[FrontPage] [TitleIndex] [WordIndex

Note: You are looking at a static copy of the former PineWiki site, used for class notes by James Aspnes from 2003 to 2012. Many mathematical formulas are broken, and there are likely to be other bugs as well. These will most likely not be fixed. You may be able to find more up-to-date versions of some of these notes at http://www.cs.yale.edu/homes/aspnes/#classes.

What algebras are

An algebra is a set S (called the carrier) together with zero or more operations, each of which is a function from Sk→S for some k. The value k is the number of arguments to the operation, and is called the arity of the operation. Most operations that one encounters are either unary (one argument) or binary (two arguments); examples are negation (unary) and addition (binary). Some programming languages provide ternary operation with three arguments, like C's notorious ?: operation. Some algebras also include 0-ary operations: most people call these constants. The collection of operations provided by an algebra is called the algebra's signature.

For example, the integers with addition and negation have as carrier the set ℤ and signature (+, 0, -), where + is a binary operation, 0 is a 0-ary operation (i.e., constant), and - is a unary operation. Other operations (e.g. binary -) can be defined using operations in the signature.

Often we use the same letter to refer to both an algebra and its carrier when this will not cause confusion.

Why we care

The various standard algebras are abstractions that allow us to reason about lots of different situations at the same time. From the perspective of computer science, they are like base classes high up in an object-oriented type hierarchy, that provide minimal features common to all the classes that inherit from them without specifying the details. So for example in a monoid, where the single operation is associative and there is an identity e that satisfies the equations ae = a and ea = a, we can prove that there can't be a second identity e', since if ae' = a for all a we have ee' = e (from ea = a) and ee' = e' (from ae' = a), giving e = e'. This tells use that we can't extend ℤ by adding a second copy of 0 or the set of n×n matrices by adding a second copy of I; but we didn't need to know anything specific about either structure to know this.

Similarly we can sometimes recognize that our proofs (or later, our algorithms) don't require the detailed structure of ℚ, or ℝ, or ℤ, but in fact work in any algebra of the appropriate type. For example, in computing the length of a path through a graph, we take the sum of the edge weights, and in computing the shortest path, we take the min over all such paths. If we know a lot about semirings, we might recognize this as a sum over products in the (min,+) semiring, and realize that many of our shortest-path algorithms might work equally well in any other semiring.

Cheat sheet: axioms for algebras (and some not-quite algebras)

Axioms

Name

Anything that satisfies this and above is a

0 binary operations

set

1 binary operation +

magma

x+(y+z)=(x+y)+z

Associativity

semigroup

0+x=x+0=x

Identity

monoid

x+(-x)=(-x)+x=0

Inverse

group

x+y=y+x

Commutativity

abelian group

2 binary operations + and ⋅

x(y+z)=xy+xz, (y+z)x=yx+yz

Distributivity

x(yz)=(xy)z

Associativity (multiplicative)

rng (ring without the multiplicative identity)1

1x=x1=x

Identity (multiplicative)

ring

xy=yx

Commutativity (multiplicative)

commutative ring

xx-1=x-1x=1 when x≠0

Inverse (multiplicative)

field

Semirings don't quite fit into this progression; they act like rings, but the + operation gives a commutative monoid instead of an abelian group. An example is the (max,+) semiring, where max acts as the addition operator and + (which distributes over max) acts as multiplication.

See below for examples.

Classification of algebras with a single binary operation (with perhaps some other operations sneaking in later)

The simplest interesting algebras consist of a set together with a single binary operation (*, say). These algebras are classified depending on what axioms * satisfies. Each of the classes below adds an additional constraint to the previous one.

Magmas

If * satisfies closure (which just says that its codomain is S), then (S,*) is a magma (or sometimes groupoid, but groupoid is more often used to mean something else that is more useful: see Groupoid). There really isn't very much to say about magmas, but we can define one interesting magma: the free magma on a set A has as its elements (1) all the elements of A, and (2) (x*y) for any x and y in the magma. This effectively generates all binary trees whose leaves are elements of A.

