ω1 = ω + is the next largest cardinal number after
ω, and (2ω )+ is the successor cardinal of the cardinal of the continuum.
Recently R. Willard proved the following important theorem about finite
axiomatizability, which generalizes Baker’s Theorem.
6.8 Exercises
6.8.1. Verify that the pair (Id, M od) forms a Galois-connection between the
sets Alg(τ ) and Wτ (X)2 .
6.8.2. Prove that the set of all fully invariant congruence relations Conf i A
of an algebra A forms a sublattice of the lattice ConA of all congruence
relations on A.
6.8.4. Let L be the variety of all lattices. Determine all elements of FL ({x})
and of FL ({x, y}).
6.8.6. Show that ISP(K) is the smallest class containing K and closed under
I, S and P.
6.8.7. Let V be a variety and let Y and Z be non-empty sets with |Y | ≤ |Z|.
Show that FV (Y ) can be embedded in FV (Z) in a natural way.
114 CHAPTER 6. IDENTITIES AND VARIETIES
6.8.8. Prove that in the variety of all algebras of type τ = (3) defined by the
identities
6.8.9. Using the normal form for semigroup terms from Example 6.4.5, de-
scribe the free semilattice on the set Xn of n generators.
Chapter 7
Quotient algebras and relatively free algebras are examples of a more gen-
eral construction. Given any set A and any equivalence relation on A, we
can form the quotient set A/θ of all the equivalence classes with respect
to θ. Elements of A/θ are classes or sets of equivalent elements of A, but
calculations on such classes are always done by choosing a representative ele-
ment from each class, and calculating with these representatives. This means
that it is important to be able to check whether two elements belong to the
same equivalence class. Given any two elements a and b of A, we must check
whether the pair (a, b) is in our original equivalence relation.
115
116 CHAPTER 7. TERM REWRITING SYSTEMS
7.1 Confluence
We shall be interested in testing for equivalence of elements with respect to
an equivalence relation on a set A. Any equivalence relation on A is a binary
relation on A, and any binary relation generates an equivalence relation. To
describe the equivalence relation generated by a relation ρ on A, we need
the following notation:
7.1. CONFLUENCE 117
ρSRT = (ρ ∪ ρ−1 ∪ 4A )T .
Lemma 7.1.1 For any two binary relations ρ1 and ρ2 on a set A, we have
ρ2 = 4A ∪ ρ1 ◦ ρ2 ⇒ ρRT
1 ⊆ ρ2 .
(0)
Proof: Let ρ2 = 4A ∪ ρ1 ◦ ρ2 . Then by definition both ρ1 = 4A ⊆ ρ2 and
(n−1) (n) (n−1)
ρ1 ◦ ρ2 ⊆ ρ2 . Inductively, if ρ1 ⊆ ρ2 then ρ1 = ρ1 ◦ ρ1 ⊆ ρ1 ◦ ρ2 ⊆ ρ2 .
(n)
Thus ρ1 ⊆ ρ2 for all natural numbers n, and we have
ρRT
1 = ∪ ρ(i) ⊆ ρ2 .
i≥0
ρ1 ⊆ ρ2 ⇒ ρRT RT
1 ⊆ ρ2 and (ρRT
1 )
RT
= ρRT
1 .
Also, transitivity of ρRT means that ρRT RT = ρRT . This shows us that
1 ◦ ρ1 1
ρ1 ◦ ρRT
2 ⊆ ρRT RT
2 ◦ ρ1 ⇒ (ρ1 ◦ ρRT
2 )
(n)
⊆ ρRT RT
2 ◦ ρ1 ,
for all n ∈ IN (see Exercise 7.5.2). We have now proved the following result.
ρ1 ◦ ρRT
2 ⊆ ρRT RT
2 ◦ ρ1 ⇒ (ρ1 ◦ ρRT
2 )
RT ⊆ ρRT ◦ ρRT .
2 1
Proposition 7.1.4 For any two binary relations ρ1 and ρ2 on a set A the
following conditions are equivalent:
(i) ρ1 ◦ ρRT RT RT
2 ⊆ ρ2 ◦ ρ1 ;
(ii) ρRT RT RT RT
1 ◦ ρ2 ⊆ ρ2 ◦ ρ1 ; and
(0)
Proof: (i) ⇒ (ii): Clearly, ρ1 ◦ ρRT2 = 4A ◦ ρRT 2 = ρRT2 ⊆ ρRT RT
2 ◦ ρ1 . In-
(n−1)
ductively, if for some n ≥ 0 we have ρ1 ◦ ρRT
2 ⊆ ρRT RT
2 ◦ ρ1 , then also
(n) (n−1)
ρ1 ◦ ρRT
2 = ρ1 ◦ (ρ1 ◦ ρRT RT
2 ) ⊆ (ρ1 ◦ ρ2 ) ◦ ρ1
RT ⊆ ρRT ◦ (ρRT ◦ ρRT ) =
2 1 1
(n)
RT RT RT
ρ2 ◦ ρ1 . This means that for all n ≥ 0, we have ρ1 ◦ ρ2 ⊆ ρ2 ◦ ρRT RT
1 .
(i) RT RT RT RT RT
From this we have ∪ ρ1 ◦ ρ2 = ρ1 ◦ ρ2 ⊆ ρ2 ◦ ρ1 .
i≥0
Our aim is to obtain a decision procedure for the equivalence relation ρSRT
generated by a binary relation ρ on a base set A. We shall see that this will
require some finiteness restrictions. We introduce now a change of notation:
instead of the algebraic notation ρ for a binary relation on a set A, we will
use the notation → more commonly used in Computer Science. The state-
ment (a, b) ∈ ρ becomes a → b. For the equivalence relation generated by
→, we first form the inverse relation ← := →−1 and the symmetric closure
RT
←→ defined as → ∪ ←; then we use the reflexive transitive closure ←→ of
this.
a = x0 → x1 → · · · → xm ↓ .
a = x0 → x1 → · · · xs → · · · .
It is easy to see that if the relation → contains any pairs (a, a), or any pairs
(a, b) and (b, a), for elements a and b of A, non-terminating reductions will
be possible, and so → will not be terminating. As we will see in Section
7.3, we usually start with relations → which are both irreflexive and anti-
symmetric, to avoid this problem, and later form the reflexive, symmetric,
transitive closure of the relation.
a ←→ b
↓ &.
c d
RT RT RT
Here c and d are in normal form, and we have a ←→ b, a −→ c and b −→ d,
but c 6= d.
7.1. CONFLUENCE 121
In general then we want to start with an element a in our base set, and
reduce it to some normal form c. However, it is possible, in an arbitrary
relation →, that there may be many different reductions starting from a,
which may or may not converge to the same result. In particular, an element
a may have no normal form, if the reduction relation → is not terminating,
or it may have more than one normal form reached by different reductions
from a. Some kinds of restrictions used to ensure a unique normal form for
every element are described in the next definition.
y
/ ^
x
RT RT
w/
z
Theorem 7.1.8 can be used to decide whether any pair is in the equivalence
RT
relation ←→. But in order to use this test we have to know that → has
the Church-Rosser property, and in general it is a hard problem to show
that a relation has this property. In the next sections some lemmas are
derived which allow us to reduce the “global” problem of proving the Church-
Rosser property to a “local” problem. But first we prove that having the
Church-Rosser property is equivalent to being confluent, a result known as
the “Church-Rosser-Theorem.”
RT RT RT
(← ◦ −→)n ⊆ −→ ◦ ←− for all n ∈ IN. This is clearly satisfied for n = 0. If
RT RT RT
(← ◦ −→)n−1 ⊆ −→ ◦ ←−, then
RT RT RT
(← ◦ −→)n = ← ◦ −→ ◦(← ◦ −→)n−1
RT RT RT
⊆ ← ◦ −→ ◦ −→ ◦ ←− by induction hypothesis
RT RT RT RT RT
= ← ◦ −→ ◦ ←− since −→ ◦ −→ = −→
RT RT RT
⊆ −→ ◦ ←− ◦ ←− by confluence
RT RT RT RT RT
= −→ ◦ ←− since ←− ◦ ←−=←− .
RT RT RT
Therefore, we get (← ◦ −→)n ⊆ −→ ◦ ←− for all n ∈ IN, and so
RT RT RT RT RT RT
{(← ◦ −→)n | n ∈ IN} ⊆ −→ ◦ ←−. Consequently ←→ ⊆ −→ ◦ ←−,
S
x R y
RT R/ RT
z
r r r - r
-
For the proof of our next result we need the following Principle of Noetherian
Induction.
RT RT
P(a) iff ∀x, y ∈ A (a −→ x and a −→ y ⇒ x ↓RT y).
We verify that
Since a complete reduction system is one which is both confluent and termi-
nating, we can rephrase Theorem 7.2.4 as follows.
Reduction systems will be used to help us to solve the word problem. Given
an equivalence relation ∼ on A, we have to find a complete reduction system
which generates the relation ∼. Of course, different reduction systems can
generate the same equivalence relation. Such reduction systems are called
equivalent. Thus we want to find, among all the equivalent reduction systems
which generate our equivalence relation, any complete reduction systems.
Definition 7.2.6 Two reduction systems (A; →) and (A; ⇒) are called
RT RT
equivalent if ←→ = ⇐⇒. A complete reduction system which is equivalent
to (A; →) is called a completion of (A; →).
126 CHAPTER 7. TERM REWRITING SYSTEMS
for all x, y ∈ A.
RT RT
Clearly ⇐⇒ = ←→, so that the reduction systems (A; →) and (A; ⇒)
RT
are equivalent. Moreover, if x ⇐⇒ y then there is an element z with
RT RT
x =⇒ z ⇐= y, namely z = S(x) if x 6= y and x = z = y otherwise. Therefore
⇒ has the Church-Rosser property and is confluent by Theorem 7.1.10. The
relation ⇒ is also terminating and therefore complete. This shows that any
reduction system does have a completion.
For a terminating reduction system (A; →), we know by Corollary 7.2.5 that
local confluence is enough to guarantee completeness. To find a completion,
therefore, we can try the following construction, based on the strategy of
locating all cases in which local confluence is violated. We consider the set
of all so-called “critical pairs,” that is, pairs (x, y) ∈ A × A for which there
is an element z ∈ A such that z → x and z → y, but no element z 0 ∈ A
RT RT
satisfying x −→ z 0 and y −→ z 0 . If there are no such pairs, then the relation
→ is locally confluent and hence confluent and complete. Otherwise, if there
exists at least one critical pair (x, y) we can try to fix the local confluence
violation by adding either (x, y) or (y, x) to the reduction relation →. That
is, we add either x → y or y → x. However, it is possible that with this
addition the new enlarged relation is no longer terminating. If both possible
7.2. REDUCTION SYSTEMS 127
additions destroy the termination property, our procedure stops with “fail-
ure.”
The problem with this procedure is that in general there can be infinitely
many critical pairs. So this procedure does not give an algorithm for con-
struction of the completion of a terminating reduction system, and in general
there is no such algorithm.
The reduction system (A; →) is terminating since any reduction step de-
creases the length of the words. But it is not confluent, since we have the
reductions
yyxyyxy
. &
yxy yyx
128 CHAPTER 7. TERM REWRITING SYSTEMS
but it is not possible to reduce yxy and yyx to a common element. Applying
our procedure to the critical pair (yxy, yyx), we check whether we should
add yxy → yyx or yyx → yxy. Since both words have equal length, we
use the lexicographical order induced by x > y to make our decision: since
yxy > yyx we add the rule yxy → yyx. Now we consider the relation →1
generated by the rules r0 = yyxy →1 e and r1 = yxy →1 yyx, where e is the
empty word (having the property we ≈ ew ≈ w for all words w). Our new
relation →1 is defined by w →1 w0 if there are words w1 , w2 ∈ A such that
either
The reduction system (A; →1 ) is still not confluent. The overlappings of the
left hand sides of the two rules give new rules. The overlapping of the left
hand side of r0 with itself was already examined. Consider now r0 and r1 .
yyxy
. &
e yyyx
yyxyxy
. &
yyxyxy yx
yyxyxy
. &
xy yyxyyx
yxyyxy
. &
yyxxy yxyyx
yyxyxy
. &
xy yx
But if we now add the rule r6 = xy →3 yx, it is easy to show that no other
overlappings produce critical pairs. We denote by →3 the relation generated
by r0 , r1 , r2 , r3 , r4 , r5 and r6 . Then (A; →3 ) is a completion of (A; →). In fact
we can omit the rules r0 to r5 , since (A; →3 ) is equivalent to the complete
reduction system (A; ⇒) where ⇒ is generated by yyyx → e and xy → yx.
Thus (A; ⇒) is also a completion of (A; →).
The motivation for this algorithm lies in the deductive method for equational
theories in Universal Algebra. That is, we want to form our deduction rules
for terms from some identities of an equational theory. There is an impor-
tant difference, however, in that identities t ≈ t0 are symmetric, while our
deduction rules t → t0 are one-way only. We usually start with an irreflexive
and anti-symmetric set of rules as our set →, and later form the reflexive,
symmetric and transitive closure of this, as in the previous sections.
As we saw in Section 6.3, there are five rules of deduction for equational
theories, which allow us to deduce new equations from given ones. Of these
rules, the first three correspond to the three basic properties of an equiva-
RT
lence relation. As before, these are taken care of by taking the closure ←→
of →. We will focus now on the other two rules, the substitution rule and
the so-called “replacement” rule which allows us to apply fundamental op-
eration symbols or (from Theorem 5.2.4) arbitrary terms to equations to
produce new equations. We will call a relation invariant if it is closed under
the replacement rule, rule (4), and fully invariant if it is closed under both
the replacement rule and the substitution rule. The word problem for fully
invariant congruences can thus be transformed into a reduction problem for
fully invariant relations, and the first condition we impose on our relation
→ is that it should be fully invariant.
Then a relation → on the free algebra Fτ (X) is fully invariant iff for all
terms t, t0 and t00 , for all substitutions s : X → Wτ (X), and all addresses u
in t00 ,
It is clear that these two conditions express the deduction rules (4) and (5),
the replacement rule and the substitution rule, respectively. The following
proposition gives a necessary condition for a relation on Wτ (X) to be ter-
minating. We use the notation var(t) for the set of all variables which occur
in a term t.
Since z does not occur in t it is clear that ŝ[t] = t for the extension of s. We
set t0 := ŝ[t] and t1 := ŝ[t0 ]. Then from t → t0 and condition (2) for term
reductions we get t0 → t1 .
t0 → t1 → t2 → · · · .
132 CHAPTER 7. TERM REWRITING SYSTEMS
This necessary condition for a terminating reduction is built into our defini-
tion of a term reduction system, along with the requirement of full invariance.
Assume conversely that t →R t0 . Then there are terms t01 , t02 with t01 →R t02 ,
an address u in t and a substitution s : X → Wτ (X), such that t/u = ŝ[t01 ]
and t0 /u = ŝ[t02 ]. Using (2) and (1) we have t0 [u/ŝ[t02 ]] = t0 . Continuing in
this way we come finally to a reduction rule t1 → t2 ∈ R with this property
and our condition is satisfied.
7.3. TERM REWRITING 133
All the properties of reduction systems described in Section 7.2 can be ap-
plied to term rewriting systems. Given a term rewriting system R, we have
to find an equivalent complete terminating rewriting system. In this section
we assume that our system is terminating. The termination of term rewrit-
ing systems will be considered in the next section. As we did in the example
given in Section 7.2, we proceed by transforming critical pairs into new re-
duction rules.
To check the local confluence of our term rewriting system R, any two diverg-
ing one-step reductions t → t0 and t → t00 must be inspected for a common
reduction t such that t0 and t00 converge to t. The easiest way is to find a
substitution s under which t0 and t00 are equal, making t = ŝ[t0 ] = ŝ[t00 ].
tq
t0 R t00
s s
R/
ŝ[t0 ] = ŝ0 [t00 ] = t̄
s1 ∼ s2 :⇔ s1 ≤ s2 and s2 ≤ s1 .
Proof: We consider the set UM of all unifiers for M and its quotient set
UM / ∼ when we factorise by the relation ∼. Clearly a most general unifier
for M is a maximal element with respect to the partial order ≤ on this
quotient set. Assume that s is not maximal in UM / ∼. Then there is an
element s1 in UM / ∼ such that s < s1 . If s1 is maximal, we are finished.
Otherwise, by iteration of this procedure (using the axiom of choice) we get
a sequence s1 , s2 , s3 , . . . such that
Since < is noetherian this sequence must terminate, at some sm for which
there is no element sm+1 with sm < sm+1 . Therefore sm is a most general
unifier.
2. Let τ = (2) and let Σ be the equational theory of the variety of all
commutative semigroups, that is, the equational theory generated by
Consider the terms t = f (a, x1 ) and t0 = f (b, x2 ), where a and b are variables
different from x1 and x2 . Then the substitution s1 defined by s1 (x1 ) = b,
s1 (x2 ) = a and s1 (xj ) = xj for all j ≥ 3 is a Σ-unifier for t and t0 , since ŝ1 [t]
= f (a, b) ≈ f (b, a) = ŝ1 [t0 ] is an identity in Σ.
The next example illustrates the process of replacing critical pairs by new
reduction rules, in our procedure for finding the completion of a given term
rewriting system.
Example 7.3.7 Consider the type τ = (2, 1, 0) with the binary operation
symbol ·, the unary operation symbol −1 and the nullary operation symbol
e. We start with three rules r0 , r1 and r2 .
r0 : (x1 · x2 ) · x3 → x1 · (x2 · x3 ),
r1 : x−1 · x → e, and
r2 : e · x → x.
These rules clearly arise from the axioms of a group, with the imposition
of a definite (one-way) orientation. We will show later, in Example 7.4.5,
that the term rewriting system R defined by these three rules is termi-
nating. But it is not confluent, since we can find some critical pairs. For
instance, we have (using r1 ) that (x−1 1 · x1 ) · x3 → e · x3 , and (using r0 ) that
(x1 · x1 ) · x3 → x1 · (x1 · x3 ). But the term x−1
−1 −1
1 · (x1 · x3 ) cannot be further
reduced with our three rules; so the pair (e · x3 , x−1 1 · (x1 · x3 )) is a critical
pair. We can reduce the term e · x3 to x3 , using r2 , and we see from this
that x−1
1 · (x1 · x3 ) ≈ x3 is an identity in the variety of all groups. But the
corresponding rule x−11 · (x1 · x3 ) → x3 cannot be derived from the rules r0 ,
r1 and r2 . Therefore, we add a new reduction rule:
r3 : x−1
1 · (x1 · x3 ) → x3 .
For notational convenience, we shall not distinguish here between rules from
the original R and new rules in the successive enlargements of R. Now from r3
and r1 we can get (x−1 −1 −1 −1 −1 −1 −1 −1
2 ) ·(x2 ·x2 ) → x2 and (x2 ) ·(x2 ·x2 ) → (x2 ) ·e.
This gives a new critical pair (x2 , (x−1
2 )
−1 · e) which cannot be further re-
r4 : (x−1
2 )
−1 · e → x .
2
136 CHAPTER 7. TERM REWRITING SYSTEMS
r5 : e−1 · x3 → x3 .
We now use the concept of a most general unifier for a pair of terms, to make
our definition of a critical pair.
7.3. TERM REWRITING 137
Table 1
138 CHAPTER 7. TERM REWRITING SYSTEMS
Definition 7.3.9 Let R be a term rewriting system and let t1 → t01 and
t2 → t02 be two reduction rules of R. We may assume that var(t1 )∩ var(t2 ) =
∅, since otherwise the variables can be renamed in an appropriate way. Let
v be an address of t1 such that t1 /v is a subterm which is not a variable. If
there is a most general unifier s of t1 /v and t2 (so that ŝ[t1 /v] = ŝ[t2 ]), then
the pair (ŝ[t01 ], ŝ[t1 ][v/ŝ[t02 ]]) is called a critical pair in R.
Remark 7.3.10 If we reduce ŝ[t1 ] using the rules t1 → t01 and t2 → t02 , then
we obtain ŝ[t1 ][v/ŝ[t02 ]].
Example 7.3.11 Consider the reduction rules r3 and r12 from Example
7.3.7:
x−1
1 · (x1 · x3 ) → x3 and x1 · (x2 · (x1 · x2 )−1 ) → e.
The next lemma shows that all other “critical situations” of the kind we
have been considering are pairs which can be obtained by substitutions from
a critical pair.
Lemma 7.3.12 Let R be a term rewriting system and let t1 → t01 and t2 →
t02 be two reduction rules of R. Let v be an address of t1 such that t1 /v is a
subterm which is not a variable. If there exist substitutions s1 and s2 such
that ŝ1 [t1 /v] = ŝ2 [t2 ], then there exist a critical pair (t0 , t00 ) and a substitution
s such that
ŝ[t0 ] = ŝ1 [t01 ] and ŝ[t00 ] = ŝ1 [t1 ][v/ŝ2 [t02 ]].
Now we define a critical pair (t0 , t00 ) by t0 = m̂[t01 ] and t00 = m̂[t1 [v/r̂[t]]]. Then
we have ŝ1 [t01 ] = ˆs̄[t01 ] = ŝ[m̂[t01 ]] = ŝ[t0 ], and ŝ1 [t1 ][v/ŝ2 [t02 ]] = ˆs̄[t1 ][v/ˆs̄[r̂[t02 ]]]
= ˆs̄[t1 [v/r̂[t02 ]]] = ŝ[m̂[t1 [v/r̂[t02 ]]]] = ŝ[t00 ].
Proposition 7.3.13 A term rewriting system is locally confluent iff all its
critical pairs are convergent.
If conversely t0 and t00 have a common normal form, for all critical pairs
(t0 , t00 ), then R is locally confluent by Proposition 7.3.13. Since R is termi-
nating, it is confluent by Theorem 7.2.4. Moreover in this case the normal
form of any term is unique.
140 CHAPTER 7. TERM REWRITING SYSTEMS
The main problem here is that termination has to be preserved. For this the
following Lemma is needed.
To use this result in our algorithm, it must be decidable for each pair (t, t0 )
of terms whether t > t0 holds or not. If a critical pair (t, t0 ) is produced
during the completing process for which the two components t and t0 are
incomparable with respect to >, then the Knuth-Bendix algorithm gives an
error-message.
The inputs of the Knuth-Bendix algorithm are a finite term rewriting system
R = {(t1 , t01 ), . . . , (tm , t0m )} and a terminating fully invariant relation >. The
output is a complete term rewriting system equivalent to R, or an error
message.
The sequence (Rn | n ∈ IN) of finite term rewriting systems is produced, if
possible, as follows:
begin R−1 := ∅; n := 0;
for i = 1, . . . , m do
if ti , t0i are incomparable then stop with error message
else if ti > t0i then add ti → t0i to R0
else add t0i → ti to R0 ;
while Rn 6= Rn−1 do
7.4. TERMINATION OF TERM REWRITING SYSTEMS 141
(ii) ϕ̂(fi (t1 , . . . , tni )) = ϕ(fi ) + ϕ̂(t1 ) + · · · + ϕ̂(tni ), for any term
fi (t1 , . . . , tni ) where fi is an ni -ary operation symbol and t1 , . . . , tni
are terms in Wτ (X).
We will denote by occx (t) the number of occurrences of the variable x in the
term t.
(i) ϕ̂(t) > ϕ̂(t0 ) and occx (t) > occx (t0 ) for all x ∈ X, or
(ii) ϕ̂(t) > ϕ̂(t0 ) and occx (t) = occx (t0 ) for all x ∈ X and either
Then we can verify that this induced order is indeed a reduction order.
Proof: Let τ be a fixed type, and let ≤ be a total order on the set {fi | i ∈ I}
of operation symbols and ϕ : {fi | i ∈ I} → IN be a weight function. We
consider the induced Knuth-Bendix order >KB on Wτ (X). We have to show
that >KB is fully invariant and terminating.
Assume that t1 and t2 are terms from Wτ (X) with t1 >KB t2 , and let
t ∈ Wτ (X) be an arbitrary term. We have to prove that t[u/t1 ] >KB t[u/t2 ]
for every address u of t. By the transitivity of >KB it suffices to verify this
for all addresses u of t with u ∈ IN \ {0}.
Since t1 >KB t2 , either case (i) or case (ii) of Definition 7.4.3 holds. If case
(i) holds, and ϕ̂(t) > ϕ̂(t0 ) and occx (t) > occx (t0 ) for all x ∈ X, then also
ϕ̂(t[u/t1 ]) > ϕ̂(t[u/t2 ]) and occx (t[u/t1 ]) > occx (t[u/t2 ]) for all x ∈ X. Then
again by case (i) of Definition 7.4.3 we have t[u/t1 ] >KB t[u/t2 ]. If instead
case (ii) holds, and ϕ̂[t1 ] = ϕ̂[t2 ] and occx (t[u/t1 ]) = occx (t[u/t2 ]) for all
x ∈ X, then by the same case we have t[u/t1 ] >KB t[u/t2 ]. This shows that
>KB is closed under the replacement rule. In a similar way it can be shown
that >KB is closed under all substitutions, so that >KB is fully invariant.
Since each nullary operation symbol has positive weight, the weight w of
any nullary term is greater than or equal to k0 . Therefore for a fixed weight
w there is only a finite number of choices for k0 , k2 , k3 , · · ·. If each unary
operation symbol has a positive weight we have w ≥ k1 and therefore there
are only finitely many nullary terms of weight w. That means an infinite
chain
t0 >KB t1 >KB t2 >KB · · · >KB · · ·
zero. We define a mapping h from the set of all unary terms into itself such
that h(t) is the unary term obtained from t by replacing all occurrences of f0
by another unary term. Clearly the mapping h preserves the weight and there
are only finitely many nullary terms of weight w (by the previous remark).
Now we show that there is no infinite chain t0 >KB t1 >KB > t2 >KB · · · of
nullary terms such that h(t0 ) = h(t1 ) = h(t2 ) = · · ·. Each nullary term
t can be regarded as a word over the set of all nullary operation sym-
bols; that means t can be written in the form t = f0r1 α1 f0r2 α2 · · · f0r1 or
t = α1 f0r1 α2 f0r2 · · · αn f0rn αn , where r1 , . . . , rn are natural numbers and
α1 , . . . , αn are words built up by nullary operation symbols except f0 . Let
r(t) = (r1 , . . . , rn ) be the n-tuples consisting of the exponents of the occur-
rences of f0 .
It can be shown that if h(t) = h(t0 ) then t >KB t0 iff r(t) >lex r(t0 ), where
>lex is the lexicographic order on n-tuples. It can easily be shown that >lex
is terminating (on words of equal length). This completes the proof of Propo-
sition 7.4.4.
(x · y) · z >KB x · (y · z).
Also, (x · y)−1 >KB y −1 · x−1 by Definition 7.4.3 (ii)(b), first case, and
(y −1 )−1 > y by Definition 7.4.3 (ii)(a). This order can be used to prove
the termination of the term rewriting system for groups from Example 7.3.7
(see [66]).
The following proposition generalizes Lemma 7.3.15.
7.5. EXERCISES 145
Assume that > is a terminating invariant proper order on Wτ (X), such that
ŝ(t) > ŝ(t0 ) for all t → t0 ∈ R and all substitutions s : X → Wτ (X). Then S
is a subset of >, and since > is invariant we have → a subset of > too. Thus
R
→ is terminating.
R
Unfortunately, there are many cases for which the Knuth-Bendix ordering is
not suitable. Suppose we want to show the termination of t = x · (y −1 · y) −→
(y · y)−1 · x, using the same order on the fundamental operation symbols and
the same weight function as in Example 7.4.5. The only possible case is
Definition 7.4.3 (ii)(b), second case, but then we must have x >KB (y · y)−1 ,
and this is a contradiction.
Several other orderings have therefore been developed; see for instance the
Handbook of Formal Languages, Vol. 3 ([54]).
7.5 Exercises
7.5.1. Let ρ be a relation on a set A. Prove that
ρSRT = (ρ ∪ ρ−1 ∪ 4A )T .
7.5.2. Let ρ1 and ρ2 be relations on a set A. Prove that for all n ∈ IN,
ρ1 ◦ ρRT RT RT RT (n)
2 ⊆ ρ2 ◦ ρ1 ⇒ (ρ1 ◦ ρ2 ) ⊆ ρRT RT
2 ◦ ρ1 .
7.5.3. Let Substτ be the set of all substitutions of type τ . Let ≤ be the
relation defined on Substτ by s1 ≤ s iff there is a substitution s such that
s1 = s ◦ s2 . Prove that ≤ is a reflexive and transitive order on Substτ , and
that the relation ∼ defined by s1 ∼ s2 iff s1 ≤ s2 and s2 ≤ s1 is an equiva-
lence relation on Substτ .
146 CHAPTER 7. TERM REWRITING SYSTEMS
Algebraic Machines
147
148 CHAPTER 8. ALGEBRAIC MACHINES
U V := {uv | u ∈ U, v ∈ V }.
Definition 8.1.1 The set RegX of all regular languages over an alphabet
X is the smallest set R such that
It is easy to see from this definition that all finite languages are regular. The
set RegX is the smallest set of languages over X which contains all the finite
languages and is closed under the three regular language operations.
Example 8.1.2 Let X2 = {x1 , x2 }. Some elements of RegX2 are {x1 }, {x2 },
{x1 x2 }, {x1 x2 }∪ {x2 x2 } = {x1 x2 , x2 x2 }.
150 CHAPTER 8. ALGEBRAIC MACHINES
(i) δ̂(z, e) = z, for each state z ∈ Z, where e is the empty word, and
(ii) δ̂(z, wx) = δ(δ̂(z, w), x), for each state z ∈ Z, each letter x ∈ X and each
word w ∈ X ∗ .
The inductive step (ii) of this definition says that to compute which state
the machine is in, when it has started in state z with input word wx, we
first find what state it is in after input w, then move from that state with
input x. That is, words are read in one letter at a time, with a change of
state each time.
Now that we know how an automaton reads in words, we can describe how
an automaton acts as a language recognizer, over the alphabet X. We input
a word w to the machine, in the initial state z0 , and compute the resulting
state δ̂(z0 , w). If this state is a final state from the set Z 0 , the input word w is
recognized or accepted by the automaton; otherwise, it is said to be rejected
8.2. FINITE AUTOMATA 151
L(H) = {w ∈ X ∗ | δ̂(z0 , w) ∈ Z 0 }.
Notice that the finiteness of the state set is important here; otherwise, every
language over X would be recognizable. We now have two classes of languages
over an alphabet: the regular languages of the previous section, and the new
recognizable languages. The first important result of finite automata theory
is Kleene’s Theorem (see S. C. Kleene, [63]), which says that these two classes
are in fact the same. We can easily prove one direction of this theorem now,
that any recognizable language is regular, but the other direction will take
us more work.
Now assume that all the sets Lkab are regular. Then Lk+1 ij is the union of Lkij
with the set of all words w such that δ̂(zi , w) = zj and the state zk is used at
least once in the computation. We show that this latter set is also regular.
152 CHAPTER 8. ALGEBRAIC MACHINES
Now words in this set use state zk at least once, and we can describe them
in terms of the occurrences of zk . Up to the first occurrence of zk , no state
zm with m ≥ k is used, and the same is true between any two consecutive
uses of zk and after the last use of zk . Thus we can express any such word
w in the form uw1 · · · wl v, where u ∈ Lkik , v ∈ Lkkj and wr ∈ Lkkk for each r.
This means that we can express Lk+1 ij = Lkij ∪ Lkik (Lkkk )∗ Lkkj . By the induction
hypothesis the sets used in this expression are each regular, and thus
so is Lk+1
ij .
δ 0 1
z0 z1 z3
z1 z2 z3
z2 z3 z0
z3 z4 z4
z4 z4 z4
The graph below describes the work of this automaton. Notice that an arrow
from state zi to state zj is labelled with a letter x when δ(zi , x) = zj .
z0 0 z1
u - u
@
6@
1@
@
1 @ 1
@
0
@
1 - z4
0
u - u - u
R ?
@
z2 0 z3 0 1
Consider the word w = 0001101. The machine reads this word in one letter
at a time, from left to right, starting with input 0 in the initial state z0 . This
8.2. FINITE AUTOMATA 153
We want to show now that the language accepted by this automaton is reg-
ular, by finding a regular expression for it. We notice that the final state z3
can be reached only after a final step from states z0 , z1 or z2 , while if state z4
is reached there is no way to leave it for another state. Thus we may travel
through the states z0 , z1 and z2 as many times as we wish, but must then
finish with a transition to z3 . For instance, if the last transition used goes
from z0 to z3 , the word accepted has the form (001)n 1, for some n ≥ 1. This
gives a set of words represented by the regular expression (001)∗ 1. If the last
transition used goes from z1 or z2 to z3 , we get words of the form (001)n 01
or (001)n 000, respectively.
Thus we guess that the language L(H) is represented by the regular expres-
sion (001)∗ (1 + 01 + 000). Now we must prove that this is indeed the case,
by induction.
Next we prove the converse of the first claim, that if δ̂(z0 , w) = z0 , then
w = (001)n for some n ≥ 0. This is obvious if w is the empty word e; so
we will assume that w = w1 r. Then since δ(δ̂(z0 , w1 ), r) = δ̂(z0 , w) = z0 ,
we see that r = 1 and δ̂(z0 , w1 ) = z2 . In a similar way we get w = w2 001,
where δ̂(z0 , w2 ) = z0 . Since w2 is a shorter word than w, we conclude by the
induction hypothesis that w2 = (001)n for some n, so that w = (001)n+1 .
Our next aim is to prove the other direction of Kleene’s Theorem, by show-
ing that for any given regular expression there is a finite automaton which
recognizes the corresponding language. To do this we need the concept of a
derived language. Let L ⊂ X ∗ be a language over an alphabet X. For any
letter a ∈ X we define the derived language La of L with respect to a by
La := {w ∈ X ∗ | aw ∈ L}.
Lw := {v ∈ X ∗ | wv ∈ L}.
regular expression, since any single word has this property. We show by in-
duction on the construction of the regular expression for L that there are
only finitely many distinct languages of the form LE , and from this we get
the required result.
We consider first the three base cases, that L = ∅, L = {e} or L = {a} for
some letter a ∈ X. If L = ∅ then LE can only be the empty set too. If L =
{e}, then LE is either {e} if e ∈ E or the empty set again if not. If L = {a}
for a letter a, then LE is one of the following:
{e, a}, if a, e ∈ E;
{e}, if a ∈ E and e 6∈ E;
{a}, if e ∈ E and a ∈6 E;
∅, if a, e 6∈ E.
From this it will follow, as in the case for unions, that there are no more
than mn distinct languages of the form (L1 L2 )E (with m and n as before).
For the first direction of (*) let v ∈ (L1 L2 )E . Then wv = w1 w2 for some
w1 ∈ L1 , w2 ∈ L2 and w ∈ E. There are two cases possible, depending on
the length of w:
For our final step we consider languages L∗ , assuming that there are only
finitely many different languages of the form LE . We show that (L∗ )E is
either LEL∗ L∗ , if the empty word e is not in E, or the union of this set with
{e}, if e ∈ E.
Now we note that there are at most twice as many different languages of the
form (L∗ )E as there are of the form LE , and hence only finitely many.
This result can now be used to finish our proof of Kleene’s Theorem.
Proof: Since one direction of this was proved in Theorem 8.2.1, we now
have to prove that any regular language L is recognizable. By Lemma 8.2.3,
there are only finitely many different languages of the form Lw ; we will label
these as L1 , . . ., Lk . We construct our new automaton H with states indexed
by these Li . We have a transition δ(Li , a) = Lj if Lj = (Li )a . This gives a
complete (deterministic) definition of our transition function δ. Our initial
state is that indexed by L itself, and the final states are those Li for which
the language Li contains the empty word e. We now claim that the language
accepted by this automaton H is precisely our given language L. This is
because for any word w, Lw is the language which labels the state δ̂(z0 , w).
Therefore
(i) λ̂(z, e) = z,
(ii) λ̂(z, x) = λ(z, x) for any letter x, and
(iii) λ̂(z, xw) = λ(z, x)λ̂(δ̂(z, x), w).
δ 0 1 λ 0 1
z0 z0 z1 z0 0 1
z1 z2 z3 z1 2 3
z2 z4 z0 z2 4 0
z3 z1 z2 z3 1 2
z4 z3 z4 z4 3 4
The reader should draw the graph of this automaton. We will illustrate the
behaviour of the extended output function λ̂ by calculating the output of
this automaton when it is started in state z0 with input word w = 110. We
have
Other concepts regarding subalgebras from Section 1.3 can similarly be gen-
eralized to the multi-based setting. Let H = (Z, X, B; δ, λ) be a finite deter-
ministic automaton with output, and let {Hj = (Zj , Xj , Bj ; δj , λj ) | j ∈ J}
be an indexed family of subautomata of H. Assume that the intersections
T T T
Zj , Xj and Bj are non-empty. Then we define a new subautoma-
j∈J j∈J j∈J
Hj , defined by
T
ton called the intersection
j∈J
Bj ; δ 0 , λ0 ),
\ \ \ \
Hj := ( Zj , Xj ,
j∈J j∈J j∈J j∈J
{H0 = (Z 0 , X 0 , B 0 ; δ 0 , λ0 ) | Z0 ⊆ Z 0 , X0 ⊆ X 0 , B0 ⊆ B 0 },
\
fS (δ1 (z, x)) = δ2 (fS (z), fI (x)) and fO (λ1 (z, x)) = λ2 (fS (z), fI (x)).
The reader should check that our congruence property means precisely that
these two mappings are well defined. As before, congruences occur as kernels
of homomorphisms.
f
H - H0
@ 6
@
@ (=)
natkerf @ h
@
@
@
@
R
@
H/kerf
All of these results for automata with output can be extended to the varia-
tions of initial and weak initial automata. In the initial case, we require that
a homomorphism f : H1 → H2 satisfy the additional requirement that fS
maps the initial state of H1 to the initial state of H2 , and maps final states
of H1 to final states of H2 .
In the case that H1 and H2 are initial, so Z0i = {z0i }, this condition reduces
to H1 ∼ H2 iff z01 ∼ z02 . We also leave it as an exercise for the reader to show
8.3. ALGEBRAIC OPERATIONS ON FINITE AUTOMATA 163
Z 0 = {[z]∼ | z ∈ Z}.
We show first that these functions are well defined, then check that H0 does
recognize the same language as H. Suppose that ([z1 ]∼ , x1 ) = ([z2 ]∼ , x2 ), so
that x1 = x2 and z1 ∼ z2 . By the definition of equivalent states this means
that for any input w ∈ X ∗ we have
By (∗) with w = x, the first part in each of these concatenations is the same;
so the remaining parts must also be equal:
The equivalence of two automata was defined only in terms of their behaviour
on the same input words. But such equivalence also turns out to have an
algebraic interpretation.
A = (A, Σ, X, α, A0 ),
where
A := (A; ΣA ) is a finite algebra,
Σ is a ranked alphabet of operation symbols,
X is a set of individual variables,
α : X ∪ Σ0 −→ A ∪ ΣA 0 is a mapping, called the evaluation mapping, and
A0 is a subset of the (finite) set A.
166 CHAPTER 8. ALGEBRAIC MACHINES
Example 8.4.2 We take Σ = Σ1 ∪Σ2 , where Σ1 = {h} and Σ2 = {f, g}, and
alphabet X = {x1 , x2 }. We consider the finite algebra A = ({0, 1}; ∧, ∨, ¬),
where hA = ¬, f A = ∧ and g A = ∨. We define an evaluation α by
α(x1 ) = 1 and α(x2 ) = 0. We also designate A0 = {1}. Consider the term
t = f (h(f (x2 , x1 )), g(h(x2 ), x1 )). The tree corresponding to this term is ac-
8.4. TREE RECOGNIZERS 167
Let P(A) be the power set of the set A. For n ≥ 1, an n-ary mapping
f A : An −→ P(A)
f A : {∅} −→ P(A).
168 CHAPTER 8. ALGEBRAIC MACHINES
f P(A) (A1 , . . . , An ) :=
[
{f A (a1 , . . . , an ) | a1 ∈ A1 , . . . , an ∈ An },
A = (A, Σ, X, α, A0 ),
where
A:= (A; ΣA ) is a finite non-deterministic algebra,
Σ is a ranked alphabet of operation symbols,
X is a set of individual variables,
α : X ∪ Σ0 −→ P(A) ∪ ΣA 0 is a mapping (called the evaluation mapping),
and
A0 is a subset of the (finite) set A.
The set
T (A) := {t | t ∈ WΣ (X) and α̂(t) ∩ A0 6= ∅}
is called the language recognized by A.
It turns out that deterministic and non-deterministic tree-recognizers are
equivalent, in the sense that the families of languages which they recognize
are equal. This can be useful, since non-deterministic tree-recognizers are
sometimes easier to work with than deterministic ones.
G = (N, Σ, X, P, a0 ),
where
N is a finite non-empty set, called the set of non-terminal symbols,
Σ is a ranked alphabet of operation symbols,
X is an alphabet of individual variables,
P is a finite set of productions (or rules of derivation) which have the form
a → r for some a ∈ N and r ∈ WΣ (N ∪ X), and a0 ∈ N is called an initial
symbol.
We also specify that N ∩ (Σ ∪ X) = ∅.
Let G be a regular Σ − X-tree grammar. We think of a production a → r
from G as a rule allowing us to replace the non-terminal symbol a by the
term (tree) r. Such a replacement is done within terms, as follows. Let s be
a term from WΣ (X ∪ N ) in which the non-terminal symbol a occurs as a
subterm. If we have a production a → r in G, we can change s to a new
term t by replacing the subterm a by the term r. In this case we write
s ⇒G t,
We write
s ⇒∗G t
if either s = t or there is a sequence (t0 , . . . , tn ) of terms from WΣ (X ∪ N )
with t0 = s, tn = t and
t0 ⇒G t1 ⇒G · · · ⇒G tn−1 ⇒G tn .
Then the tree f (h, f (x, f (x, f (x, x)))) has the derivation
a ⇒G f (h, a) ⇒G f (h, f (x, f (x, b))) ⇒G f (h, f (x, f (x, f (x, x)))),
and the tree f (x, f (x, f (x, x))) has the derivation
Next let a → r be a production with depth greater than one. Then r has the
form fi (r1 , . . . , rni ), where ni ≥ 1, fi ∈ Σni and depth(rj ) < depth(r) for all
j = 1, . . . , ni . In this case we delete a → r, and replace it with productions
of the form
ai → rj , (∗∗)
for each 1 ≤ j ≤ ni .
In this way, step by step, any production with a depth greater than 1 can be
replaced by productions of lower depth, until we reach a normal form. None
of these steps changes the language generated by the grammar; so we obtain
an equivalent normal-form grammar.
These productions have the form required to ensure that the grammar G
is in normal form. Conversely, given any grammar G, we may construct an
equivalent normal form grammar and then use the productions to define a
8.5. REGULAR TREE GRAMMARS 173
We proceed by induction on the depth of the tree t. For the base case, if
t has depth zero it is either a variable x or a nullary operation symbol f .
If t = x ∈ X and a ∈ α̂(x) then a ⇒∗G x since a → x is a production. If
conversely a ⇒∗G x then to start this derivation we can only use a production
of the form a → x. But then a ∈ α(x). A similar argument holds in the case
that t = f for some f ∈ Σ0 , using the second kind of productions in P .