Semigroups

A magma (S,*) becomes a semigroup if its operator is associative, that is, if (x*y)*z = x*(y*z) for all x, y, and z in S. Semigroups show up quite often in ComputerScience. Examples are:

Most magmas are not semigroups and can't easily be turned into semigroups without destroying information.

Monoids

A semigroup (S,*) is a monoid if it has an identity element e, that is, if there is an element e such that e*x = x and x*e = x for all x. In algebraic terms, we usually think of the identity element as being provided by a 0-ary operation, also known as a constant.

If there is an identity element, it is easy to prove that it is unique: Let e and e' be identities. Then e*e' = e' (by the e*x=x rule) and e*e' = e (by the x*e'=e' rule). Thus e=e'.

Groups

A monoid (S,*) is a group if for each a and b in S, there are solutions x and y to the equations ax=b and ya=b. This fact is equivalent to the existence of a unary inverse operation taking x to x-1 (or -x when the binary operation is written as +), where x-1x=xx-1=e. The existence of an inverse operation can be derived from the more fundamental definition by proving a sequence of properties of groups that are interesting in their own right:

  1. Let y∈S. If there exists any x∈S such that xy=x or yx=x, then y=e. (In other words, any y that ever acts like the identity is the identity; this is not true in general in monoids.) Proof: Suppose xy=x; the other case is symmetric. We will show that y = ey = e. Let q be a solution to qx=e. Then y = ey = (qx)y = q(xy) = qx = e.
  2. If xy = e, then yx = e. (This says that if y is a right inverse of x, it is also a left inverse.) Proof: Let xy = e, and observe that yxy = ye = y. Now let q be a solution to yq=e. Then (yxy)q = yq, but (yxy)q = (yx)(yq) = yx and yq = e, so we have yx = e.

  3. If xy = e and xy' = e, then y = y'. (I.e., right inverses are unique.) Proof: Consider yxy'. Grouping the operations as y(xy') we get yxy' = y(xy') = ye = y. Grouping them the other way gives yxy' = (yx)y' = ey' = y'. So we have y = y'.

It follows that if xy=e, then yx=e and y is uniquely defined by x. Thus the inverse operation x-1 is well-defined (it identifies a unique element of S satisfying the required property xx-1 = x-1x = e). The inverse of the identity is always the identity, since e-1 = e-1e = e.

Note that if we start with an inverse operation, we can easily solve ax=b or ya=b by computing a = axx-1 = bx-1 or a = y-1ya = y-1b.

Examples of groups:

Groups are about the weakest algebraic structure that allow calculations similar to what we expect with ordinary arithmetic: we can add and subtract, solve for unknown variables, etc. Note that the operation in a group does not have to be commutative: it may be that xy ≠ yx for some x and y. For example, in the permutation group on three elements, the function that swaps the first two elements and the function that swaps the last two elements do not commute: 123 → 213 → 231 does not give the same result as 123 → 132 → 312. Similarly, matrix multiplication is in general not commutative.

Abelian groups

If a group is commutative, which means that xy=yx for all x and y, then it is an abelian group (after Niels Henrik Abel, a 19th-century Norwegian mathematician---but by convention spelled with a lowercase "a" despite its etymology).

All of the groups listed previously except the permutation group (on three or more elements) and the free group are abelian. By convention, the binary operation in an abelian group is often (but not always) written as +.

For more on the structure of groups and abelian groups, see GroupTheory.

Operations on algebras

Like sets or graphs, we can define operations that connect one algebra to another or allow us to construct new algebras. When considering multiple algebras, we generally assume that all share the same set of operations (or at least operations with the same names and arities). Most operations on an algebra are defined in terms of operations on the carrier (the set of elements of the algebra).

Subalgebras

A subalgebra of an algebra A is obtained by taking a subset S' of the carrier that is closed under all operations of the algebra; i.e. for any n-ary operation f, f(x1, ..., xn) is in S' if all the xi are. If B is a subalgebra of A, we write B⊆A, as with subsets.

For specific classes of algebras (such as semigroups, monoids, groups, or abelian groups), any axioms satisfied by the parent algebra are inherited by its subalgebras: e.g., if xy = yx in A, then xy = yx in B⊆A.

One important way to get a subalgebra is the subalgebra generated by a particular element or set of elements. This is the smallest subalgebra of a given algebra that includes the specified elements (which are called generators); formally, this is defined by taking the intersection of all subalgebras that contain the generators. (The proof that taking the intersection of subalgebras gives a subalgebra is left as an exercise.)

Examples:

Homomorphisms

A function from the carrier of A to the carrier of B is a homomorphism, written f:A→B, if for any n-ary operation g, g(f(x1), ..., f(xn)) = f(g(x1, ..., xn)). Note that the g on the left-hand side is B's version of g and the g on the right-hand side is A's. If this equation holds for a particular operation g, f is said to preserve g. A homomorphism preserves all operations.

Examples:

In the last case the image of f, defined as f(A) = { f(x) | x in A }, is a subgroup of the codomain ℤ. This turns out always to be the case for images of algebra homomorphisms:

Theorem 1
Let f:A→B be an algebra homomorphism. Then f(A) is a subalgebra of B.
Proof

To show that f(A) = { f(x) | x in A } is a subalgebra of B, we need to show that it is closed under every operation of B. Let g be an n-ary operation of B, and consider g(y1, ..., yn) where each yi is in f(A). Because each yi is in f(A), there exists some xi such that yi = f(xi). So we can rewrite g(y1, ..., yn) as g(f(x1), ..., f(xn,)) = f(g(x1, ..., xn)) ∈ f(A). (The equality holds because f is a homomorphism.)

As with graphs, homomorphisms on algebras give rise to isomorphisms (bijective homomorphisms, the inverses of which will also be homomorphisms) and automorphisms (isomorphisms between an algebra and itself). Every algebra has at least one automorphism, the identity homomorphism that maps every element to itself. An injective homomorphism is called an embedding (sometimes imbedding); any embedding f of A into B gives an isomorphism between A and B's subalgebra f(A).

More examples:

Free algebras

We've seen examples of various kinds of algebras that have been called "free", such as free magmas, free semigroups, etc. There is in fact a single definition of a free algebra (from a particular class of algebras) that produces each of these specific free algebras.

The essential idea of a free algebra A over a set S is that (a) it contains all the elements of S, (b) it contains g(x1, ..., xn) for any n-ary operation g whenever x1,...,xn are in A (which is just the closure requirement), and (c) for particular xi and x'i, g(x1,....,xn) = g(x'1,...,xn) only if required by the axioms.

So in a free magma, x(yz) is never equal to (xy)z, no matter what values x, y, and z have. In a free semigroup, x(yz) = (xy)z, always, because of the associativity axiom, but xy is never equal to x for any x and y, xy is never equal to yx, etc. In a free monoid, xe = ex = x, but xy is never equal to x unless y = e, and so forth.

The formal definition that yields these properties is this:

Let C be a class of algebras, where the names and arities of the operations are the same for all algebras in the class, and let S be a set. Then a free algebra in C over S, written F(S) or FC(S), is an algebra in C such that (a) S is a subset of the carrier of F(S), and (b) for any algebra B in C, and any function from S to the carrier of B, there is a unique homomorphism f* from F(S) to B such that f*(x) = f(x) for any x in S. (In this case f is said to lift to f*; another way of expressing this is that f is a subset of f* when both are viewed as sets of ordered pairs.)

How does this definition yield all the specific cases of free algebras? Let's take free monoids as an example. Recall that a free monoid F(S) over S consists of all finite sequences of elements of S, with + = concatenation and 0 = the empty sequence. It's trivially the case that S is a subset of the carrier of F(S) (assuming we identify single elements with one-element sequences). Now let's show that any function f from S to the carrier of some semigroup B lifts to a unique f* from F(S) to S.