For the inductive step, let t = fi (t1 , . . . , tni ) and suppose that
This proves that the equivalence (*) holds. Now for every tree t we have
t ∈ T (A) ⇔ α̂(t) ∩ A0 6= ∅ ⇔ ∃a ∈ A0 (a ∈ α̂(t)) ⇔ ∃a ∈ A0 (a ⇒∗G t) ⇔
t ∈ T (G).
C1 = (C, Σ, X, γ, A0 × B 0 ),
C2 = (C, Σ, X, γ, A0 × B ∪ A × B 0 ), and
C3 = (C, Σ, X, γ, A0 × (B \ B 0 )),
Now we consider mappings which transform trees from one language into
trees from another one. Let Σ := {fi | i ∈ I} be a set of operation symbols
of type τ1 = (ni )i∈I , where fi is ni -ary, ni ∈ IN and let Ω = {gj | j ∈ J}
be a set of operation symbols of type τ2 = (nj )j∈J where gj is nj -ary. We
denote by WΣ (X) and by WΩ (Y ) the sets of all terms of type τ1 and τ2 ,
respectively, where X and Y are alphabets of variables.
(ii) h(fi (t1 , . . . , tni )) = hni (fi )(ξ1 ← h(t1 ), . . . , ξni ← h(tni )), where ξj ←
h(tj ), means to substitute h(tj ) for ξj .
P 0 := {a → h0 (t) | a → t ∈ P }.
The theorem will be proved if we show that T (G0 ) = h(T ). This in turn can
be proved by showing that a ⇒∗G0 t iff (∃s ∈ WΣ (X)) (h(s) = t and a ⇒∗G s
for all a ∈ N , t ∈ WΩ (Y )). We leave the remaining details to the reader (see
for instance F. Gécseg and M. Steinby, [48]).
(iii) ϕ−1 (B 0 ) = A0 .
8.7. MINIMAL TREE RECOGNIZERS 177
Notice that by definition the set C(A) is a subset of the set ConA of all
congruences on the algebra A.
{% | % ∈ C(A)} ∈ C(A).
S
(iii)
C(A)
The details of this proof are left to the reader; see Exercise 8.11.5.
The join in (iii) is the greatest and thus the generating element of C(A).
Later on we will give a more useful description of the greatest element of
C(A).
Using the natural homomorphism and Lemma 8.7.2 we have the following
conclusion.
(i) reduced if ∼A = ∆A ;
(ii) connected if every state is reachable, in the sense that for every a ∈ A
there exists a tree t ∈ WΣ (X) such that α̂(t) = a;
The quotient recognizer A/∼A is called the reduced form of A. Tree recog-
nizers which have isomorphic reduced forms are equivalent. For connected
tree-recognizers the converse is also true.
2. We show that ϕ is well defined. Assume that α̂(s) = α̂(t) and that β̂(s) 6=
β̂(t). Then β̂(s) and β̂(t) are non-equivalent since A and B are reduced.
Then there exists a unary polynomial operation f of the algebra A such
that f (β̂(s)) ∈ B 0 and f (β̂(t)) 6∈ B 0 , or vice versa. Furthermore, there is a
tree p ∈ WΣ (B ∪ {ξ}), where ξ is an auxiliary variable and ξ 6∈ B ∪ X, such
that for all b ∈ B we have f (b) = pB (βb ) where βb : B ∪ {ξ} → B is defined
by βb |B = 1B and βb (ξ) = b. Since B is connected, for each b ∈ B there
exists a Σ − X-tree pb such that β̂[pb ] = b. Now let
q = p(b ← pb | b ∈ B) (∈ WΣ (X ∪ {ξ})).
and
β̂(qt ) = pB (ββ̂(t) ) = f (β̂(t)) 6∈ B 0 .
8.7. MINIMAL TREE RECOGNIZERS 181
We remark that the previous results can be used to prove that an arbitrary
Σ − X-language T is recognizable iff there exist a finite Σ-algebra A, a
homomorphism ϕ : FΣ (X) → A and a subset A0 ⊆ A such that ϕ−1 (A) =
T . This proposition allows us to give a new definition of recognizability for
subsets of the universes of arbitrary algebras, not just free algebras. Let A
be any algebra. Then a subset T ⊆ A is called recognizable if there exist a
finite algebra B of the same type, a homomorphism ϕ : A → B and a subset
B 0 ⊆ B such that ϕ−1 (B 0 ) = T . This definition gives us the recognizability of
Σ − X-languages when A is the free algebra of its type, and the recognizable
languages of Section 8.1 when A is the free monoid X ∗ generated by X.
8.6 we denote by WΣ (X) and WΩ (X) the sets of all terms of type τ1 and τ2 ,
respectively. Then we define a tree transformation as a binary relation
The most important tree transformations are those which can be given in
an effective way.
This means that tree transformations of the form TA , induced from a tree
transducer, can be described in an effective (algorithmic) way. Now we con-
sider the following example of a tree transducer.
Example 8.8.3 Let A = (Σ, {x}, {a0 , a1 }, Ω, P, {a1 }) be the tree trans-
ducer with Σ = Σ2 = {f }, Ω = Ω2 = {g}, and where P consists of the
productions x → a0 x, f (a0 , a0 ) → a1 g(ξ1 , ξ2 ), f (a0 , a1 ) → a0 g(ξ1 , ξ2 ),
f (a1 , a0 ) → a1 g(ξ1 , ξ2 ) and f (a1 , a1 ) → a1 g(ξ1 , ξ2 ).
Then the term t = f (f (x, x), x) has the following derivation:
f (f (x, x), x) ⇒∗A f (f (a0 x, a0 x), a0 x) ⇒∗A f (a1 g(x, x), a0 x) ⇒∗A
a1 g(g(x, x), x).
P = {x → ahX (x) | x ∈ X} ∪
{fi (a(ξ1 ), . . . , a(ξni )) → ahni (fi )(ξ1 , . . . , ξni ) | fi ∈ Σni },
= h(x) and therefore (x, h(x)) ∈ TA . Inductively, let t = fi (t1 , . . . , tni ) and
suppose that (tj , h(tj )) ∈ Th implies tj ⇒∗A ah(tj ) for all j = 1, . . . , nj . Then
t has the following derivation in A:
The state µ1 is called the initial state, and we set Z1 = {µ1 }, while the state
µ0 is called the final state.
L(T ) := {w ∈ X ∗ | w is accepted by T }.
(i) For each 1 ≤ i ≤ k and for each r ∈ {0, 1}, T contains exactly one
instruction µi rsT γ.
(ii) No state symbol µj for which j > k occurs in the instructions in T .
A configuration of a Turing machine T is a triple Q = (t, n, γ), where t is a
tape, n is an integer and γ is a state symbol. This encodes the information
that the machine is in state γ, reading the n-th square on tape t. If there is
an instruction γt(n)sT γ 0 , then the machine will move into state γ 0 , and the
read-write head moves either right or left according to T . In addition, the
tape t is converted to a tape t0 , for which t0 (n) = s and t0 (k) = t(k) for all
k 6= n. The result is described by a new configuration, either (t0 , n − 1, γ 0 )
8.10. UNDECIDABLE PROBLEMS 187
In this way, starting with Q0 and applying the transition to new configura-
tions, we obtain a sequence Q0 , Q1 , Q2 , . . . of configurations. This process
will only stop if we reach a configuration Qm which uses the final state µ0 ,
since there is no instruction for this state. In this case the sequence of con-
figurations stops at some finite stage m; otherwise it will continue infinitely.
We say that the Turing machine T halts iff its sequence of configurations is
finite. In the next section we will use the well-known fact that the problem of
deciding whether a Turing machine will halt, the so-called Halting Problem,
is recursively undecidable. More details may be found in S. C. Kleene [63].
The main idea of McKenzie’s proofs involves the construction of a finite al-
gebra A(T ) which encodes the computation of a Turing machine T . We will
outline here the construction of this algebra, and its use in the proof, but
the full proofs are beyond the scope of this book. Complete details may be
found in [79], [80] and [81].
U = {1, 2, H},
W = {C, D, C, D},
A = {0} ∪ U ∪ W ,
k
Vi , where Vi = Vi0 ∪ Vi1 and Vir = Vir0 ∪Virs ,
S
V =
i=0
with Virs s , D s , M r , C s , D s , M r }, for 0 ≤ i ≤ k and {r, s} ⊆ {0, 1}.
= {Cir ir i ir ir i
The unbarred symbols are used to encode configurations of the Turing ma-
chine, while the barred versions control the finite subdirectly irreducible
algebras. The universe of A(T ) is defined to be the set A ∪ V . We point
out that the cardinality of this set is 20k + 28, where k is the number of
non-halting states of the Turing machine.
On this base set, we define the following operations. The first is a semilattice
operation ∧, defined by
(
x, if x = y
x∧y =
0, otherwise.
2 · D = H · C = D, 1 · C = C,
2 · D = H · C = D, 1 · C = C,
x · y = 0 for all other pairs (x, y).
x,
if x = y 6= 0,
J(x, y, z) = x ∧ z, if x = y ∈ V ∪ W,
0, otherwise.
x ∧ z,
if x = y 6= 0,
0
J (x, y, z) = x, if x = y ∈ V ∪ W,
0, otherwise.
8.10. UNDECIDABLE PROBLEMS 189
(
(x ∧ y) ∨ (x ∧ z), if u ∈ V0 ,
S0 (u, v, x, y, z) =
0, otherwise.
(
(x ∧ y) ∨ (x ∧ z), if u ∈ {1, 2},
S1 (u, v, x, y, z) =
0, otherwise.
(
(x ∧ y) ∨ (x ∧ z), if u = v ∈ V ∪ W,
S2 (u, v, x, y, z) =
0, otherwise.
x · y,
if x · y = z · u 6= 0 and x = z and y = w,
T (x, y, z, u) = x · y, if x · y = z · u 6= 0 and x 6= z or y =
6 w,
0, otherwise.
For each instruction µi rsLµj of the Turing machine T and each t ∈ {0, 1},
we have an operation Lirt which will describe the operation of the machine
when it is reading symbol r in state µi , and the square just to the left of the
current one has symbol t on it.
s0 , 0
s for some s0 ,
Cjt if x = y = 1, u = Cir
Mjt , t ,
if x = H, y = 1, u = Cir
s if x = 2, y = H, u = Mir ,
Djt ,
0 s0 for some s0 ,
Lirt (x, y, u) =
Ds ,
jt if x = y = 2, u = Dir
v, if u ∈ V and Lirt (x, y, u) = v ∈ V,
according to the previous lines
0, otherwise.
For each instruction of the form µi rsRµj in T and for each t ∈ {0, 1}, we
have an operation Rirt , given by
190 CHAPTER 8. ALGEBRAIC MACHINES
s0 , 0
s for some s0 ,
Cjt if x = y = 1, u = Cir
s, if x = H, y = 1, u = Mit ,
Cjt
t if x = 2, y = H, u = Ditr ,
Mj ,
0 s0 for some s0 ,
Rirt (x, y, u) =
Ds ,
jt if x = y = 2, u = Dir
v, if u ∈ V and Rirt (x, y, u) = v ∈ V,
according to the previous lines
0, otherwise.
Let L be the collection of all these operations Lirt , and dually for R. We will
assume that F1 , . . . , Fc is a complete list of all these operations. We define
a binary relation on U by
Now we use this to define operations Ui1 and Ui2 , for each 1 ≤ i ≤ c, by
Fi (x, y, u),
if x z, y 6= z, Fi (x, y, u) 6= 0,
Ui1 (x, y, z, u) = Fi (x, y, u) if x z, y = z, Fi (x, y, u) 6= 0,
0, otherwise.
Fi (y, z, u),
if x z, x 6= y, Fi (y, z, u) 6= 0,
Ui2 (x, y, z, u) = Fi (y, z, u) if x z, x = y, Fi (y, z, u) 6= 0,
0, otherwise.
Now the two possibilities for T , that it halts or not, are considered. McKen-
zie showed that if T does not halt, the variety generated by A(T ) contains a
denumerably infinite subdirectly irreducible algebra, and the residual bound
of A(T ) satisfies κ(A(T )) ≥ ω1 . However, if T halts, then A(T ) is residu-
ally finite, with a finite cardinal m such that κ(A(T )) ≤ m. Since it is not
8.11. EXERCISES 191
Now we can apply Willard’s theorem, Theorem 6.7.5. The variety V (A(T ))
can be shown to be congruence meet-semidistributive, and if T halts, the
variety is also residually finite. Therefore, if T halts, the algebra A(T ) is
finitely axiomatizable. But it is algorithmically undecidable if T halts; so it
is also undecidable whether A(T ) is finitely axiomatizable or not.
8.11 Exercises
8.11.1. Draw a directed graph illustrating the action of the automaton from
Example 8.2.5.
8.11.4. Prove that for any weakly initial deterministic finite automaton H,
the reduced automaton H0 constructed in the proof of Theorem 8.3.10 is
unique up to isomorphism.
8.11.9. Let Xn be a fixed finite alphabet. Let R be the relation between the
set of all languages on the alphabet Xn and the set of all finite automata on
Xn , defined by (L, H) ∈ R iff H recognizes L. Describe the Galois-connection
induced by this relation R.
Chapter 9
Mal’cev-Type Conditions
193
194 CHAPTER 9. MAL’CEV-TYPE CONDITIONS
It is easy to show that for any two equivalence relations θ and ψ defined on
a set A, the union θ ∪ ψ is again an equivalence relation defined on A iff
θ and ψ are permutable. In this case θ ◦ ψ is the least equivalence relation
containing θ and ψ, and we have θ∪ψ = θ◦ψ. When θ and ψ are congruences
on an algebra A, this makes the join θ ∨ ψ in the congruence lattice equal to
θ ◦ ψ. (We have used this argument already in the proof of Theorem 4.1.4.)
This means that in FV ({x, y}) we have x = p(x, y, y), and hence that
p(x, y, y) ≈ x ∈ Id V . A similar argument shows that p(x, x, y) ≈ y is also
in Id V .
For the converse, assume now that there is a term p such that p(x, x, y) ≈
y, p(x, y, y) ≈ x ∈ Id V . To show that V is congruence permutable, we
let A be any algebra in V , let θ and ψ be any congruences on A, and let
(a, b) ∈ ψ ◦ θ. By definition there is an element c ∈ A such that (a, c) ∈ θ
and (c, b) ∈ ψ. Using (a, a), (b, c), (b, b) ∈ ψ and (a, a), (c, a), (b, b) ∈ θ,
and the compatibility property of terms from Theorem 5.2.4, we have
Example 9.1.2 Consider the class of all groups, viewed as algebras of type
(2,1,0). This is a variety, defined equationally by the identities
(As usual, we omit the binary multiplication symbol.) This variety has a
Mal’cev term p, given by p(x, y, z) := xy −1 z. Clearly for this term we have
p(x, x, y) = xx−1 y ≈ ey ≈ y and p(x, y, y) = xy −1 y ≈ xe ≈ x, in any group.
Theorem 9.1.1 then tells us that the variety of all groups is congruence
permutable. Similarly, the class of all rings forms a variety with a Mal’cev
term given by p(x, y, z) = x − y + z, since in any ring p(x, x, y) = x − x + y
≈ y and p(x, y, y) = x − y + y ≈ x. So the variety of rings is also congruence
permutable.
Example 9.2.4 We will use Theorem 9.2.3 to show that the variety of all
lattices is congruence distributive. Let m be the ternary lattice term
m(x, y, z) = (x ∨ y) ∧ (x ∨ z) ∧ (y ∨ z).
Then in any lattice, using the idempotency and absorption laws, we have
m(x, x, y) = (x ∨ x) ∧ (x ∨ y) ∧ (x ∨ y) ≈ x ∧ (x ∨ y) ≈ x,
m(x, y, x) = (x ∨ y) ∧ (x ∨ x) ∧ (y ∨ x) ≈ (x ∨ y) ∧ x ≈ x, and
m(y, x, x) = (y ∨ x) ∧ (y ∨ x) ∧ (x ∨ x) ≈ (y ∨ x) ∧ x ≈ x.
Thus m is a two-thirds majority term, and the variety of all lattices is con-
gruence distributive.
When the conditions (∆n ) from this theorem are satisfied for a variety V ,
the variety is said to be n-distributive. It is possible to show that an n1 -
distributive variety is also n2 -distributive for any n2 ≥ n1 . But the converse
direction is false, since n2 -distributivity does not imply n1 -distributivity (see
K. Fichtner, [44]).
(θ(x, y) ∧ θ(x, z)) ∨ (θ(y, z) ∧ θ(x, z)) = (θ(x, y) ∨ θ(y, z)) ∧ θ(x, z). (D)
Since (x, y) and (y, z) are both in θ(x, y) ∨ θ(y, z), and this relation is a
congruence, we have (x, z) ∈ θ(x, y) ∨ θ(y, z). Hence the pair (x, z) is in
the relation on the right hand side of the equation (D) above, and because
of the equality must be in the relation on the left hand side too. By the
join criterion from Lemma 9.2.2, there are a natural number n and finitely
many elements d0 , . . . , dn of FV ({x, y, z}) such that d0 = x, dn = z, and
(d0 , d1 ) ∈ θ(y, z), (d1 , d2 ) ∈ θ(x, y), and so on, with the last pair (dn−1 , dn )
in either θ(x, y) or θ(y, z), depending on whether the number n is even or
odd.
9.2. CONGRUENCE DISTRIBUTIVITY 199
Assume now that all the identities (∆n ) are satisfied for some number n
and some terms t0 , . . ., tn . Let A ∈ V and let θ, ψ and Φ be congruences
on A. Since we always have (θ ∧ Φ) ∨ (θ ∧ ψ) contained in θ ∧ (Φ ∨ ψ),
it will suffice to prove the converse inclusion. We let (x, y) ∈ θ ∧ (Φ ∨ ψ).
Then (x, y) ∈ θ, and from Lemma 9.2.2 there are a number m and elements
d0 , . . . , dm ∈ A with x = d0 , (d0 , d1 ) ∈ ψ, (d1 , d2 ) ∈ Φ, . . ., (dm−1 , dm ) ∈ Φ
and dm = y. For each j = 1, . . . , n − 1, applying the term tj to these
pairs and the pairs (x, x) and (y, y) gives us (tj (x, x, y), tj (x, d1 , y)) ∈ ψ,
(tj (x, d1 , y), tj (x, d2 , y)) ∈ Φ, . . ., (tj (x, dm−1 , y), tj (x, y, y)) ∈ Φ. We
also have each pair (tj (x, dk , y), tj (x, dk+1 , x)) ∈ θ, by the same argu-
ment as in the proof of Theorem 9.2.3. Combining these facts we have
(tj (x, x, y), tj (x, d1 , y)) ∈ θ ∧ ψ, (tj (x, d1 , y), tj (x, d2 , y)) ∈ θ ∧ Φ, . . .,
(tj (x, dm−1 , y), tj (x, y, y)) ∈ θ ∧ Φ. Hence by Lemma 9.2.2 we have
(tj (x, x, y), tj (x, y, y)) ∈ (θ ∧ Φ) ∨ (θ ∧ ψ). Altogether we have
The next famous theorem of Baker and Pixley ([6]) shows that the presence
of a majority term operation has some far-reaching consequences.
We will give an inductive proof that for all k ≥ 2, any set D ⊆ An of size k
has this property. Since A is finite, this will suffice.
When k = 2 we can write D = {(x1 , . . . , xn ), (y1 , . . . , yn )}.
Since by assumption f preserves every subalgebra of A2 , the pair
(f (x1 , . . . , xn ), f (y1 , . . . , yn )) belongs to the subalgebra of A2 which is gen-
erated by {(x1 , y1 ), . . . , (xn , yn )}. Then by Theorem 1.3.8, there is a term
operation p of A for which f (x1 , . . . , xn ) = p(x1 , . . . , xn ) and f (y1 , . . . , yn ) =
p(y1 , . . . , yn ). Let J be the collection of all sets of n-tuples such that
f (x1 , . . . , xn ) = p(x1 , . . . , xn ). Thus D belongs to J.
Assume now that D has size k+1 ≥ 3 and that (x1 , . . . , xn ), (y1 , . . . , yn ), (z1 ,
. . . , zn ) are pairwise different elements of D. By induction f agrees on
D1 := D \ {(x1 , . . . , xn )} with some term operation p1 of A, f agrees on
D2 := D \ {(y1 , . . . , yn )} with some term operation p2 of A, and f agrees on
D3 := D \ {(z1 , . . . , zn )} with some term operation p3 of A.
Example 9.3.3 In this example we will show that the variety of all
Boolean algebras is arithmetical. It can be shown that the variety of
all Boolean algebras is generated by the two-element Boolean algebra
2B = ({0, 1}; ∧, ∨, ¬, ⇒, 0, 1), where the operation symbols here denote the
usual Boolean operations on the set {0, 1}. This means that every iden-
tity which is satisfied in the two-element Boolean algebra 2B is satisfied in
any Boolean algebra. To prove arithmeticity by Theorem 9.3.2 (iii), we look
for a ternary term q(x, y, z) on the set {0, 1} which satisfies the identities
q(x, y, y) ≈ q(x, y, x) ≈ q(y, y, x) ≈ x in 2B . Considering the truth table of
such an operation, as shown below, we see that there is exactly one such
ternary operation q. This operation q is indeed a term operation of 2B , since
it can be expressed as q(x, y, z) = (¬x)(¬y)z ∨ x(¬y)(¬z) ∨ x(¬y)z ∨ xyz.
Thus the variety is arithmetical.
9.4. N -MODULARITY AND N -PERMUTABILITY 203
x 0 0 0 0 1 1 1 1
y 0 0 1 1 0 0 1 1
z 0 1 0 1 0 1 0 1
q(x, y, z) 0 1 0 0 1 1 0 1
θ0 ⊇ θ2 ⇒ θ0 ∩ (θ1 ∪ θ2 ) ⊆ (θ0 ∩ θ1 ) ∪ θ2 .
g0 (x, y, z) ≈ x
gi (x, y, x) ≈ x for all 0 ≤ i ≤ n,
gi (x, x, y) ≈ gi+1 (x, x, y) if i is even,
gi (x, y, y) ≈ gi+1 (x, y, y) if i is odd,
gi (x, y, y) ≈ d(x, y, y)
d(x, x, y) ≈ y.