Define f*(x1...xk) = f(x1)f(x2)...f(xk). It is easy to see that f* is a (monoid) homomorphism, and that f*(x) = f(x) when x is in S. Now let g be any other monoid homomorphism from F(S) to B such that g(x) = f(x) for any x in S. We'll show that for sequence s=x1...xk in F(S), g(s) = f*(s), by induction on the length k of s. The cases k=0 and k=1 are the base cases. For k=0, g(<>) = f*(<>) = the identity of B because both g and f* are monoid homomorphisms (and therefore must preserve the identity). For k=1, we have g(x1) = f(x1) = f*(x1) by the requirement that g(x) = f(x) when x in S. For k > 1, we have g(x1...xk-1xk) = g(x1...xk-1)g(xk) = f*(x1...xk-1)f*(xk) = f*(x1...xk-1xk).

Note that there may be more than one way to define a free algebra. For example, we could define the free monoid over S by setting + to be reverse concatenation: xyz+abc = abcxyz. This again gives sequences of elements of S, but to translate them back to the previous definition we have to reverse the sequences. Fortunately this reversal operation is an isomorphism between the two monoids, and in general we can show that any free algebra for a given class with a given base set is unique up to isomorphism, which means that while there may be more than one such algebra, there is only one isomorphism equivalence class.

Theorem 2
If F and G are free algebras over S in a class C, then F and G are isomorphic.
Proof

The proof is by repeated application of the uniqueness of f*:F(S)→A for any f:S→A.

  1. Let f:S→G be defined by f(x)=x. Then there is a unique homomorphism f*:F→G.

  2. Similarly, let g:S→F be defined by g(x) = x. Then there is a unique homomorphism g*:G→F.

  3. We will now show that g*=(f*)-1. Consider the homomorphism h=g*∘f*. For any x in S, we have h(x) = g*(f*(x)) = g(f(x)) = x. So h is a homomorphism from F to F, such that for any x in S, h(x)=x. Since F is a free algebra, h is unique among all such homomorphisms: if somebody shows up with another homomorphism h':F→F with h'(x) = x for all x in S, then h'=h. Now out of the box let us pull IF:F→F, the identity homomorphism on F defined by IF(x)=x for all x in F. Clearly IF(x)=x for all x in S, so IF=h. It follows that g*∘f* is the identity---i.e., that g* = (f*)-1. Since f* is invertible and thus bijective, it's an isomorphism, and F and G are isomorphic.

The nice thing about this proof is that at no time did we have to talk about specific operations of the algebras; everything follows from the definition of a free algebra in terms of homomorphisms.

Applications of free algebras

The selling point for free algebras is that we can define homomorphisms just by defining what happens to elements of the underlying set S. This operation is done all the time. Some examples:

Product algebras

Let F and G be algebras with the same signatures. Then F×G is an algebra whose elements consist of a pair of one element from F and one from G (i.e., is the usual cartesian product F×G), and an operation f on (F×G)k carries out the corresponding F and G operations on both sides of the ordered pairs.

Examples:

Congruences and quotient algebras

A congruence is an equivalence relation on elements of an algebra that respects the structure of the algebra. In particular, a relation ~ is a congruence on A if

  1. It is an equivalence relation on elements of A, and
  2. for any n-ary operation g of A, if x1 ~ y1, x2 ~ y2, ..., xn ~ yn, then g(x1, x2, ... xn) ~ g(y1, y2, ... yn).

An example of a congruence is the relation x~y if x and y are both even or x and y are both odd, where x and y are elements of additive group of the integers (ℤ,+,0,-).

Given a congruence ~ on an algebra A, we can define a quotient algebra A/~ whose elements are the equivalence classes C of ~, and whose operations are defined by letting g(C1...Cn) be the equivalence class of g(x1...xn) where each xi is some element of the corresponding Ci. The requirements for ~ to be a congruence guarantee that this defines a unique equivalence class no matter which representatives xi we pick.