We noted in Section 9.1 that if two congruences θ and ψ are permutable, then
θ ∪ ψ = θ ◦ ψ. In this case we say that the congruences are 2-permutable. We
can generalize this, to say that two congruences θ and ψ are n-permutable,
for n ≥ 2, if
(
θ if i is even
θ ∪ ψ = θ1 ◦ . . . ◦ θn where θi = .
ψ if i is odd
p0 (x, y, z) ≈ x,
pi (x, x, y) ≈ pi+1 (x, y, y) for all 0 ≤ i < n, (Pn )
pn (x, y, z) ≈ z.
In [21] K. Denecke proved the following result for the special case of varieties
generated by two-element algebras.
algebras was given by E. L. Post ([95]) in 1941. We shall study Post’s lattice
of representatives in the next chapter, making use of the congruence prop-
erties we determine here.
x 0 0 0 0 1 1 1 1
y 0 0 1 1 0 0 1 1
z 0 1 0 1 0 1 0 1
t1 (x, y, z) 0 0 0 1 0 1 1 1
t1 (x, x, y) ≈ t1 (x, y, x) ≈ x
t2 (x, x, y) ≈ y, t2 (x, y, x) ≈ x.
Again we turn to tables of values to see what choices are possible for t1 and
t2 . Note that the identities above determine the value of each of t1 and t2
on six of the eight triples in A3 , leaving us to consider possible values on the
two remaining triples (0, 1, 1) and (1, 0, 0).
x 0 0 0 0 1 1 1 1
y 0 0 1 1 0 0 1 1
z 0 1 0 1 0 1 0 1
t1 (x, y, z) 0 0 0 ? ? 1 1 1
t2 (x, y, z) 0 1 0 ? ? 1 0 1
(x ∧ ¬y). But using t2 we can define a term m(x, y, z) = t2 (x, t2 (x, y, z), z),
which we can easily see acts as a majority term:
Thus in this case our algebra is in fact 2-distributive, by the previous Lemma,
and so 3-distributive as well.
For 4-distributivity, we can start with the same argument for t1 as in the 3-
distributive case. If a two-element algebra has terms t0 , . . ., t4 which satisfy
conditions (∆4 ), then in particular the term t1 has its value constrained on
six of the eight triples, and we have 4 possibilities for t1 , as we did in the
3-distributive case.
Case 2: t1 = x. Here we must consider the values possible for the remaining
terms t2 and t3 (with t4 = z). Making a table of values possible for these
two, and using the 3-distributive identities forcing values of t3 , shows that
their values are also determined on six triples, and must be equal on the
remaining two triples. This leads to four subcases:
Combining our analysis of all the cases for n-distributivity, we have the
following conclusion.
We will show that in the three remaining cases for p1 , we can use the exis-
tence of the term p1 to produce a Mal’cev term p.
(i) A is 2-permutable, or
9.7 Exercises
9.7.1. Prove Corollary 9.5.4.
9.7.2. Prove the assertion following the statement of Theorem 9.2.4, that
when V has a majority term m, we may take n = 2, and terms t0 = x1 , t1
= m and t2 = x3 , to obtain the identities (∆2 ) of Jonsson’s Condition.
9.7.3. Prove that for any two-element algebra A, the variety V (A) generated
by A is congruence modular iff it is congruence distributive or congruence
permutable.
215
216 CHAPTER 10. CLONES AND COMPLETENESS
operations on our set, one binary, three unary, and one nullary. Subalgebras
of this algebra are then called iterative algebras. The binary operation ∗ is
clearly associative. This makes any iterative algebra a semigroup, with three
additional unary operations and a nullary operation. It is not hard to prove
that universes of subalgebras of O(A) are clones in our sense: they are closed
under composition and contain all the projections. Conversely, any clone of
operations can be shown to be closed under the iterative algebra operations
∗, ζ, τ , ∆ and e21 . Thus any clone occurs as the base set of a subalgebra of
a full iterative algebra. Universes of clones are also called closed classes or
superposition-closed classes, and we refer to superposition of functions.
for the (heterogeneous) full clone of all operations defined on the set A. A
clone on A is then any subalgebra of this algebra. Each such clone belongs
to a variety K0 of heterogeneous algebras defined by the following identities
(C1), (C2), (C3):
p (z, S n (y , x , . . . , x ), . . . , S n (y , x , . . . , x )) ≈
(C1) Sm m 1 1 n m p 1 n
n (S p (z, y , . . . , y ), x , . . . , x ),
Sm n 1 p 1 n (m, n, p ∈ IN+ ),
n (en , x , . . . , x ) ≈ x ,
(C2) Sm i 1 n i (m ∈ IN+ , 1 ≤ i ≤ n),
For any set F ⊆ O(A) of operations and any set Q ⊆ R(A) of relations we
consider
(For convenience we usually write P olA % and InvA f for P olA {%} and
InvA {f }, and if the base set is clear from the context we usually omit the
subscript A.)
Next we consider the relation ρ := {(0, 0), (0, 1), (1, 1)}, which can be defined
by (x, y) ∈ ρ iff x ≤ y. A 2×n matrix has all columns in ρ whenever y1j ≤ y2j
for all j = 1, . . . , n and thus f preserves ρ if
f (y11 , . . . , y1n ) ≤ f (y21 , . . . , y2n ) provided y11 ≤ y21 , . . . , y1n ≤ y2n .
10.3. THE LATTICE OF ALL BOOLEAN CLONES 219
In this section we will give a proof using our results from Section 9.6 on the
properties of the congruence lattices of two-element algebras. Clearly, if all
two-element algebras are known, then all closed classes of Boolean operations
are known. For the set A = {0, 1}, we will give a complete classification
of the varieties V (A) generated by algebras with base set A. Our analysis
will be broken into a number of cases, beginning with whether V (A) is
congruence modular or not. In the case that V (A) is congruence modular,
we will consider two further cases, depending on whether V (A) is congruence
220 CHAPTER 10. CLONES AND COMPLETENESS
distributive or not. The case that V (A) is congruence distributive will again
be broken into cases, based on whether V (A) is 2-distributive, 3-distributive
but not 2-distributive or finally 4-distributive but not 3-distributive. Note
that as in Section 9.6 we do not distinguish between isomorphic or equivalent
(having the same clone of term operations) algebras. We will use the notation
cloneA for the term clone, or clone of all term operations, of A. We also use
the standard notation for Boolean operations, as introduced in Section 9.6.
In addition, we will denote by c20 and c21 the binary constant operations with
values 0 and 1, respectively.
Next we claim that any at least essentially ternary operation g from cloneA
must be monotone. Monotonicity means that if we define (a1 , a2 , . . . , an )
(b1 , b2 , . . . , bn ) iff ai ≤ bi for all i ∈ {1, . . . , n}, for ai , bi ∈ {0, 1}, then
for arbitrary n-tuples (a1 , a2 , . . . , an ) and (b1 , b2 , . . . , bn ), if (a1 , . . . , an )
(b1 , . . . , bn ) then g(a1 , . . . , an ) ≤ g(b1 , . . . , bn ). Let g be an essentially at
least ternary operation. If g was not monotone, there would be a pair of
n-tuples
a = (a1 , a2 , . . . , ai−1 , 0, ai+1 , . . . , an )
b = (a1 , a2 , . . . , ai−1 , 1, ai+1 , . . . , an )
for which g(a) = 1 > 0 = g(b). We can assume that g is not constant, and
therefore that g(0, . . . , 0) = 0 and g(1, . . . , 1) = 1. But then by identifica-
tion of variables we can make a ternary operation f with f (0, 0, 0) = 0,
f (1, 0, 1) = 0, f (0, 0, 1) = 1 and f (1, 1, 1) = 1. Consideration of all possible
values for f on the remaining four triples in A3 shows that we would get
one of the operations x + y + z, (x ∨ ¬ y) ∧ z or (x ∧ ¬ y) ∨ z. But these are
impossible, as we saw above. Hence g must be monotone.
Now we will show that the n-ary term operations of cloneA are n-ary pro-
jections, negations of n-ary projections, n-ary constant operations, or can
be written in the form f (x1 , . . . , xn ) = xi1 ∧ . . . ∧ xin with {i1 , . . . , in } ⊆
{1, . . . , n}. For the binary term operations this is clear. For n > 2, we have
seen that any essentially n-ary term operation h is monotone. Suppose that
h is not constant and not an n-ary projection. Then h(0) = 0 and h(1) = 1.
In addition, there must exist an n-tuple a = (a1 , . . . , ai , . . . , aj , . . . , an ) with
a 6= 0 and a 6= 1, but h(a) = 1; that is, there is at least one i and a num-
ber j, with 1 ≤ i, j ≤ n, such that either ai = 0 and aj = 1 or aj = 0
and ai = 1. Since h is not an n-ary projection, there is an n-tuple b =
(b1 , . . . , bi , . . . , bj , . . . , bn ) with h(b) 6= bj . By identification and permutation
of variables we can make a binary term operation h0 with h0 (0, 0) = 0,
h0 (0, 1) = 1, h0 (bi , bj ) 6= bj and h0 (1, 1) = 1. Since h0 is monotone we must
have bj = 0 and thus bi = 1, and so h0 (x, y) = x ∨ y.
But this means we can produce the join operation ∨ as a term operation of
A. If the binary term operations of A are contained in M1 , we have both ∨
and ∧, and from these we obtain (x ∧ y) ∨ (x ∧ z) ∨ (y ∧ z). Similarly, if the set
of all binary term operations of A is contained in the set M3 , we have both
¬ and ∨ as term operations, and again can produce (x ∧ y) ∨ (x ∧ z) ∨ (y ∧ z).
But this makes A congruence modular.
222 CHAPTER 10. CLONES AND COMPLETENESS
Analysis of the sets M1 and M3 yields the following description of all two-
element algebras which generate varieties which are not congruence modular.
As usual, we use Post’s names for the algebras.
P6 = ({0, 1}; ∧, c20 , c21 ), P3 = ({0, 1}; ∧, c20 ), P5 = ({0, 1}; ∧, c21 ),
P1 = ({0, 1}; ∧), O9 = ({0, 1}; ¬, c20 ),
O8 = ({0, 1}; e21 , c20 , c21 ), O6 = ({0, 1}; e21 , c20 ),
O4 = ({0, 1}; ¬), O1 = ({0, 1}; e21 ).
Having found all two-element algebras which generate varieties which are
not congruence modular, we turn now to the other case, of two-element
algebras which generate congruence modular varieties. This case will also
be treated in two subcases, depending on whether the variety generated
is congruence distributive or not. First, suppose that V (A) is congruence
modular but not congruence distributive. In this case, we know that V (A) is
congruence permutable. By Theorems 9.6.3 and 9.6.4, the operation x+y +z
is a term operation of A, but none of (x ∧ y) ∨ (x ∧ z) ∨ (y ∧ z), x ∧ (y ∨ z),
x ∨ (y ∧ z), (x ∨ ¬ y) ∧ z and (x ∧ ¬ y) ∨ z are term operations of A. Clearly,
the clone h{x + y + z}i is the least clone of Boolean operations to contain
the operation x + y + z. From a well-known result of R. McKenzie in [78] it
follows that the greatest clone of Boolean operations to have this property,
that is, to contain x + y + z but none of (x ∧ y) ∨ (x ∧ z) ∨ (y ∧ z), x ∧ (y ∨ z),
x ∨ (y ∧ z), (x ∨ ¬ y) ∧ z and (x ∧ ¬ y) ∨ z, is the clone consisting of all
linear Boolean operations. Linear Boolean operations are those of the form
f (x1 , . . . , xn ) = c0 + c1 x1 + . . . + cn xn , for ci ∈ {0, 1}. It is easy to show that
10.3. THE LATTICE OF ALL BOOLEAN CLONES 223
the interval [h{x + y + z}i, h{+, c21 }i] contains exactly the following clones of
Boolean operations:
h{+, c21 }i, h{x + y + 1}i, h{¬, x + y + z}i, h{+}i, h{x + y + z}i.
This tells us all the two-element algebras which generate a congruence mod-
ular but not congruence distributive variety, namely
Consider first the case that V (A) is 2-distributive. Then by Lemma 9.6.1,
the majority operation (x ∧ y) ∨ (x ∧ z) ∨ (y ∧ z) is a term operation of A.
Using the Baker-Pixley Theorem, Theorem 9.2.6, we can describe cloneA
as the clone consisting of exactly all operations defined on A which pre-
serve all subalgebras of the direct square A2 . By examining all sublattices
of the lattice of all subsets of A2 which can be subalgebra lattices of A2
we can determine all two-element algebras with a majority term operation
(x ∧ y) ∨ (x ∧ z) ∨ (y ∧ z). To describe them all, we need the following five
relations on A:
It is not hard to determine a generating system for each clone in this list,
and from this we produce our list of all two-element algebras which generate
224 CHAPTER 10. CLONES AND COMPLETENESS
a 2-distributive variety:
For the next case, we assume now that V (A) is congruence distributive, but
not 2-distributive. Then x ∧ (y ∨ z) or (x ∨ (y ∧ z)) is a term operation,
but (x ∧ y) ∨ (x ∧ z) ∨ (y ∧ z) is not a term operation. We consider two
further subcases here, depending on whether the term clone cloneA can be
represented as P ol ρ for some u-ary relation ρ on {0, 1}. The next lemma,
due to M. Reschke and K. Denecke ([99]), handles the case where cloneA =
P olρ for some at least ternary relation. It uses a function hµ with properties
corresponding to the generalized Baker-Pixley Theorem (see Remark 9.2.7).
Such functions are called near-unanimity functions. The generalized Baker-
Pixley Theorem shows that a function f is a term function of A if and only
if f preserves all subalgebras of Aµ .
Lemma 10.3.3 ([99]) Let cloneA = P ol ρ for a µ-ary relation ρ ⊆ {0, 1}µ ,
where µ ≥ 3. Then V (A) is congruence distributive but not 2-distributive if
and only if the following operation hµ or its dual h0µ is a term operation in
10.3. THE LATTICE OF ALL BOOLEAN CLONES 225
cloneA:
µ+1
_
hµ (x1 , . . . , xµ+1 ) = (x1 ∧ . . . ∧ xi−1 ∧ xi+1 ∧ . . . ∧ xµ+1 ).
i=1
a11 aµ+11
.. ..
a1 = . ∈ ρ, . . . , aµ+1 = ∈ ρ.
.
a1 µ aµ+1 µ
Since cloneA = P olρ, it is enough to show that
b1
..
hµ (a1 , . . . , aµ+1 ) = . = b ∈ ρ.
bµ
The least clone of the form cloneA = P ol ρ such that V (A) is 4-distributive
but not 2-distributive is the clone generated by hµ . It can be shown that
this clone is P ol{Rµ , ≤, {1}}, where Rµ is the relation {0, 1}µ \ {(1, . . . , 1)}
226 CHAPTER 10. CLONES AND COMPLETENESS
for µ ≥ 3. Dually, we have h{h0µ }i = P ol{Rµ0 , ≤, {0}} for the relation Rµ0 =
{0, 1}µ \ {(0, . . . , )}. There are four clones of this kind, for µ ≥ 3:
We turn now to our last case, where we have V (A) congruence distribu-
tive but not 2-distributive, but there is no finitary relation ρ on {0, 1} such
that cloneA = P ol ρ. By Lemma 10.3.3 we know that hµ 6∈ cloneA. The
least clone cloneA such that V (A) is 4-distributive and not 3-distributive is
the clone generated by x ∧ (y ∨ z) or x ∨ (y ∧ z), and the least clone such
that V (A) is 3-distributive and not 2-distributive is the clone generated by
(x ∧ ¬ y) ∨ z or (x ∨ ¬ y) ∧ z. It is easy to show moreover that the only other
clones with the first property are h{x ∧ (y ∨ z), c10 }i and h{x ∨ (y ∧ z), c11 }i,
and the only other clones of the second type are h{(x ∧ ¬ y) ∨ z, c10 }i and
h{(x ∨ ¬ y) ∧ z, c11 }i.
Cr1
C2 r rC3
Cr4
Ar1
A2 A3
r r
r
F24 r A4 rF2
F23 F27 8
r r r r
2
F1 r r F25
F22 F26
F34 r F33 F 3 rF3
8
r r r 7 r
F31 r Dr3 r F35
F32 F36
r
D 1
r r rF∞
F∞
4 F∞
3 D2 F∞
7 8
r r r r ∞
L1
F∞
1 r r r F 5
F∞
2 r rL5 r F∞
6
rS6 L2 L3 P6 r
r
S3 r rS5
L4 P5 r r P3
rS r
1 Or9 P1
r
O8
O5 r r r O6
r O4
O1
This completes our survey of all two-element algebras, since all our cases
have now been considered. The diagram shows the lattice of all clones of
Boolean operations.
In the case that A is a two-element set, we consider the clone of all Boolean
functions. We know already that in this case O(A) is generated by {∨, ¬}
and thus is finitely generated. Post’s results show that there are exactly 5
maximal subclones of O(A): these are C2 , C3 , A1 , D3 and L1 . Hence a set C
of Boolean functions is functionally complete iff for each of these five maxi-
mal subclones there exists a function f ∈ C which is not an element of this
maximal subclone.
Then h{+ , · , (χa )a∈A }iO(A) = O(A), and any operation f : An → A may
be expressed as
X Y
f (x1 , . . . , xn ) = f (a1 , . . . , an ) · χai (xi ),
(a1 ,...,an )∈An i≤n
Q
where Σ means iteration of + in a canonical way and yi is defined simi-
i≥1
yi ) · yn+1 .
Q Q Q
larly by yi := y1 and yi := (
i≤1 i≤n+1 i≤n
Note that in the Boolean case A = {0, 1}, we have χ0 (x) = ¬x and χ1 (x) =
x. If we take + to represent disjunction and · to represent conjunction, the
representation given in Theorem 10.4.1 is just the usual disjunctive normal
form.
Now that we know that O(A) is finitely generated when A is finite, we must
look for the maximal subclones of O(A). These are in fact known for every
finite A. The deep theorem which describes explicitly all maximal clones on
O(A) is due to I. G. Rosenberg ([103], [102]). The special cases where |A| is
2, 3 or 4 were solved earlier by E. L. Post ([95]), S.V. Jablonskij ([59]), and
A.I. Mal’cev (unpublished). New proofs of Rosenberg’s Theorem were also
given by R. W. Quackenbush ([97]) and D. Lau ([70]).
Corollary 10.4.2 Let A be a finite set and let F be a subset of O(A). Then
hF iO(A) = O(A) if and only if F is not contained in one of the maximal
subclones of O(A).
(ii) Let ρ ⊆ A2 be a partial order with least and greatest elements 0 and 1,
respectively. Then P olρ is a maximal clone.
The well-known Slupecki criterion says that f ∈ P olιn iff f is not surjective
or f depends essentially on at most one variable ([109]). P ol% is a maximal
clone for every t-universal relation.
It follows from the definition that any primal algebra is functionally com-
plete. Another connection between these two properties is based on the fol-
lowing construction. For any finite algebra A, we can form a new algebra A+
from A by adding as fundamental operations every constant operation ca , for
a ∈ A. (The finiteness of A means that we still have only finitely many fun-
damental operations in this algebra.) That is, A+ = (A; F A ∪ {ca | a ∈ A}).
Then the following result is easily verified, and its proof is left as an exercise
for the reader.
The functional completeness criterion from Section 10.4 can be used to obtain
examples of primal algebras. It is very easy to see in this way that the two-
element Boolean algebra is primal. We list here some more examples of
primal and functionally complete algebras; we leave the verification, using
Theorem 10.4.1 and Corollary 10.4.2, to the reader.
E. L. Post showed in [94] that Ak is primal. (Note that for k = 2, this is the
algebra ({0, 1}; ∧, ¬).)
10.5. PRIMAL ALGEBRAS 233
Theorem 10.5.5 For every prime number p, the prime field modulo p is
primal.
Proof: The prime field modulo p is up to isomorphism the algebra Zp =
(Zp ; +, −, ·, 0, e) of type τ = (2, 1, 2, 0, 0) where Zp is the set of the residue
classes modulo p. We can use the fundamental operations of Zp to construct,
for each k ∈ Zp , the unary operation χk defined by χk (x) := e − (x − k)p−1 .
These operations satisfy the following property:
(
e if x = k
χk (x) =
0 otherwise.
Note that the Baker-Pixley Theorem, Theorem 9.2.6, also gives a character-
ization of primal algebras. It can be formulated as follows:
Proposition 10.5.7
(i) Let A be a primal algebra. Then
1. A has no proper subalgebras,
2. A has no non-identical automorphisms,
3. A is simple, and
4. A generates an arithmetical variety.
(ii) Every functionally complete algebra is simple.
10.5. PRIMAL ALGEBRAS 235
Proof: (i) Since every operation defined on the set A is a term operation of
the algebra A, for each proper subset B ⊂ A, each non-identical permutation
ϕ on A and each non-trivial equivalence relation θ ⊆ A2 there exist term
operations tA A A
1 , t2 , t3 of A such that
A
t1 (b1 , . . . , bn ) 6∈ B for some elements b1 , . . . , bn ∈ B,
ϕ(tA A
2 (a1 , . . . , an ) 6= t2 (ϕ(a1 ), . . . , ϕ(an )) for some a1 , . . . , an ∈ A, and
A A
(t3 (a1 , . . . , an ), t3 (b1 , . . . , bn )) 6∈ θ for some pairs (ai , bi ) ∈ θ, i = 1, . . . , n.
The variety V (A) is then arithmetical since there is a term q satisfying the
identities of Theorem 9.3.2 (iii) which induces a term operation on A.
It turns out that the four properties of primal algebras given in Proposition
10.5.7 are sufficient to characterize primal algebras. An additional character-
ization was given by H. Werner in [118] (see also A. F. Pixley, [87]), using
the so-called ternary discriminator term:
(
z if x = y
t(x, y, z) =
x otherwise.
Theorem 10.5.8 For a finite algebra A the following propositions are equiv-
alent:
(i) A is primal.
Proof: (i) ⇒ (iii): This is clear since when A is primal every operation, in-
cluding the ternary discriminator, induces a term operation on A. The other
two properties were proved in the previous theorem.
Step 1a). We show first that B has the following property, which is called
rectangularity:
Step 1b). Using the projection homomorphisms we define our algebras and
congruences:
Step 1c). Now we define a mapping ϕ : B1 /θ1 → B2 /θ2 , by [a]θ1 7→ [d]θ2 for
all (a, d) ∈ B. We will show in this step that ϕ is an isomorphism.
First we check that ϕ is well defined. If [a]θ1 = [a0 ]θ1 , and (a, d) is in B, then
(a, a0 ) ∈ θ1 and so (a, e) and (a0 , e) ∈ B for some e ∈ A. Then by definition
of θ2 , having both (a, d) and (a, e) in B means that (d, e) ∈ θ2 . From this we
have ϕ([a]θ1 ) = [d]θ2 = [e]θ2 = ϕ([a0 ]θ1 ).
To see that the mapping ϕ is injective, suppose that [d]θ2 = ϕ([a]θ1 ) =
ϕ([a0 ]θ1 ) = [e]θ2 . Then (a, d) and (a0 , e) are in B, and (d, e) ∈ θ2 means that
(c, d) and (c, e) are in B for some c ∈ A. Then (a, c) and (a0 , c) are both in θ1 .
By symmetry and transitivity of θ1 we get (a, a0 ) ∈ θ1 , and so [a]θ1 = [a0 ]θ1 .
and
f B2 (ϕ([a1 ]θ1 ), . . . , ϕ([an ]θ1 )) = f B2 ([d1 ]θ2 , . . . , [dn ]θ2 ).
Step 1d). Now we show that B = {X × ϕ(X) | X ∈ B1 /θ1 }. First,
S
opposite inclusion, assume now that (a, d) ∈ X × ϕ(X) for X ∈ B1 /θ1 , say
X = [b]θ1 and ϕ(X) = [e]θ2 with (b, e) ∈ B. Then (a, d) ∈ [b]θ1 × [e]θ2 ,
so that (a, b) ∈ θ1 and (d, e) ∈ θ2 . The definitions of θ1 and of θ2 re-
spectively give us elements d0 and a0 ∈ A for which (a, d0 ), (b, d0 ), (a0 , d)
and (a0 , e) are in B, and by the rectangularity property from Step 1a)
we have (b, d) ∈ B and (a, d) ∈ B. This means that X × ϕ(X) ∈ B and
{X × ϕ(X) | X ∈ B1 /θ1 } ⊆ B. Altogether we have equality.
S
Combining the result of Step 1 with our assumption that A has no proper
subalgebras, we see that B1 = B2 = A. This means that θ1 and θ2 are ac-
tually congruences on A, and since A is simple we see that θ1 and θ2 must
each be one of ∆A or A2 . If θ1 = A2 , then the isomorphism ϕ : A/θ1 → A/θ2
shows that θ2 = A2 also, and in this case we have B = A2 . We also have B
= {(a, ϕ(a)) | a ∈ A}. If instead θ1 = ∆A , then the isomorphism ϕ shows
that θ2 must also equal ∆A . Again B has the form {{a} × ϕ({a}) | a ∈ A},
S
We point out that for the proof of the claim of Step 1 we used only con-
gruence permutability. We will make use of this fact later, in the proof of
Theorem 10.6.5.
x + y := t(x, 0, y),
x · y := t(y, 1, x),
χ0 (x) := t(y, 1, x),
χa (x) := t(t(0, a, x), 0, 1), for a ∈ A \ {0}.
This theorem can be proved very easily using Rosenberg’s description of all
maximal classes of operations defined on a finite set by relations, and we
leave it as an exercise for the reader.
There are two other kinds of algebras which are defined in a different way. A
finite non-trivial algebra is called cryptoprimal if A is simple, has no proper
subalgebra and generates a congruence distributive variety. Also, A is called
paraprimal if each subalgebra of A is simple and A generates a congruence
permutable variety.
(i) A is quasiprimal.
(iii) ⇒ (i): By Theorem 9.3.2 the variety V (A) is arithmetical and there is a
majority term operation in A. Therefore we can apply Theorem 9.2.6. The
variety V (A) is congruence permutable since it is arithmetical, and the claim
from Step 1 in the proof of Theorem 10.5.8 is satisfied. Thus there are sub-
algebras B1 and B2 of A, congruence relations θ1 ∈ ConB1 and θ2 ∈ ConB2
and an isomorphism ϕ : B1 /θ1 → B2 /θ2 such that the universe of B can be
written as B = {X × ϕ(X)|X ∈ B1 /θ1 }. Since by assumption each subal-
S
(iv) No two distinct subalgebras with more than one element are isomor-
phic.
Now we will prove that from conditions (i) - (iv) the semiprimality of A
can be deduced. By Theorem 10.6.5 the algebra A is quasiprimal. There are
no isomorphisms between non-trivial subalgebras of A, and therefore each
operation preserving all subalgebras of A is a term operation of A. Conse-
quently, A is indeed semiprimal.
Conversely, from conditions (i) - (iv) and Theorem 10.6.5 we see that A is
quasiprimal. Since A has no proper subalgebras, each isomorphism between
non-trivial subalgebras of A is an automorphism of A , and A is demipri-
mal.
Theorem 10.6.9 Let A be a finite non-trivial algebra and let V (A) be the
variety generated by A. Then:
Proof. “⇒”: Assume that A generates an arithmetical variety and the three
conditions of Case 1 are satisfied. Since A is simple and the only other sub-
algebra B of A is also simple, since it is either primal or trivial, we conclude
that all subalgebras of A are simple. The algebra A has no non-identical
automorphisms and the subalgebra B, as a trivial or primal algebra, also has
no non-identical automorphisms; hence all subalgebras of A have no proper
automorphisms. This means that all conditions of Theorem 10.6.6 are satis-
fied, and A is semiprimal. Since A has only one proper subalgebra, by the
definition of semiprimality every operation defined on A which preserves the
subset B ⊂ A is a term operation of A, and cloneA is a clone of operations
preserving a central relation with h = 1 (class (v)), making A preprimal.
Next assume that A generates an arithmetical variety and that the three
conditions of Case 2 are satisfied. Let the automorphism group of A be the
cyclic group of prime order p. We want to apply Theorem 10.6.7. Clearly,
AutA is generated by an automorphism s of prime order p 6= 1. The auto-
morphism s has no fixed points, since the fixed points of an automorphism
form a subalgebra but A has no proper subalgebras, forcing p = 1, a contra-
diction. Therefore A is demiprimal.
248 CHAPTER 10. CLONES AND COMPLETENESS
Now assume that A generates an arithmetical variety and that the conditions
of Case 3 are satisfied. Then all conditions of Theorem 10.6.8 are fulfilled;
so A is hemiprimal and cloneA = P olθ for a non-trivial congruence relation
θ of A. Then A is preprimal of type (iv).
Preprimal algebras A for which cloneA = P olθ for some non-trivial equiva-
lence relation θ are the only non-simple preprimal algebras. But such algebras
have only one non-trivial congruence relation; so they are subdirectly irre-
ducible. This shows that all preprimal algebras are subdirectly irreducible.
The question of describing all other subdirectly irreducible algebras in the
variety V (A) generated by a preprimal algebra A was solved for most classes
of preprimal algebras, independently by K. Denecke in [20] and A. Knoebel
in [65]. This is closely connected to the problem of describing the subvariety
lattice of varieties V (A) generated by preprimal algebras A. The results are
10.8. EXERCISES 249
summarized in the following table; the proofs may be found in [20] and [65].
Number of Number of
Class Relation Subvarieties Subdirectly
of V (A) Irreducibles
(i) fixed point free permutation 2 1
(ii) bounded partial order 2 ?
(iii) elementary abelian p-group 3 2
(iv) equivalence relation 3 2
(v) h = 1, subset |B| = 1 2 1
(v) h = 1, subset |B| > 1 3 2
(v) h>2 2 1
(vi) 5 or 6 finite
10.8 Exercises
10.8.1. Prove directly that the pair (P ol, Inv) of operators introduced in
Section 10.2 forms a Galois connection.
10.8.7. Prove that for preprimal algebras A corresponding to class (i), the
variety V (A) has no non-trivial subvarieties and A is the only subdirectly
irreducible algebra in V (A).
10.8.8. Prove that for preprimal algebras A corresponding to class (iv), the
variety V (A) has only one non-trivial subvariety and two subdirectly irre-
250 CHAPTER 10. CLONES AND COMPLETENESS
ducible algebras.
10.8.9. Prove that for preprimal algebras A corresponding to class (v), when
|B| = 1 the variety V (A) has no non-trivial subvarieties.
10.8.10. Let A be the two-element set {0, 1}. Prove that an operation f :
An → A is a linear Boolean function iff it is in P olρ, where
ρ = {(y1 , y2 , y3 , y4 ) ∈ A4 | y1 + y2 = y3 + y4 }.
Chapter 11
An algebra A is called finite when its universe A is finite and every funda-
mental operation is finitary. Finite algebras are important in many areas
where finiteness plays a crucial role, for instance in Computer Science. A
major area of research activity has been to try to classify all finite algebras
of a given type. For instance, the classification of all finite groups has been
a longstanding mathematical problem.
251
252 CHAPTER 11. TAME CONGRUENCE THEORY
Our first theorem of this chapter shows that under certain conditions, the
restriction of a congruence on A to a subset U of A is also a congruence
relation, on the induced algebra A|U .
Theorem 11.1.2 Let A be an algebra and let e ∈ E(A). Let U be the image
set U := e(A). Then the mapping ϕU : Con A → Con(A|U ), defined by
θ 7→ θ|U , is a surjective lattice homomorphism.
Proof: We must show first that if θ ∈ Con A, then θ|U is in Con(A|U ),
so that our mapping ϕU is well defined. Let (a1 , b1 ), . . . , (an , bn ) be in θ|U ,
so that (a1 , b1 ), . . . , (an , bn ) are all in θ ∩ U 2 . The fundamental operations
of the induced algebra A|U have the form f A |U , where f A is a polynomial
operation on A. Then we have
11.1. MINIMAL ALGEBRAS 253
Next we must show that the mapping ϕU is compatible with the lattice op-
erations ∧ (which is just ∩) and ∨ on the congruence lattices. For ∧, we let
θ and ψ be in ConA, and let a, b ∈ A. Then (a, b) ∈ θ|U ∩ ψ|U ⇔ (a, b) ∈
(θ ∩ U 2 ) ∩ (ψ ∩ U 2 ) ⇔ (a, b) ∈ θ ∩ ψ ∩ U 2 ⇔ (a, b) ∈ (θ ∩ ψ)|U . It follows that
ϕU (θ) ∩ ϕU (ψ) = ϕU (θ ∩ ψ).
x a b c d
f (x) b c c d
The mapping f is not idempotent, but nevertheless we can form the set
U := f (A) = {b, c, d}, and compare the lattices of congruences on the orig-
inal algebra A and the induced algebra A|U . It is straightforward to work
out that lattice Con A has the Hasse diagram shown below, with θ0 and θ1
denoting the trivial congruences ∆A and A × A, respectively.
11.1. MINIMAL ALGEBRAS 255
θq1
θ2 q q θ5
θ3 q q θ4
q
θ0
Our induced algebra A|U has universe U , and fundamental operation set
P (A)|U . There is only one unary term operation on A which preserves the
subset U = {b, c, d}, namely the term g := f |U = (f ◦ f )|U , given by the
table below.
x b c d
g(x) c c d
Thus P (A)|U is generated by g and the three constant mappings with values
in U . Using this, we can work out that this algebra A|U has four congru-
ences:
θ20 s s θ0
3
s
θ00
θ1 7→ θ1 |U = θ10 , θ0 7→ θ0 |U = θ00 ,
θ2 7→ θ2 |U = θ10 , θ3 7→ θ3 |U = θ30
θ4 7→ θ4 |U = θ20 , θ5 →
7 θ5 |U = θ20 .
(ii) for all g ∈ P 1 (A) with g(A) ⊆ f (A) and |g(A)| > 1, we have g(A) =
f (A).
x m1 (x) m2 (x)
0 0 0
1 2 0
2 2 2
Then cloneA = P (A) is the set of all operations which are monotone with
respect to the usual order 0 ≤ 1 ≤ 2. For each of the following monotone
operations f1 , . . ., f6 ,
11.1. MINIMAL ALGEBRAS 257
x f1 f2 f3 f4 f5 f6
0 0 0 0 0 1 1
1 0 1 0 2 1 2
2 1 1 2 2 2 2
the corresponding image set fi ({0, 1, 2, }) is a minimal set. Notice that these
are all idempotent operations.
Theorem 11.1.6 Let A be a finite algebra and let U ∈ M in(A). Then every
unary polynomial operation of the induced algebra A|U is either a permuta-
tion or a constant operation. Moreover, the non-constant unary polynomial
operations of A|U form a permutation group on U .
This means that the unary polynomial operations of A|U are exactly the
mappings of the form g|U for some g ∈ P 1 (A) such that g(U ) ⊆ U . Now
suppose that such a mapping is not a permutation of the set U . Since
U is minimal, it has the form U = f (A) for some f ∈ P 1 (A). Then
g(f (A)) = g(U ) ⊂ U = f (A). But again since U is minimal we must have
|g(U )| = |g(f (A))| = 1. Therefore g|U is constant on U .
Algebras with this property, that every non-constant unary polynomial op-
eration is a permutation, are called permutation algebras. Our theorem thus
says that any minimal algebra of a finite algebra A is a permutation algebra.
Later we will show that there are exactly five types of permutation algebras,
and thus five types of minimal algebras.
(ii) for any g ∈ P 1 (A) with g(A) ⊆ f (A) and g(β) 6⊆ ∆A , we have g(A) =
f (A).
The set of all β-minimal sets of an algebra A is denoted by M inA (β), or
when the algebra is clear from the context, just M in(β). The corresponding
induced algebras A|U , for U ∈ M in(β), are called β-minimal algebras of A,
or minimal algebras with respect to β.
β = {(a, a), (b, b), (c, c), (d, d), (b, c), (c, b), (b, d), (d, b), (c, d), (d, c)}.
For the operation g given by the table
x a b c d
g(x) c c c d
we have an image set U = g(A) = {c, d}. The image g(β) =
{(c, c), (d, d), (c, d), (d, c)} contains a non-diagonal pair, and moreover there
is no other unary polynomial with this property which gives a smaller image.
Thus our set g(A) meets both conditions regarding our congruence, and is a
β-minimal set of A.
Definition 11.1.7 is in fact a generalization of Definition 11.1.4. A minimal
set for an algebra A is always a β-minimal set for β equal to the largest
congruence A×A on A. The first condition for β-minimal sets, in this special
case, merely says that the cardinality of the minimal set is greater than
one, while the second condition translates directly. We thus have M in(A) =
M inA (A × A).
We now generalize the characterization from Theorem 11.1.6 of minimal
algebras of a finite algebra.
11.1. MINIMAL ALGEBRAS 259
Now let N be a β|U congruence class of A|U . Every unary polynomial oper-
ation of (A|U )|N has the form h|N for some h ∈ P 1 (A) with h(U ) ⊆ U and
h(N ) ⊆ N . Then h|U is either a permutation on U , or satisfies h(β|U ) ⊆ ∆U .
In the first case h|N is a permutation on N , while in the second case h|N is
constant.
α = {(0, 0)(1, 1), (2, 2), (3, 3), (4, 4), (5, 5), (0, 3), (3, 0), (1, 4), (4, 1),
(2, 5), (5, 2)},
β = {(0, 0), (1, 1), (2, 2), (3, 3), (4, 4), (5, 5), (0, 2), (2, 0), (0, 4), (4, 0)),
260 CHAPTER 11. TAME CONGRUENCE THEORY
(2, 4), (4, 2), (1, 3), (3, 1), (1, 5), (5, 1), (3, 5), (5, 3)}.
The unary polynomial operations of our algebra have the form f (x) = ax+b,
for some a, b ∈ Z6 . From this information we can calculate all the minimal
sets of this algebra:
M in(Z6 × Z6 ) = {{0, 3}, {1, 4}, {2, 5}, {0, 2, 4}, {1, 3, 5}},
M in(α) = {{0, 3}, {1, 4}, {2, 5}},
M in(β) = {{0, 2, 4}, {1, 3, 5}}.
Lemma 11.1.14 Let L be a finite lattice with least element 0, greatest el-
ement 1 and the property that the meet of arbitrary coatoms (elements di-
rectly below 1) is zero. Then L has no non–constant strongly extensive meet–
endomorphisms.
r r r r p p p r
(ii) θ ⊃ ∆A ⇒ θ|U ⊃ ∆U ,
(iii) θ ⊂ A × A ⇒ θ|U ⊂ U × U .
Let us remark that the concept of a tame congruence can be generalized even
further, in the following way. Let A be finite and α and β be congruences
on A with α ⊂ β. Then the interval [α, β] ⊆ ConA is called tame if the con-
gruence relation β/α of the quotient algebra A/α is tame. The β/α-minimal
sets, and the corresponding induced algebras, of the quotient algebra A/α
are called [α, β]-minimal. (Note that by the Second Isomorphism Theorem,
Theorem 3.2.2, the intervals [α, β] ⊆ ConA and [∆A/α , β/α] ⊆ ConA/α are
isomorphic.)
(Z2) θ(β|U ) = β; that is, for all (x, y) ∈ β there are pairs
(a1 , b1 ), . . . , (ak , bk ) ∈ β|U and polynomials f1 , . . . , fk ∈ P 1 (A) with
x = f1 (a1 ), fi (bi ) = fi+1 (ai+1 ) for i = 1, . . . , k − 1, and fk (bk ) = y.
(Z1) ⇒ (ii): Assume now that (Z1) is satisfied and that θ ∈ ConA with
∆A ⊂ θ ⊆ β. Then for any (x, y) ∈ θ with x 6= y, by (Z1) there is a polyno-
mial f ∈ P 1 (A) with (f (x), f (y)) ∈ θ ∩ U 2 = θ|U and f (x) 6= f (y). Thus (ii)
is satisfied.
(Z2) ⇔ (iii): The condition (iii), that if θ ⊂ β then θ|U ⊂ β|U , for all
θ ∈ [∆A , β], is by Theorem 11.1.2 equivalent to the fact that the congruence
relation on A generated by β ∩ U 2 is equal to β; that is, θ(β|U ) = β. But
since θ(β|U ) ⊆ β and β|U ⊆ θ(β|U ), we do have θ(β|U ) = β.
11.2. TAME CONGRUENCE RELATIONS 265
For all θ ∈ ConA, it follows from f ∈ P 1 (A) that f (θ) ⊆ θ. The equation
f (B) = C gives f (θ|B ) = f (θ ∩ B 2 ) ⊆ θ ∩ C 2 = θ|C . Similarly, we obtain
g(θ|C ) ⊆ θB and then θ|C = f (g(θ|C )) ⊆ f (θ|B ). Altogether we have equal-
ity.
266 CHAPTER 11. TAME CONGRUENCE THEORY
(ii) For all U ∈ M inA (β), the conditions (Z1), (Z2) from Theorem 11.2.4
and (i),(ii),(iii) from Definition 11.2.3 are satisfied.
(iii) Let U ∈ M inA (β), with U = f (A) for some f ∈ P 1 (A) with
f (β|U ) 6⊆ ∆A . Then f (U ) is also in M inA (β), and f |U is a polynomial
isomorphism from U onto f (U ).
Proof: (i) Since β is tame there is a set W ∈ M inA (β) satisfying con-
ditions (i), (ii), (iii) from Definition 11.2.3 and (Z1), (Z2) from Theorem
11.2.4. Moreover, we may assume that W = e(A) for an idempotent unary
polynomial e. We show that for every U ∈ M inA (β), the sets U and W
are polynomially isomorphic. Since U is β-minimal, there is a polynomial
operation s ∈ P 1 (A) with s(A) = U and s(β) 6⊆ ∆A . Then there is a pair
(x, y) ∈ β with s(x) 6= s(y). Applying condition (Z2) for W to the pair
(x, y), we obtain the existence of a pair (a, b) ∈ β|W and of a polynomial
h ∈ P 1 (A) such that s(h(a)) 6= s(h(b)) (since otherwise s(x) = s(y)). Now
we define s1 to be the composition mapping she. Then we have s1 (A) = U ,
since s(h(e(A))) = s(h(W )) ⊆ U because of the β-minimality of U . Ap-
plying (Z1) to the pair (s1 (a), s1 (b)) gives us a polynomial t ∈ P 1 (A) with
t(A) = W and ts1 (a) 6= ts1 (b). By the β-minimality of W again, we have
ts1 (W ) = W . Now we use s1 and t to construct the required polynomial iso-
morphism between U and W . The mapping s1 t|U satisfies s1 (t(U )) ⊆ U and
because of the minimality of U we have s1 (t(U )) = U . This shows that s1 t|U
is a permutation on the finite set U . Therefore there is an element k ∈ IN with
(s1 t)k = idU . Taking f := t and g := (s1 t)k−1 s1 , we have f (U ) = t(U ) = W ,
g(W ) = (s1 t)k−1 s1 (W ) = (s1 t)k−1 (U ) = U and gf = (s1 t)k−1 s1 t = idU .
Since every w ∈ W can be written as f (u) for some u ∈ U , and
(ii) Let U be a β-minimal set. With the help of the polynomial isomorphisms
f |U and g|W from the proof of (i), the properties (Z1) and (Z2) are satisfied
11.2. TAME CONGRUENCE RELATIONS 267
for U . For (Z2) we use the additional fact that f (β|U ) = β|W , which follows
from Lemma 11.2.6. The conditions (i), (ii) and (iii) of Definition 11.2.3 then
follow by Theorem 11.2.4.
(iii) Since f (β|U ) 6⊆ ∆A , and since by (i) of this proof all U ∈ M inA (β)
have the same cardinality, we get f (U ) ∈ M inA (β). For the polynomial iso-
morphism, we have to show the existence of a polynomial g ∈ P 1 (A) with
g(f (U )) = U , gf |U = idU , and f g|f (U ) = idf (U ) . Assume that (a, b) ∈ β|U
with f (a) 6= f (b). Then by condition (Z1) there is an element h ∈ P 1 (A)
with h(A) = U and h(f (a)) 6= h(f (b)). This gives h(f (U )) = U . Then as in
the proof of part (i) there is a k ∈ IN with (hf )k = idU and the mapping
g := (hf )k−1 h has the desired property.
lattices are 0-1-simple. By Lemma 11.1.14, all finite 0-1-simple lattices whose
coatoms have meets equal to zero are tight.
r r r r p p p r
are tight.
α = {(0, 0), (1, 1), (2, 2), (3, 3), (4, 4), (5, 5), (0, 3), (3, 0), (1, 4), (4, 1), (2, 5),
(5, 2)}
11.3. PERMUTATION ALGEBRAS 269
and
β = {((0, 0), (1, 1), (2, 2), (3, 3), (4, 4), (5, 5), (0, 2), (2, 0), (0, 4), (4, 0), (2, 4),
(4, 2), (1, 3), (3, 1), (1, 5), (5, 1), (3, 5), (5, 3)}
are tame, since both intervals [∆Z6 , α], [∆Z6 , β] are simple. But the algebra
Z6 is not tame, since neither congruence has {0} or {1} as singleton blocks.
All two-element lattices are tight. For any algebra A, we can look at what are
called prime intervals [α, β] ∈ ConA, intervals which contain only the two
congruences α and β. For minimal sets which correspond to prime intervals,
the conditions (Z1) and (Z2) are satisfied. This means that when A is finite
we can consider all prime intervals. In this way tame congruence theory gives
a structure theory for all finite algebras.
Before we begin the proof of this theorem we need the following fact:
(Q1) x \ (x + y) ≈ y
(Q2) (x + y)/y ≈ x
(Q3) x + (x \ y) ≈ y
(Q4) (x/y) + y ≈ x.
11.3. PERMUTATION ALGEBRAS 271
Now we claim that this forces all the polynomials La0 , for a0 6= a, to be
non-constant. For suppose that there was some a0 6= a for which La0 was
constant. Then La0 (x) = La0 (c) = a0 + c = s, so that La0 always gives the
value s. Thus La0 = La , and for every x ∈ A we have a0 + x = a + x. But
then for every x ∈ A we have Rx (a0 ) = a0 + x = a + x = Rx (a). Then Rx
is not injective, hence not a permutation, hence must be constant. Now we
have all right-additions Rx constant, contradicting the fact that our opera-
tion x + y is essentially binary. Therefore we see that every La0 , for a0 6= a,
is non-constant.
Proof: We recall that a loop is an algebra (L; +, 0) of type (2, 0) such that
(L; +) is a quasigroup and 0 is a neutral element with respect to the operation
272 CHAPTER 11. TAME CONGRUENCE THEORY
Proof: Suppose that for some f and g in G we have f (a) = g(a) and
f (b) = g(b) with a 6= b. Let h be the unary polynomial on A defined by
h(x) := f (x)/g(x), where / is the right subtraction belonging to the loop
operation +, as in Step 2. Then h(a) = f (a)/g(a) = 0 = h(b) = f (b)/g(b), so
h must be constant with value 0. But this means f (x) = g(x) for all x ∈ A,
and f = g.
Step 4. (A; +) is an abelian group.
ture, we use the fact that the set of endomorphisms of an abelian group
forms a ring. The multiplication operation in this ring is composition of
mappings, while addition of operations is defined pointwise, by (f1 + f2 )(a)
= f1 (a) + f2 (a) for all a ∈ A. Let K := {k ∈ G | k(0) = 0} ∪ {0}, where 0 is
the constant mapping with 0(x) = 0 for all x ∈ A.
Step 5. The set K is the universe of a subring of the ring of all endomor-
phisms of the abelian group (A; +, −, 0). Moreover since K is finite and all
elements of K \ {0} are permutations, K = (K, +, −, 0, ◦,−1 , idA ) is a field
and (A; +, −, 0, K) is a vector space over K.
(x1 , . . . , xn ) 7→ a + k1 x1 + . . . + kn xn ,
Proof: By definition every operation of this form belongs to P (A). For the
converse, let f be a polynomial on A of arity n ≥ 0. For n = 0 the claim is
clear, so we assume now that n ≥ 1 and proceed inductively.
Class 1. O1 , O5 , O8 :
- all two-element algebras with unary term operations;
- polynomially equivalent to O8 = ({0, 1}; c10 , c11 ).
Class 2. L1 , L3 , L4 , L5 :
- all two-element algebras which generate a congruence permutable, but not
congruence distributive variety;
- polynomially equivalent to L1 = ({0, 1}; c10 , c11 , +, ¬).
Class 5. P1 , P3 , P6 , P5 :
- polynomially equivalent to P6 = ({0, 1}; c10 , c11 , ∧).
C1
r
A1 r
rL1
P6 r
r
O8
Definition 11.4.1 Let A be a finite tame algebra. If one (and hence all) of
the minimal algebras of A is of type i, for i ∈ {1, 2, 3, 4, 5}, then A is said to
be of type i, and we write type A = i.
the corresponding minimal algebras of the quotient algebra A/α are called
[α, β]-minimal.
Definition 11.4.2 Let C be [θ, Θ]- minimal and let [α, β] be tame in A. A
[θ, Θ]- trace is a set N ⊆ C such that N is a Θ equivalence class [x]Θ for
which [x]Θ 6= [x]θ . An [α, β]- trace N is a set N ⊆ A such that there is some
[α, β]-minimal set U such that N ⊆ U and N is an [α/U, β/U ]- trace of A/U
(so N = [x]β∩U for some x ∈ U with [x]β∩U ∩ U 6⊆ [x]α ).
Theorem 11.4.3 Let [α, β] be a tame interval for the algebra A and let N
be an [α, β]-trace. Then (A|N )/(α|N ) is a minimal algebra. Moreover for all
[α, β]-traces N the minimal algebras (A|N )/(α|N ) have the same type 1–5.
This result allows us to extend our definition of type from minimal and tame
algebras to tame intervals in a congruence lattice.
Definition 11.4.4 Let [α, β] be a tame interval of the finite algebra A. The
type of the interval is the type of the minimal algebra (A|N )/(α|N ) for an
arbitrary [α, β]-trace N .
We saw earlier that all two–element lattices are tight. In a congruence lattice
Con A, a two-element interval is called a prime interval. It follows that on a
finite algebra A all prime intervals [α, β] ⊆ Con A are tame. In this case, for
any [α, β]-minimal set N , the induced algebra A|N is minimal with respect
to (α|N , β|N ).
Definition 11.4.5 The type of a prime interval [α, β] is the type of the
((α|N )/(β|N ))-minimal algebra A|N , where N is an [α, β]-minimal set.
We now present some examples. For the first one, we recall that a finite
algebra A = (A; F A ) is called primal if its term clone is all of OA .
0 ∧ 0 = 0 ∧ 1 = 1 ∧ 0 = 0, 1 ∧ 1 = 1,
0 ∨ 1 = 1 ∨ 0 = 1 ∨ 1 = 1, 0 ∨ 0 = 0,
¬0 = 1, ¬ 1 = 0.
Then the algebra A|U is polynomially equivalent to the two-element Boolean
algebra, which has type 3.
tA (c1 , c2 , . . . , cn ) = tA (d1 , d2 , . . . , dn ) ⇒
tA (b, c2 , . . . , cn ) = tA (b, d2 , . . . , dn ).
tA (a, c2 , . . . , cn ) = tA (a, d2 , . . . , dn ) ⇒
tA (b, c2 , . . . , cn ) = tA (b, d2 , . . . , dn ).
f 0 1 2 3 g 0 1 2 3
0 0 0 1 1 0 0 0 1 1
1 0 0 1 1 1 0 0 1 1
2 0 0 1 1 2 0 0 1 1
3 2 2 3 3 3 2 2 0 0.
It can be verified that for all evaluations f satisfies the implication of the
strong term condition, while g satisfies the implication of the term condition.
Abelian and strongly abelian algebras will be studied further in the next
chapter. For the moment, we characterize the types of such algebras.
Proof: We have shown that the lattices Mn with m ≥ 3 are tight, and
therefore every finite algebra with Con A ∼
= Mn is tame. By Theorem 11.4.8
(i) we have type A ∈ {1, 2}, and then Theorem 11.4.12 (ii) shows that A is
abelian. If n − 1 is not a prime power then by Lemma 11.4.9 and Theorem
11.4.12 A is strongly abelian.
Hobby and McKenzie ([57]) proved the following “omitting type” theorem,
which characterizes the types which cannot occur for a variety by means of
the Mal’cev-condition properties of the variety.
Theorem 11.5.1 Let A be a finite algebra, with V (A) the variety generated
by A. Then:
282 CHAPTER 11. TAME CONGRUENCE THEORY
(iii) V (A) is n-permutable for some n ≥ 2 iff type {V (A)} ⊆ {2, 3};
(v) type {V (A)} ∩ {1, 2} = ∅ iff V (A) is meet-semidistributive iff the class
of all lattices isomorphic to a sublattice of Con B for some B ∈ V (A)
does not contain the lattice M3 .
Lemma 11.5.4 Let V (A) be a locally finite variety. Then type {V (A)} =
{1} iff for every algebra B in V (A) the set Sn1 (B) = ∅.
For algebras which generate congruence modular varieties we have the fol-
lowing useful lemma, due to E.W. Kiss and E. Pröhle ([62]).
11.5. MAL’CEV CONDITIONS AND OMITTING TYPES 283
Case 2: V (A) is not congruence modular. In this case V (A) is neither con-
gruence distributive nor congruence permutable. From our classification of
two-element algebras in Section 10.3, we know that exactly the following
two-element algebras have this property:
P1 = ({0, 1}; ∧), P3 = ({0, 1}; ∧, c10 ), P5 = ({0, 1}; ∧, c11 ),
P6 := ({0, 1}; ∧c10 , c11 ), O1 = ({0, 1}; c11 ), O4 = ({0, 1}; ¬),
O8 = ({0, 1}; c10 , c11 ), O9 = ({0, 1}; ¬, c10 )
If M has type 2, then by Theorem 11.5.1 parts (i) and (v) V (M) cannot
be congruence distributive or meet-semidistributive. Hobby and McKenzie
showed in [57] that M has a Mal’cev operation as a polynomial operation,
so that M+ has a Mal’cev operation as a term operation and thus V (M+ )
is congruence permutable (that is, 2-permutable).
Varieties which are not residually small are called residually large. A variety
is called locally finite if every finitely generated algebra in the variety is finite.
Proposition 11.6.2 Let A be a finite algebra such that ConA has a sublat-
11.7. EXERCISES 287
tice of the form given by the Hasse diagram below (the pentagon) in which
β covers α (there are no congruences properly between them), and such that
type [α, β] = 2 and type [∆A , δ] ∈ {3, 4}.
t
β t
2 t δ
α t 3 or 4
t
∆A
Theorem 11.6.3 Every locally finite variety which omits the types 1 and 5,
and is residually small, is congruence-modular.
11.7 Exercises
11.7.1. Let L be the lattice given by the following Hasse diagram:
1s
c
s
a s s
s d
0 s b
Let µ : L 7→ L be the mapping which maps 0 and a to c and all other ele-
ments to 1. Show that µ is strictly extensive and a meet-endomorphism.
11.7.2. Verify that the lattices from Example 11.2.9 are tight.
11.7.3. Verify that the two lattices shown below are not tight.
288 CHAPTER 11. TAME CONGRUENCE THEORY
s
s
s s s
s s s
s s
s s s
s s s
s
s
11.7.6. Let (S, ·) be a finite semigroup. Show that for each element a ∈ S
there is some integer k ≥ 1 such that e = ak is an idempotent, that is,
a2k = ak . Moreover, there is an integer k such that a2k = ak holds for every
a ∈ S.
In this chapter we show how concepts such as an abelian group, the commu-
tator of a group and related concepts like solvable groups may be generalized
to arbitrary universal algebras. The commutator of two elements a and b in
a group G is the element [a, b] := a−1 b−1 ab. The commutator group of G is
the normal subgroup of G which is generated by the set {[a, b] | a, b ∈ G}
of all commutators of G. In a sense, the commutator subgroup of a group G
measures how far the group is from being commutative. More generally, if
M and N are two normal subgroups of G then the commutator group of M
and N , written as [M, N ], is the normal subgroup of G generated by the set
{[a, b] | a ∈ M, b ∈ N }. It is well known that [M, N ] is the least normal sub-
group of G with the property that gh = hg for all g ∈ M/[M,N ] , h ∈ N/[M,N ] .
289
290 CHAPTER 12. TERM CONDITION AND COMMUTATOR
d2 , . . . , dn ∈ G
The next theorem shows that the property of being abelian can be charac-
terized by properties of the congruence lattice of the algebra A2 .
Since in this sense the condition (*) is equivalent to the term condition,
and (*) is also equivalent to the fact that 4A is a block of the congruence
h4A iConA2 , we have proved our proposition.
The basic term condition (TC) can be generalized to a term condition for
congruences, which we shall denote by (TCC).
tA (a, c2 , . . . , cn ) = tA (a, d2 , . . . , dn ) ⇒
A A
t (b, c2 , . . . , cn ) = t (b, d2 , . . . , dn ) (TCC)
is satisfied.
Lemma 12.1.5 The following two statements are equivalent for normal sub-
groups M and N of a group G:
Proof: (ii) ⇒ (i): Suppose that (ii) is satisfied, and let a ∈ M and b ∈ N .
Consider the term operation tG (x, y, z) := yxz over G. Since (e, b) ∈ θN and
(e, a), (a, e) ∈ θM , for e the identity element of the group, we have tG (e, e, a)
= a = tG (e, a, e); using the term condition (TCC) on this gives tG (b, e, a) =
tG (b, a, e), meaning ab = ba.
Proof: By Lemma 5.3.3, it is enough to show that for all a, b ∈ A and every
unary polynomial operation pθ on A × A we have
(1) For all n ∈ N, for all n-ary term operations tA of A and all (a, b) ∈ β
and (c2 , d2 ), . . . , (cn , dn ) ∈ α,
294 CHAPTER 12. TERM CONDITION AND COMMUTATOR
For a given α and β in ConA, we will consider the set of all δ ∈ ConA
satisfying (1). We note first that this set is non-empty, since for instance δ =
A × A satisfies (1). It is also true that this set is closed under intersection. If
the congruences δ1 and δ2 satisfy the implication of condition (1), for fixed
congruences α and β, then for any (a, b) ∈ β and (c2 , d2 ), . . . , (cn , dn ) ∈ α
we have
The following theorem connects the commutator with the term condition for
congruences.
G = D0 G ⊇ D1 G · · · ⊇ Di G ⊇ · · · (N ),
in which each Di+1 G is a normal subgroup of Di G and each quotient (or
factor) group Di G/Di+1 G is abelian. If this chain stops at some finite stage
n with Dn G equal to the trivial subgroup E of G, then the series (N) is called
a normal series with abelian factors, and G is said to be solvable. Thus a
group G is solvable if and only if there is a natural number n such that Dn G
= E.
To generalize this process for arbitrary algebras, we make the following def-
inition.
5◦A := 5A , 5k+1
A := [5kA , 5kA ],
(i) A is abelian;
(ii) A satisfies the term condition (1) from the beginning of Section 12.2;
(iii) The diagonal 4A is a block of a congruence relation on A × A;
(iv) [5A , 5A ] = 4A ;
(v) A is nilpotent of degree 1.
(i) A is abelian,
(ii) A is polynomially equivalent to a module over a ring.
Proof: (ii) ⇒ (i): it is an easy exercise to show that the polynomial oper-
ations of a module satisfy the term condition (TC) from Definition 12.1.1.
Therefore modules are abelian. This is also true for algebras which are poly-
nomially equivalent to a module over a ring.
x + y : = p (x, 0, y),
−x : = p (0, x, 0).
Then (A; +, −) is an abelian group.
Proof: We note first that 0 is a neutral element for the operation +, since
for any a ∈ A we have a + 0 = p(a, 0, 0) = a = p(0, 0, a) = 0 + a.
Next we use the fact that the term condition is satisfied for all polynomial
operations, including all operations built up from +, − and 0. Then for
any a ∈ A, we have p(0, 0, −a) = p(0, 0, p(0, a, 0)) = p(0, a, 0), and by ap-
plying the term condition (TC) to the equation p(0, 0, −a) = p(0, a, 0) we
can replace the first 0 by any element from A. Replacing 0 by a we ob-
tain p(a, 0, −a) = p(a, a, 0), and therefore a + (−a) = 0. Similarly we have
p(−a, 0, 0) = p(p(0, a, 0), 0, 0) = p(0, a, 0), and using the term condition to
replace 0 by a gives p(−a, 0, a) = p(0, a, a), and (−a) + a = 0. Here we are
also using commutativity, which follows from b = p(a, a, b) = p(b, a, a) by
the term condition if we replace a by 0.
f A (a1 + 0, . . . , an + 0) + f A (0, . . . , 0) =
f A (a1 , . . . , an ) + f A (0 + 0, . . . , 0 + 0)
f A (a1 + b1 , . . . , an + 0) + f A (0, . . . , 0) =
f A (0 + b1 , . . . , 0 + 0) + f A (a1 , . . . , an ),
298 CHAPTER 12. TERM CONDITION AND COMMUTATOR
f A (a1 + b1 , . . . , an + bn ) + f A (0, . . . , 0) =
f A (a1 , . . . , an ) + f A (b1 , . . . , bn ).
Proof: This follows from the fact that (R; +, −, 0, ◦) is a ring of endomor-
phisms of (A; +, −, 0).
Step 6. The polynomial operations of the algebra A are exactly the opera-
tions of the form f (x1 , . . . , xn ) = c + r1 x1 + · · · + rn xn , for n ∈ N, c ∈ A and
r1 , . . . , rn ∈ R.
Proof: It is clear that all operations of the given form are polynomial op-
erations of A. Assume that f A is a polynomial operation of A. From Step 2
we have
12.3. EXERCISES 299
f A (a1 , . . . , an ) = f A (a1 + 0, 0 + a2 , . . . , 0 + an )
= (f A (a1 , 0, . . . , 0) − f A (0, . . . , 0))+
f A (0, a2 , . . . , an )
..
.
= (f A (a1 , 0, . . . , 0) − f A (0, . . . , 0)) + · · ·
+(f A (0, . . . , 0, an ) − f A (0, . . . , 0))+
f A (0, . . . , 0).
Now the polynomial operations induced by the polynomials
12.3 Exercises
12.3.1. Verify that any left-zero, right-zero or zero-semigroup is abelian, as
claimed in Example 12.1.2 (2).
12.3.2. Prove that any subalgebra of an abelian algebra is abelian, and that
every direct product of abelian algebras is abelian.
12.3.3. Prove that the collection of all abelian algebras in a congruence per-
mutable variety forms a subvariety.
f a b c d
a b a c d
b a d b c
c d c a b
d c b d a
12.3.5. Complete the proof given in this chapter that a group is abelian iff
it is commutative. Prove that a ring is abelian iff it satisfies xy = 0.
Complete Sublattices
We have seen that the collection of all varieties of a given type forms a com-
plete lattice, as does the collection of all clones of operations defined on a
fixed set. These two lattices play an important role in universal algebra, but
their study is made difficult by the fact that the lattices are large (usually
uncountably infinite) and very complex. Thus we look for new approaches
or tools to use in their study. One such approach is to try to study some
smaller parts of the large lattice. Such smaller parts should have the same
algebraic structure, so we are interested in the study of complete sublattices
of a complete lattice.
In this chapter we describe some new methods for producing complete sub-
lattices of a given complete lattice. As we saw in Chapter 2, the two complete
lattices we are interested in both arise as the lattice of closed sets under a
closure operation, which can be obtained via a Galois-connection. This leads
us to the study of new closure operators and Galois-connections, which pro-
duce sublattices of the original lattice of closed sets. We develop the theory
of such sublattices in this chapter. In the next chapter we will apply this
general theory to our two specific examples of the lattices of varieties and
clones.
301
302 CHAPTER 13. COMPLETE SUBLATTICES
that any such closure system forms a complete lattice. In this lattice, the
meet operation, also the greatest lower bound or infimum with respect to
the partial order of set inclusion, is the operation of intersection. The join
operation however is not usually just the union: we have
_ \ [
B= {H ∈ Hγ | H ⊇ B}
for every B ⊆ Hγ . One situation when we do have the join operation equal
to union is the following.
We can show easily that when γ is an additive closure operator, the least
upper bound operation on the lattice Hγ agrees with B (see M. Reichel, [98]
S
all subsets of A.
Definition 13.1.2 Let γ1 be a closure operator defined on the set A and let
γ2 be a closure operator defined on the set B. Let R ⊆ A × B be a relation
between A and B. Then γ1 and γ2 are called conjugate with respect to R if
for all t ∈ A and all s ∈ B, γ1 (t) × {s} ⊆ R iff {t} × γ2 (s) ⊆ R.
(ii) µγ (T ) ⊆ µ(T ),
(iii) γ2 (µγ (T )) = µγ (T ),
Proof: We will prove only (i)-(v), the proofs of the other propositions being
dual.
(i) By definition,
µγ (T ) = {b ∈ B | ∀a ∈ T ((a, b) ∈ Rγ )}
= {b ∈ B | ∀a ∈ T (γ1 (a) × {b} ⊆ R)}
= {b ∈ B | ∀a ∈ γ1 (T ) ((a, b) ∈ R)} = µ(γ1 (T )).
(v) µγ (ιγ (S)) = µ(γ1 (ιγ (S))) = µ(ιγ (S)) = µ(ι(γ2 (S))).
The next theorem is our “Main Theorem for Conjugate Pairs of Closure
Operators.” It shows that when we consider sets which are closed under the
original Galois-connection from R, there are four equivalent conditions for
such sets to also be closed under the new connection from Rγ .
closure operators with respect to the relation R. Then for all sets T ⊆ A
with ι(µ(T )) = T the following propositions (i) - (iv) are equivalent; and
dually, for all sets S ⊆ B with µ(ι(S)) = S, propositions (i0 ) - (iv0 ) are
equivalent:
(ii) γ1 (T ) = T ,
(iii) µ(T ) = µγ (T ),
(i 0 ) S = µγ (ιγ (S)),
(ii 0 ) γ2 (S) = S,
Proof: We prove the equivalence of (i), (ii), (iii) and (iv); the equivalence
of the four dual statements can be proved dually.
(iv) ⇒ (i): Since the ιγ µγ -closed sets are exactly the sets of the form
ιγ (S), we have to find a set S ⊆ B with T = ιγ (S). But we have
ιγ (µ(T )) = ι(γ2 (µ(T ))) = ι(µ(T )) = T , by 13.1.4 (i0 ) and our assumption
that T is ιµ-closed.
Proof: We prove only (i0 ) and (ii0 ); the others are dual.
Now we are ready to produce our complete sublattices. We know that from
the original relation R and Galois-connection (µ, ι) we have two (dually
isomorphic) complete lattices of closed sets, the lattices Hµι and Hιµ . We
also get two complete lattices of closed sets from the new Galois-connection
(µγ , ιγ ) induced by Rγ . Our result is that each new complete lattice is in
fact a complete sublattice of the corresponding original complete lattice.
operators with respect to R. Then the lattice Hµγ ιγ of sets closed under µγ ιγ
is a complete sublattice of the lattice Hµι , and dually the lattice Hιγ µγ is a
complete sublattice of the lattice Hιµ .
Proof: As a closure system Hµγ ιγ is a complete lattice, and we have to
prove that it is a complete sublattice of the complete lattice Hµι . We begin
by showing that it is a subset. Let S ∈ Hµγ ιγ , so that µγ (ιγ (S)) = S. Then
µ(ι(S)) = µ(ι(µγ (ιγ (S)))) = µ(ι(µ(ι(γ2 (S))))) = µ(ι(γ2 (S))) = µγ (ιγ (S)) =
S by 13.1.4 (v), and thus S ∈ Hµι . This shows Hµγ ιγ ⊆ Hµι . Since every S
in Hµγ ιγ satisfies µ(ι(S)) = S, we can apply Theorem 13.1.5 (ii0 ), to get
S ∈ Hµγ ιγ ⇐⇒ S = µγ (ιγ (S)) ⇐⇒ S = γ2 (S) ⇐⇒ S ∈ Hγ2 .
As we remarked after Definition 13.1.1, the fact that γ2 is an additive closure
operator means that the corresponding closure system is a complete sublat-
tice of the lattice (P(B); ∩, ∪) of all subsets of B; that is, on our lattice
Hµγ ιγ the meet operation agrees with ordinary set-intersection and the join
agrees with union. We already know that the meet operation in Hµι also
agrees with intersection, so we only need to show that Hµγ ιγ is closed under
the join operation of Hµι . Let (Sk )k∈J be an indexed family of subsets of B.
Then
W W S
µγ (ιγ ( Sk )) = γ2 ( Sk ) = γ2 (µ(ι( Sk )))
k∈J S k∈J S k∈J S W
= γ2 (µ(ιγ ( Sk ))) = µ(ιγ ( Sk )) = µ(ι( Sk )) = Sk ,
k∈J k∈J k∈J k∈J
1) R0 ⊆ R, and
13.2. GALOIS CLOSED SUBRELATIONS 309
We leave it as an exercise for the reader to verify, using the results of Section
13.1, that if γ := (γ1 , γ2 ) is a pair of additive closure operators which are
conjugate with respect to a relation R ⊆ A × B, then the relation Rγ of
Definition 13.1.3 is a Galois-closed subrelation of R.
Before we can prove our main theorem, we need the following well-known
result for Galois-connections. The proof is straightforward, and is left as an
exercise for the reader (see Exercise 2.4.2).
S T
(i) µ( Tj ) = µ(Tj ),
j∈J j∈J
S T
(ii) ι( Sj ) = ι(Sj ).
j∈J j∈J
is a Galois-closed subrelation of R.
ι µ(T ) = ι µ0 (T ) = ι0 µ0 (T ) = T.
Therefore, Hι0 µ0 ⊆ Hι µ .
Next we have to show that this subset is in fact a sublattice. This means
showing that for any family {Tj | j ∈ J} of sets in Hι0 µ0 , both the sets
{Tj | j ∈ J} and {Tj | j ∈ J} are in Hι0 µ0 .
V W
Hιµ Hι µ
We start with the meet operation. We know from above that the collection
{Tj | j ∈ J} is also a family of sets in Hι µ . Since
^ \ \
{Tj | j ∈ J} = Tj = ιµ( Tj ),
Hι µ j∈J j∈J
we have
ι0 µ0 (Tj )).
\ \
Tj = ιµ(
j∈J j∈J
Applying Lemma 13.2.3 to this, and then using Proposition 13.2.2 (iii) twice,
we get
Using the closure operator properties, and Lemma 13.2.3 once more, we get
ι0 µ0 ( Tj ) = ι0 µ0 (
T 0 0
ι µ (Tj )) = ι0 µ0 ι0 (
T S 0
µ (Tj ))
j∈J
S j∈J0 j∈J
= ι0 ( µ (Tj )) ⊆ ι0 µ0 (Tj ) = Tj ,
j∈J
312 CHAPTER 13. COMPLETE SUBLATTICES
for all j ∈ J.
Thus we have ι0 µ0 ( Tj ) ⊆
T T
Tj . The reverse inclusion is always true for
j∈J j∈J
a closure operator, so altogether we have ι0 µ0 (
T T
Tj ) = Tj . This shows
j∈J j∈J
Tj ∈ Hι0 µ0 .
T
that
j∈J
The proof that (*) holds will be divided into a number of steps. We begin
with two facts we shall need.
µ0 (T ) = {s ∈ B | ∀ t ∈ T, (t, s) ∈ RU }.
(a) Dt 6= ∅.
(c) ι0 µ0 ({t}) = Dt .
314 CHAPTER 13. COMPLETE SUBLATTICES
S
(d) T = Dt .
t∈T
T = ι0 µ0 (T ) = ι0 µ0 (
[
{t})
t∈T
⊇ ι0 µ0 ({t}) for all t ∈ T
= Dt for all t ∈ T,
[ \
µ(T ) = µ( Dt ) = µ(Dt )
t∈T t∈T
0 0
Dt ) = µ0 (T ) = S.
\ [
= µ (Dt ) = µ (
t∈T t∈T
From Fact 2, we have ι(S) = ι0 (S) = ι0 µ0 (T ) = T . This shows that (*) holds,
completing the proof of (ii) that RU is a Galois-closed subrelation of R.
13.2. GALOIS CLOSED SUBRELATIONS 315
(iii) Now we must show that for any complete sublattice U of Hι µ , and any
Galois-closed subrelation R0 of R, we have URU = U and RUR0 = R0 .
We know that URU := Hι0 µ0 , the lattice of subsets of A closed under the
closure operator ι0 µ0 induced from the relation RU . This means that T ∈ URU
iff ι0 µ0 (T ) = T . First let T ∈ URU , and let S be the set µ0 (T ). Then we have
ι0 (S) = T , and since RU is a Galois-closed subrelation of R we conclude that
For the opposite inclusion, let T ∈ U. Then using the fact that U is a
sublattice of Hι µ , along with Fact 2, we have
ι0 µ0 (T ) = ι0 µ(T ) = ι µ(T ) = T.
This shows that T ∈ Hι0 µ0 , which is equal to URU . We now have the required
equality URU = U.
To show the opposite inclusion, let T ∈ UR0 , and let S = µ(T ). Then from
Facts 1 and 2 we have
µ0 (T ) = µ(T ) = S and ι0 (S) = ι(S) = ι µ(T ) = T.
Therefore T × µ(T ) ⊆ R0 , and RUR0 ⊆ R0 . Altogether, we have RUR0 = R0 .
This completes the proof of part (iii), and of Theorem 13.2.4.
316 CHAPTER 13. COMPLETE SUBLATTICES
As before, we start with a relation R between two sets A and B, with induced
Galois-connection (µ, ι) and corresponding closure operators ιµ on A and µι
on B. We denote by Hιµ and Hµι the corresponding complete lattices of
closed sets on A and B respectively. Now we assume that γ1 : P(A) → P(A)
and γ2 : P(B) → P(B) are additive closure operators which are conjugate
with respect to the relation R. As we saw in Section 13.1, this conjugate pair
determines a new relation Rγ from A to B, with its own induced Galois-
connection (ιγ , µγ ) and closure operators ιγ µγ and µγ ιγ . We now have three
lattices of closed sets on A, corresponding to closure under the operators
ιµ, ιγ µγ and γ1 , and dually three lattices on B. From Theorems 13.1.5 and
13.1.7 we have the following connection between these lattices:
In this section we examine how the lattice Sγ1 is situated in the lattice Hιµ .
In particular, for any set T ∈ Hιµ we can look for the least γ1 -closed class
γ1 T containing T and the greatest γ -closed class
1 γ1 T contained in T . Thus
we consider two operators Γ1 : T 7→ γ1 T and Γ2 : T 7→ γ1 T , whose properties
will be studied in more detail. We will present our definitions and results for
the lattices Hιµ and Sγ1 on A, but of course these results can be dualized
for the corresponding lattices on B as well.
{T 0 ∈ Sγ1 | T 0 ⊇ T }
\
γ1
T :=
Sγ1 Hιµ
{T 0 ∈ Sγ1 | T 0 ⊆ T } = {T 0 ∈ Sγ1 | T 0 ⊆ T } .
_ _
γ1 T :=
(Note that the last equality holds because Sγ1 is a complete sublattice of
Hιµ , as we proved in Theorem 13.1.7.)
13.3. CLOSURE OPERATORS ON COMPLETE LATTICES 317
Lemma 13.3.2 For any sets T and T 0 in Hιµ and any set S in Hµι ,
(ii) Let T ∈ Hιµ and S ∈ Hµι . If S = µ(T ) ∈ Sγ2 , then ι(S) = ιµ(T ) = T
and also ιγ2 (S) = ι(S). This gives T = ιγ2 (S). Applying µ to both sides and
using 13.1.4 (i0 ) and 13.2.2 (iii), we get
Now we apply ι to both sides of this result, to get ιµγ ιγ (S) = ι(S). Finally
we have
(a) γ1 T
γ1 T = T iff = T iff γ1 (T ) = T ,
(b) γ1 T is the greatest ιγ µγ -closed set contained in T , and
(c) γ1 T = ιγ µ(T ) = ι γ2 µ(T ).
Proof: (i) This follows from the definition of γ1 T , Proposition 13.2.2 (iii)
and Proposition 13.1.4 (i).
(ii) (a) The first equivalence follows from the definitions, while the second
one follows directly from Theorem 13.1.5 and the fact that Sγ1 = Hιγ µγ .
318 CHAPTER 13. COMPLETE SUBLATTICES
= ιγ µγ ιγ µ(T ), by 13.2.2
= ιγ µ(T ).
For any set T in the lattice Hιµ , let us denote by [γ1 T, γ1 T ] the interval
between γ1 T and γ1 T in Hιµ . Such intervals will be called γ1 -intervals in
Hιµ . It is possible that different sets T may produce the same γ1 -interval.
This suggests that we define an equivalence relation ∼ on Hιµ , by T1 ∼ T2
: ⇐⇒ [γ1 T1 , γ1 T1 ] = [γ1 T2 , γ1 T2 ].
(ii) T1 ⊆ T2 implies ⊆ γ1 T2 γ1 T ⊆ γ1 T2 ,
γ1 T1 and 1
Hιµ Hιµ
(iii) γ1 T1 ∧ γ1 T2 = γ1 (T1 ∧ T2 ),
Hιµ Hιµ
(iv) γ1 T ∨ γ1 T = γ1 (T ∨ T2 ).
1 2 1
(iii) Since Sγ1 is a complete sublattice of Hιµ by Theorem 13.1.7, the set on
the left hand side is an element of Sγ1 , and it is contained in both T1 and
T2 . Therefore it is also contained in the greatest set from Sγ1 to contain T1
and T2 , which is the set on the right hand side of (iii). Conversely, the set
on the right hand side of (iii) is by definition an element of Sγ1 , and by part
(ii) it is contained in both γ1 T1 and γ1 T2 . Thus it is also contained in the set
on the left hand side.
Hιµ Hιµ
_ _
γ1
( {Tj | j ∈ J}) = {γ1 Tj | j ∈ J}.
The two equalities are generalizations of parts (iv) and (iii) of Proposition
T HV
ιµ
13.3.4, and may be proved in the same manner. Note that equals in
the lattice Hιµ .
Remark 13.3.6 From the definitions and Proposition 13.3.4 parts (iii) and
(iv), we conclude that the operators Γ1 and Γ2 have the following properties:
320 CHAPTER 13. COMPLETE SUBLATTICES
(i) The mapping Γ1 is a join-retraction from Hιµ onto Sγ1 ⊆ Hιµ ; that is, it
is an idempotent join-homomorphism which is the identity map on Sγ1 .
(iii) Note that, in general, Γ1 does not preserve meets and Γ2 does not
preserve joins.
S := F ix(ϕ) := Hϕ = {T ∈ L | ϕ(T ) = T }
of all fixed points of ϕ; and for any closure system S we have the closure
operator ϕ defined by
L
{T 0 ∈ S | T ≤ T 0 } for T ∈ L.
V
ϕ(T ) = ϕS (T ) :=
Moreover, for any closure system S and any closure operator ϕ, we have HϕS
= S and ϕHϕ = ϕ.
There is also a dual 1-1 correspondence between kernel operators and ker-
nel systems on a complete lattice L. A kernel system on L is a family S of
subsets of L which is closed under arbitrary joins. Then for kernel operators
ψ : L → L on L and kernel systems S on L, we set
S := F ix(ψ) := {T ∈ L | ψ(T ) = T } ;
L
{T 0 ∈ S | T 0 ≤ T } for T ∈ L.
W
ψ(T ) := ψS (T ) :=
We have F ix(ψS ) = S and ψF ix(ψ) = ψ, for any kernel system S and any
kernel operator ψ on L.
A result of A. Tarski ([111]) shows that for any closure operator ϕ on a com-
plete lattice L, the closure system (fixed-point set) Hϕ is always a complete
13.3. CLOSURE OPERATORS ON COMPLETE LATTICES 321
for every index set J, then the set of all fixed points under ϕ,
Hϕ = {T ∈ L | ϕ(T ) = T },
for every index set J, then the set of all fixed points under ψ,
Hψ = {T ∈ L | ψ(T ) = T },
(ii) By Remark 13.3.6, we need only show that the closure operator defined
by
L
{T 0 ∈ H | T ≤ T 0 }
^
ϕH (T ) :=
satisfies condition (*) and that ϕH (L) = H. We prove the latter fact first.
Since H is a complete sublattice of L, we have ϕH (L) ⊆ H. For the opposite
inclusion, we see that for any T ∈ H
L
{T 0 ∈ H | T ≤ T 0 } = T.
^
ϕH (T ) =
Thus H ⊆ ϕH (L), and altogether we have H = ϕH (L).
Since for each j ∈ J we have Tj ≤ ϕH (Tj ) and ϕH (Tj ) ∈ H, the set
L
_
{ϕH (Tj ) | j ∈ J}
is an upper bound of the set {Tj ∈ L | j ∈ J}. Therefore
L
_ L
_
{Tj ∈ L | j ∈ J} ≤ {ϕH (Tj ) | j ∈ J}.
Since the set on the right hand side of this inequality is an element of H,
applying ϕH on both sides gives
L
_ L
_
ϕH ( {Tj ∈ L | j ∈ J}) ≤ {ϕH (Tj ) ∈ L | j ∈ J}.
13.4. EXERCISES 323
L L
{Tj ∈ L | j ∈ J} ≥ Tj , we have ϕH ( {Tj ∈ L | j ∈ J}) ≥ ϕH (Tj )
W W
Since
L L
for all j ∈ J. Thus also ϕH ( {Tj ∈ L | j ∈ J}) ≥ {ϕH (Tj ) | j ∈ J}, giving
W W
(iii), (iv) These proofs are analogous to those of (i) and (ii).
The equations follow from Remark 13.3.6, by restricting the one-to-one map-
ping between closure operators and complete lattices to closure operators
satisfying condition (*) and to complete sublattices.
We can apply this Theorem to the special case of conjugate pairs of closure
operators studied in Section 13.1. Using the notation from 13.1 and applying
13.3.5 part (i), we take L = Hιµ , and ϕ = Γ1 (= ιγ µγ , as in 13.3.3 (i)) and
ψ = Γ2 (= ιγ µ, as in 13.3.3 (ii)). This gives an additional proof of the fact
that Sγ1 = Hιγ µγ = F ix(ϕ) = F ix(ψ) is a complete sublattice of Hιµ .
13.4 Exercises
13.4.1. Let R be a relation between sets A and B. Prove that the closure
operators µι and ιµ obtained from the Galois-connection (µ, ι) induced by
R are conjugate with respect to R.
13.4.2. Prove the additional properties for conjugate pairs of additive closure
operators listed at the end of Section 13.1.
13.4.4. Prove that for an additive closure operator γ, the least upper bound
operation on the lattice Hγ agrees with the union operation.
(i) F = A, and
S
Prove that the family of closed sets of a partial closure operator on a set
A is a partial closure system on A, and conversely that for every partial
closure system F on A there is a partial closure operator on A whose family
of closed sets is exactly F. (See B. Šešelja and A. Tepavčević, [106].)
14.1 G-Clones
In this section we apply our theory of Galois-closed subrelations to the lattice
of clones on a fixed set. We assume a fixed base set A, and denote by O(A)
the set of all finitary operations on A and by R(A) the set of all finitary
relations on set A. As our basic relation between these two sets we have the
relation R of preservation:
325
326 CHAPTER 14. G-CLONES AND M -SOLID VARIETIES
for all a1 , . . . , an ∈ A.
We use this to define, for any fixed subgroup G ⊆ SA of permutations, a
mapping γG O on operations and sets of operations. For any operation f and
This gives a map γG O on the power set of O(A), which is our first candidate
We can use the closure operator γGO to define another relation R between
G
operations and relations on A, by setting
O
RG := {(f, ρ) | f ∈ O(A), ρ ∈ R(A) and γG (f ) × {ρ} ⊆ R}.
Lemma 14.1.5 For any f ∈ O(A), for any ρ ∈ R(A) and for any subgroup
G ⊆ SA , we have:
O R
γG (f ) preserves ρ iff f preserves γG (ρ).
way.
R
RG = {(f, ρ) | f ∈ O(A), ρ ∈ R(A) and {f } × γG (ρ) ⊆ R}
Dually, the set of all G-relational clones forms a complete sublattice of the
lattice of all relational clones.
Proof: The fact that we get complete sublattices, of the lattices of clones and
of relational clones respectively, comes from Theorem 13.1.7. If G ⊆ G 0 ⊆ SA
are subgroups, then clearly γG O (F ) ⊆ γ O (F ) for every F ⊆ O(A), and if
G00
γG0 (F ) = F then also γG (F ) = F , so LG
O O G
A ⊆ LA . Moreover RG0 is a Galois-
0
closed subrelation of RG , so that LG G
A is a complete sublattice of LA .
Proposition 14.1.8 For all F ⊆ O(A) and all Q ⊆ R(A), the following
properties hold:
R (Q),
(i) GP olA Q = P olA γG
(ii) GP olA Q ⊆ P olA Q,
(iii) γGO (GP ol Q) = GP ol Q,
A A
O (F );
(iv) GP olA GInvA F = P olA InvA γG and dually,
(i0 ) O (F ),
GInvA F = InvA γG
(ii0 ) GInvA F ⊆ InvA F ,
(iii0 ) R (GInv F ) = GInv F ,
γG A A
(iv0 ) R (Q).
GInvA GP olA Q = InvA P olA γG
The Main Theorem for Conjugate Pairs of additive closure operators, The-
orem 13.1.5, gives us a characterization of G-clones and G-relational clones.
Proof: These conditions all come from Theorem 13.1.5, except for (iv) and
(iv0 ), which are simply applications of (ii0 ) and (ii).
We also have the following conditions which can be derived from Theorem
13.1.6 (see also K. Denecke and M. Reichel, [35]).
By checking the generating systems, it can be verified that all of these clones
are S2 -clones. That these are all the S2 -clones was proved by Gorlov and
Pöschel in [50].
Theorem 14.1.11 ([50]) There are exactly fourteen S2 -clones on the two-
element set A = {0, 1}:
O1 , O4 , O8 , O9 , L1 , L4 , L5 , D2 , D1 , D3 , A4 , A1 , C4 and C1 .
14.2 H-clones
In [50], V. V. Gorlov and R. Pöschel described several generalizations of
G-clones to H-clones, where H is a transformation monoid. We will use a
new approach here, by applying a closure operator on the lattice of all clones
which is different from that given in Section 14.1.
Let TA = (O1 (A), ◦, ϕid ) be the monoid of all unary mappings or transfor-
mations on A, where ◦ is the composition of unary operations and ϕid is the
identity mapping on A.
332 CHAPTER 14. G-CLONES AND M -SOLID VARIETIES
C1
u
XXXXX
uA1
X
C4u u u
D3 L1
u D1u L5u O9 uX
A4
XXX
XXu
! O8
!!
!
D2 u u O4u
!
Z L4 ! !!
Z !!
Z !!
Z !!
!
u
Z
!
Z!
O1
For every unary mapping ϕ ∈ O1 (A) and every n-ary mapping f ∈ On (A),
we define a new mapping f ϕ , by setting
for all (a1 , . . . , an ) ∈ An . We use this to define, for any set H of unary
mappings, an operator on individual mappings and on sets F ⊆ O(A) of
mappings, by
[
O
γH (f ) := {f ϕ | ϕ ∈ H} and γH
O
(F ) := O
γH (f ).
f ∈F
closed under this Galois-connection, that is the clones F such that γHO (F ) =
H
F , are called H-clones. We will denote by LA the lattice of all H-clones on
A. Since we do not have a conjugate pair of additive closure operators here,
it is not clear whether this lattice forms a complete sublattice of the lattice
LA . It is always at least a meet-subsemilattice of LA .
T{0,1}
s
H4 s
H3 s s H2 s H5
s
H1
(iii) LH 3
A is the set of all clones containing the constant c1 .
(iv) LH 4
A is the set of all clones containing both constants c0 and c1 .
It turns out that all of these lattices are complete sublattices of LA , for
A = {0, 1}. The following picture shows the structure of the set of all these
lattices.
LA
s
LH 3 s s H2 sLH5
A L A A
LH
A
4
s
T
LA{0,1}
of algebras. On the algebra side, the closed sets are the equational classes
or varieties, and we have the complete lattice L(τ ) of all varieties of type τ .
Dually, the closed sets on the identity side are the equational theories, and
we have the complete lattice E(τ ) of all equational theories of type τ . These
two lattices L(τ ) and E(τ ) are dually isomorphic, and in general are very
large and complex. Thus it is important to find some means of studying at
least portions of these lattices, such as complete sublattices.
Our goal is to introduce two new closure operators on our sets A and B,
which we shall show form a conjugate pair of additive closure operators. The
results of Section 13.1 then give us complete sublattices of our two lattices.
The new operators we use are based on the concept of hypersatisfaction of
an identity by a variety. We begin with the definition of a hypersubstitution,
as introduced by Denecke, Lau, Pöschel and Schweigert in [32]. A complete
study of hypersubstitutions and hyperidentities may be found in [37].
Here the left side of (ii) means the composition of the term σ(fi ) and the
terms σ̂[t1 ], . . . , σ̂[tni ].
σ1 ◦h σ2 := σ̂1 ◦ σ2 ,
where ◦ denotes ordinary composition of functions. We will show that this
operation is associative, and that the set of all hypersubstitutions forms a
monoid. The identity element is the identity hypersubstitution σid , which
maps every fi to fi (x1 , . . . , xni ).
(ii) Let σ1 , σ2 and σ3 be any three elements of Hyp(τ ). Then from (i) we
have σ1 ◦h (σ2 ◦h σ3 ) = σ̂1 ◦ (σ̂2 ◦ σ3 ) = (σ̂1 ◦ σ̂2 ) ◦ σ3 = (σ̂1 ◦ σ2 )ˆ◦ σ3 =
(σ1 ◦h σ2 ) ◦h σ3 .
The Galois-closed classes of algebras under this connection are precisely the
M -solid varieties of type τ , which then form a complete sublattice of the
14.3. M -SOLID VARIETIES 337
lattice of all varieties of type τ . Thus studying M -solid and solid varieties
gives a way to study complete sublattices of the lattice of all varieties of a
given type.
We now introduce some closure operators on the two sets Alg(τ ) and
Wτ (X)2 . On the equational side, we can use the extensions of our M -
hypersubstitutions to map any terms and identities to new ones. That is,
we define an operator χE
M by
χE
M [u ≈ v] = {σ̂[u] ≈ σ̂[v] : σ ∈ M }.
This extends, additively, to sets of identities, so that for any set Σ of iden-
tities we set
χE E
M [Σ] = {χM [u ≈ v] : u ≈ v ∈ Σ}.
χA
M [A] = {σ[A] : σ ∈ M}, and
A A
χM [K] = {χM [A] : A ∈ K}.
with respect to satisfaction. For this we need to show that for any algebra
A and any identity u ≈ v of type τ , we have
χA E
M [A] satisfies u ≈ v iff A satisfies χM [u ≈ v].
Once we know that our two additive closure operators form a conjugate pair,
we can apply our Main Theorem for such conjugate pairs, Theorem 13.1.5.
Translating that theorem into the specific case here, we have the following
description of the closed objects.
(ii) χA
M [V ] = V .
(iv) χE
M [IdV ] = IdV .
And dually, for any equational theory Σ of type τ , the following conditions
are equivalent:
(ii0 ) χE
M [Σ] = Σ.
(iv0 ) χA
M [M odΣ] = M odΣ.
Since for any variety V we have HM IdV ⊆ IdV , condition (iii) of this
theorem says that V is an M -solid variety, since every identity of V is an
M -hyperidentity. In analogy with the (Id, M od) case, a variety which sat-
isfies condition (i) is called an M -hyperequational class. Thus our M -solid
14.3. M -SOLID VARIETIES 339
varieties are precisely the M -hyperequational classes. Dual results hold for
M -hyperequational theories, using the second half of the theorem. In ad-
dition, this tells us that the relation of hypersatisfaction is a Galois-closed
subrelation of the satisfaction relation. Moreover, from Theorem 13.1.7 we
have the following result.
When V is a variety of type τ , we can form the lattice L(V ) of all subvarieties
of V (see Section 6.6). Then the intersection SM (V ) := SM (τ ) ∩ L(V ) is the
lattice of all M -solid subvarieties of V . Such lattices have been investigated
for a number of choices of V and M , but most work has been done for the
case that V is the type (2) variety Sem of all semigroups. We give here some
examples of results in this direction.
Theorem 14.3.7 ([91]) The largest solid variety of semigroups is the vari-
ety
One direction of this theorem is easy to prove. The variety VHS must of
course satisfy the associative law, and by applying the hypersubstitutions
taking the binary operation symbol f to the four semigroup terms x2 , xyx,
x2 y and xy 2 we get the other four identities in the claimed basis. This shows
that VHS is contained in the model class of the set of five identities given
in the theorem. The proof of the other direction involves showing that the
variety defined by these five identities is indeed hyperassociative, and is too
complex for us to give here.
Using this equational basis for the largest solid variety of semigroups, L.
Polák also gave a characterization of all solid semigroup varieties.
Theorem 14.3.9 ([27] (i) The largest presolid but not solid variety of semi-
groups is the variety
We also set
(iii) It is clear that for intervals, [ χM1 V, χM1 V ] ⊆ [ χM2 V, χM2 V ] iff
χM1
χM1 V ⊇ χM2 V and V ⊆ χM2 V ; and then we apply (i) and (ii).
By Lemma 14.4.3, the map ϕ from the lattice Sub(Hyp(τ )) of all submonoids
of Hyp(τ ) to the lattice of all varieties of type τ , which associates to each
344 CHAPTER 14. G-CLONES AND M -SOLID VARIETIES
= ϕ(M1 ) ∨ ϕ(M2 ).
Similarly, we have
14.5 Exercises
14.5.1. Prove Lemma 14.2.1.
14.5.2. Prove that for any two hypersubstitutions σ and ρ of a fixed type τ ,
we have (σ ◦h ρ)ˆ= (σ̂ ◦ ρ)ˆ= σ̂ ◦ ρ̂.
Hypersubstitutions and
Machines
347
348 CHAPTER 15. HYPERSUBSTITUTIONS AND MACHINES
Lemma 15.1.2 ([28]) Let τ = (ni )i∈I be a type of algebras with ni ≥ 1 for
all i ∈ I. Let σ be a hypersubstitution of type τ . Then the relation kerσ is a
fully invariant congruence on the absolutely free algebra Fτ (X).
In the case of type (n), for n ≥ 1, Denecke, Koppitz and Niwczyk have
completely determined all the congruence relations kerσ for any hypersub-
stitution σ of type (n). To describe their results we need some notation for
terms regarded as trees, from Section 5.1. To each node or vertex of a tree
representing a term of type (n) we can assign a sequence of integers from
the set {1, 2, . . . , n}, called the address of the node or vertex. For each value
1 ≤ i ≤ n, there is a uniquely determined variable obtained by following the
address ii · · · i in the tree until we have the address of a variable; we shall de-
note this variable by vari (t). Let K be any non-empty subset of {1, 2, . . . , n}.
An address is called a terminating K-sequence for t if it is the address of a
15.2. HYPER TREE RECOGNIZERS 349
Proposition 15.1.3 ([28]) Let τ = (n), with n ≥ 1, with one n-ary opera-
tion symbol f .
(i) If σ is a regular hypersubstitution of type τ , then kerσ is the diagonal
relation on Wτ (X); that is, σ̂[u] = σ̂[v] iff u = v.
(ii) If σ is a projection hypersubstitution, so that σ(f ) = xi for some
1 ≤ i ≤ n, then kerσ = {(u, v) ∈ Wτ (X)2 | vari (u) = vari (v)}.
(iii) Let n ≥ 2, and let σ be a non-projection hypersubstitution. Let K be
the set of variables used in the term σ(f ), with 1 ≤ |K| < n. Then a pair
(u, v) from Wτ (X)2 is in kerσ iff any terminating K-sequence of u is a ter-
minating K-sequence of v and vice versa, and any such sequence addresses
the same variable in both u and v.
In the same paper, the authors also extended their results to an arbitrary
type τ having no nullary operation symbols, for certain restricted kinds of
hypersubstitutions.
We begin by recalling some notation from Section 8.4, where algebras were
described by a set Σ of operation symbols rather than a type τ . That is,
we let Σ = Σ0 ∪ Σ1 ∪ · · · ∪ Σm be a set of operation symbols, where the
operation symbols in each Σn are n-ary. We will also assume here that Σ
is finite. As usual we have a countably infinite set X of variables, while for
each natural number n we let Xn be the set of variables x1 , . . . , xn . We will
denote by WΣ (X) and WΣ (Xn ) the sets of all terms which can be built up
from the operation symbols from Σ and the variables from X or from Xn ,
respectively. We write A = (A, ΣA ) for a (finite) algebra whose fundamental
operations correspond to the operation symbols from Σ (of the type of Σ).
As we saw in Chapter 5, in the case that both Σ and X are finite we can
consider terms as trees.
σ A [xi ] := en,A
i if xi ∈ Xn is a variable, and
σ A (fi ) = tAi for an ni -ary operation symbol fi ∈ Σni and an ni -ary term
A
operation ti of A.
(Note that n has to be greater than the greatest arity of the operation
symbols in Σ.) CnA is a subset of clonen A.
The mapping σ A thus maps each variable and each operation symbol to
an n-ary term of the algebra A. We mention that any such mapping can
15.2. HYPER TREE RECOGNIZERS 351
= {e2,A 2,A A
1 , e2 , (x1 ∧ x2 ) }. We consider the hyperrecognizer HA =
(clone2 A, Σ, X2 , σ , Cn ) in which CnA = {(x1 ∧ x2 )A } and σ A is given by
A A
(ii) for each t ∈ WΣ (Xn )\T and each sA ∈ CnA , we have σ̂[t] ≈ s ∈
/ IdA.
Proof: Suppose that T is hyperrecognizable. Then there exists a Σ−Xn -tree-
hyperrecognizer HA = (clonen A, Σ, Xn , σ A , CnA ) such that T (HA) = T .
If t ∈ T then there is an element sA ∈ CnA with (σ A )ˆ[t] = sA . But then
there is a hypersubstitution σ with (σ A )ˆ[t] = (σ̂[t])A . (Note that σ is not
uniquely determined by σ A , since every σ 0 for which σ(fi ) ≈ σ 0 (fi ) ∈ IdA
for every operation symbol fi from Σ satisfies the same equation.) From the
last equation we obtain σ̂[t] ≈ s ∈ IdA.
Conversely, suppose now that there is a finite algebra A = (A; ΣA ) with the
finite n-clone clonen A and a hypersubstitution satisfying (i) and (ii). We
will show that the hyperrecognizer HA = (clonen A, Σ, Xn , σ A , CnA ) satisfies
T (HA) = T . If t ∈ T (HA) then there is a term operation sA ∈ CnA such
that (σ A )ˆ[t] = sA , i.e., (σ[t])A = sA and σ̂[t] ≈ s ∈ IdA. Because of (ii) we
15.2. HYPER TREE RECOGNIZERS 353
The proof of Proposition 15.2.5 (i) shows that the term s which satisfies
σ̂[t] ≈ s ∈ IdA for t ∈ T is not uniquely determined. Therefore we use the
following binary relation ∼V (A) defined by J. Plonka in [90] on the set of all
hypersubstitutions:
(ii) Let A1 and A2 be Σ-algebras with IdA1 = IdA2 and let HA1 :=
(clonen A1 , Σ, Xn , σ A1 , CnA1 ) be a tree-hyperrecognizer based on the al-
gebra A1 . Then there exists a tree-hyperrecognizer T (HA2 ) based on
A2 with T (HA1 ) = T (HA2 ).
If t ∈ WΣ (Xn )\T (HA1 ) then for each sA ∈ CnA we have σˆ1 [t] ≈ s ∈
/ IdA. But
then also σˆ2 [t] ≈ s ∈
/ IdA, and by Proposition 15.2.5 we have t ∈ T (HA2 ).
A dual argument shows that T (HA2 ) ⊆ T (HA1 ).
(ii) Let V (A1 ) and V (A2 ) be the varieties generated by A1 and by A2 , re-
spectively and let FV (A1 ) (Xn ) and FV (A2 ) (Xn ) be the free algebras relative
to V (A1 ) and to V (A2 ), respectively. We have seen that the clone of an al-
gebra can be regarded as a multi-based algebra where the m-ary operations
for all 0 ≤ m ≤ n are the different sorts and where the operations are the
superposition operations. It is also well-known that clonen A1 is isomorphic
to the clone of the free algebra FV (A1 ) (Xn ). Here we have IdA1 = IdA2 ,
which tells us that FV (A1 ) (Xn ) = FV (A2 ) (Xn ) and therefore the clones are
354 CHAPTER 15. HYPERSUBSTITUTIONS AND MACHINES
0
equal. But then clonen A1 and clonen A2 are isomorphic. Let CnA2 be the im-
0
age of CnA1 under this isomorphism and let σ A2 be the composition of σ A1
with this isomorphism. Using this mapping it can be shown that T (HA1 ) =
T (HA2 ).
Now we are ready to prove that for finite alphabets the concepts of recog-
nizability and hyperrecognizability are equivalent.
HA = (clonen A, Σ, Xn , σ A , CnA )
Therefore T (HA) = m
S
i=1 T (HAi ).
Consider now an n-ary term t ∈ T (HAi ). Then we have
t ∈ T (HAi ) ⇒ (σ A )ˆ[t] = sA i
⇒ (σ̂[t])A = sA i if σ is a hypersubstitution
satisfying (σ A )ˆ[t] = (σ̂[t])A
for all t ∈ WΣ (Xn )
⇒ (σ̂[t])A (a1 , . . . , an ) = sA
i (a1 , . . . , an )
for all a1 , . . . , an ∈ A.
t ∈ T (HAi ) ⇔ (σ A )ˆ[t] = sA i
⇔ (σ̂[t])A = sA i
⇔ α̂il [σ̂[t]] = α̂il [si ] for all l = 1, . . . , ki
⇔ (α̂il ◦ σ̂)[t] = cil for all l = 1, . . . , ki
⇔ t ∈ T (Ail ) for all l = 1, . . . , ki
⇔ t ∈ kl=1
T i
T (Ail ),
Tk i
showing that T (HAi ) = l=1 T (Ail ).
(i) σ̂[x] := x,
358 CHAPTER 15. HYPERSUBSTITUTIONS AND MACHINES
Tree transducers were defined in Section 8.8. It turns out that tree transfor-
mations Tσ for a hypersubstitution σ are induced by tree transducers. The
following proposition is a special case of 8.8.5, since the extensions of hyper-
substitutions are a particular kind of tree homomorphisms as introduced in
Section 8.6.
In the remainder of this section we assume that we have only one type τ . We
denote by Tσ1 ◦ Tσ2 the composition of the tree transformations Tσ1 and Tσ2 .
We can also consider inverses, domains and ranges of tree transformations.
We define THyp(τ ) := {Tσ | σ ∈ Hyp(τ )}.
To see that ϕ is one-to-one, let Tσ1 = Tσ2 . Then for all t ∈ Wτ (X) we have
σ̂1 [t] = σ̂2 [t]. But this means that for all operation symbols fi we also have
σ̂1 [fi (xi , · · · , xni )] = σ1 (fi ) = σ2 (fi ) = σ̂2 [fi (xi , · · · , xni )],
and therefore σ1 = σ2 . Finally, since Tσ1 ◦ Tσ2 = Tσ1 ◦h σ2 , the tree transfor-
mation Tσid is an identity element with respect to the composition ◦.
In Theorem 15.1.3 it was shown that for the type τ = (n) with n ≥ 2,
any regular hypersubstitution σ has the property that kerσ is equal to the
diagonal relation 4Wτ (X) .
15.4. EXERCISES 361
15.4 Exercises
15.4.1. Let V be a variety and σ be a hypersubstitution of type τ . The set
15.4.3. Prove that TσV is surjective iff TσV ◦ (TσV )−1 = IdV .
15.4.4. Prove that TσV is injective iff (TσV )−1 ◦ (TσV ) = kerV σ = ∆Wτ (X) .
[3] Arworn, S. and K. Denecke, Left- and right-edge solid varieties of en-
tropic groupoids, Demonstratio Mathematica, Vol. XXXII No. 1 (1999),
1 - 11.
[5] Baker, K. A., Finite equational bases for finite algebras in congruence-
distributive equational classes, Adv. in Math. 24 (1977), 201 - 243.
[6] Baker, K. and A. Pixley, Polynomial interpolation and the Chinese Re-
mainder Theorem for algebraic systems, Math. Z. 143 (1975), 165 -
174.
[7] Berman, J., A proof of Lyndon’s finite basis theorem, Discrete Math. 29
(1980), 229 - 233.
[8] Birkhoff, G., The structure of abstract algebras, Proc. Cambridge Philo-
sophical Society 31 (1935), 433 - 454.
363
364 BIBLIOGRAPHY
[10] Cohn, P. M., Universal Algebra, Harper & Row, New York, 1965.
[22] Denecke, K., Clones closed with respect to closure operators, Multi. Val.
Logic, Vol. 4 (1999), 229 - 247.
BIBLIOGRAPHY 365
[40] Doner, J. E., Decidability of the weak second-order theory of two succes-
sors, Generalized finite automata, Notices Amer. Math. Soc. 12 (1965),
abstract No. 65T-468, 819.
[41] Doner, J. E., Tree acceptors and some of their applications, J. CSS 4
(1970), 406 - 451.
[43] Ésik, Z., A variety theorem for trees and theories, Publicationes Math-
ematicae, Debrecen, Tomus 54 Supplement (1999), 711 - 762.
[56] van Hoa, N., On the structure of self-dual closed classes of three-valued
logic, Diskr. Mathematika 4 (1992), 82 - 95.
[61] Jonsson, B., Algebras whose congruence lattices are distributive, Math.
Scand. 21 (1967), 110 - 121.
[65] Knoebel, A., The equational classes generated by single functionally pre-
complete algebras, Memoirs of the Amer. Math. Soc. 57, 332, Provi-
dence, Rhode Island, 1985.
[66] Knuth, D. E. and P.E. Bendix, Simple Word Problem in Universal alge-
bra, in: Computational problems in abstract algebra, Pergamon Press,
Oxford, 1970, 263 - 297.
[69] Lau, D., On closed subsets of Boolean functions, (A new proof for Post’s
theorem), J. Inform. Process. Cybernet. EIK 27 (1991), 167 - 178.
[72] Lyndon, R. C., Identities in two-valued calculi, Trans. Amer. Math. Soc.
71 (1954), 457 - 465.
[73] Lyndon, R. C., Identities in finite algebras, Proc. Amer. Math. Soc. 5
(1954), 8 - 9.
[78] McKenzie, R., On minimal, locally finite varieties with permuting con-
gruence relations, preprint, 1976.
[79] McKenzie, R., The residual bound of finite algebra is not computable, J.
of Algebra and Computation 6 No. 1 (1996), 29 - 48.
[81] McKenzie, R., The residual bounds of finite algebras, J. of Algebra and
Computation 6 No. 1 (1996), 1 - 28.
[82] McNulty, G., Residual finiteness and finite equational bases: undecidable
properties of finite algebras, Lectures 2000.
[83] Murskij, V. L., The existence in the three-valued logic of a closed class
with a finite basis not having a finite complete system of identities, So-
viet Math. Dokl. 6 (1965), 1020 - 1024.
[89] Pixley, A. F., Functional and affine completeness and arithmetical vari-
eties, in: Algebras and orders, NATO Adv. Sci. Inst. Ser. C Math. Phys.
Sci., 389, Kluwer Acad. Publ., Dordrecht, 1993, 317 - 357.
370 BIBLIOGRAPHY
[105] Schweigert, D., Hyperidentities, in: Algebras and Orders, Kluwer Aca-
demic Publishers, Dordrecht, Boston, London, 1993, 405-506.
[111] Tarski, A., A lattice theoretical fix point theorem and its application,
Pacific. J. Math. 5 (1955), 285 - 310.
373
374 INDEX
Zeitlin, G.J., 21