For example, the odd-even congruence defined previously yields a quotient group ℤ/~ = ℤ2, the integers mod 2. Associated with this congruence is a homomorphism that maps each integer to one of the two elements 0 or 1 of ℤ2, depending on whether it is even (0) or odd (1). This is not surprising; in fact, there is always such a homomorphism from any algebra to any of its quotient algebras.

Theorem 3
If ~ is a congruence on A, then the function f mapping each x in A to its equivalence class in A/~ is a homomorphism.
Proof

We need to show that g(f(x1)...f(xn)) = f(g(x1...gn)) for each n-ary operation g and sequence of arguments {xi}. For each i let Ci = f(xi) be the equivalence class of i under ~; by the definition of operations in A/~ we have g(C1...Cn) is the equivalence class of g(x1...gn) = f(g(x1...gn)).

The preceding theorem tells us we can map A onto A/~ via a homomorphism. There is a sense in which we can go the other way: given any homomorphism f:A→B, we can define a congruence ~ on A by x~y if f(x)=f(y). This congruence is called the kernel of f. (Note: in GroupTheory, the kernel is defined differently, as the subgroup Ker f of A given by f-1(e). The relation between the kernel-as-subgroup and the kernel-as-congruence is that in a group we can define a congruence a~b iff ab-1 is in Ker f, which turns out to be equivalent to the definition f(a)=f(b) for algebras in general.)

Theorem 4
Let f:A→B be a homomorphism, and let ~ be the kernel of f. Then ~ is a congruence.
Proof

It is easy to see that ~ is an equivalence relation, so we will concentrate on showing that xi~yi implies g(x1...xn)~g(y1...yn) for any {xi}, {yi}, and g. Pick some particular xi, yi, and g; then we have f(xi)=f(yi) (from the definition of the kernel) and so f(g(x1...xn))=g(f(x1)...f(xn))=g(f(y1)...f(yn))=f(g(y1)...g(yn)), which implies g(x1...xn)~g(y1)...g(yn).

We have previously observed that the image of a homomorphism is a subalgebra of its codomain. This subalgebra turns out to be isomorphic to the quotient algebra of the domain by the kernel of the homomorphism:

Theorem 5 (First Isomorphism Theorem)
Let f:A→B be a homomorphism, and let ~ be the kernel of f. Then A/~ is isomorphic to f(A).
Proof

Define h:(A/~)→B by setting h(C) = f(x), where x is any element of C. Note that h is well-defined because each C consists precisely of all x for which f(x) is some particular element y of B. We can also define an inverse h-1 by letting h-1(y) = { x | f(x) = y }, so h is a bijection. It remains only to show that it is a homomorphism. Let g be an n-ary operation of A/~, and let x1...xn be representatives of equivalence classes C1...Cn. Then h(g(C1...Cn)) = f(g(x1...xn)) = g(f(x1)...f(xn)) = g(h(C1)...h(Cn).

In the case that f is surjective, the theorem gives an isomorphism between A/~ and B. This is a fundamental tool in many branches of algebra.

Algebraic structures with more binary operations

All of the structures we have considered so far had only a single binary operation, which we usually wrote as either multiplication or addition. We now consider structures that have more binary operations. The simplest of these, rings and fields, are the natural generalization of the ways that real numbers, integers, etc. support both an addition and a multiplication operation.

Rings

A ring is a set S together with two binary operations, + and ×, called addition and multiplication (as elsewhere, they are called this because of the symbol used, even if they really do something else). Considering each of the operations separately, we require:

In addition, a distributive law relates the two operations. For all x, y, and z, we have:

Note the need for both equations since multiplication is not necessarily commutative.

In writing about rings, we adopt the usual convention that multiplication binds more tightly that addition and drop the multiplication operator, so that we can w, for example, yx+zx for (y×x)+(z×x).

Examples of rings

Properties of rings

Though rings vary widely in their properties, some properties follow from the definition and thus hold in all rings. Two important theorems about rings are that multiplying by 0 has the effect we expect from our experience with the integers, rationals, etc., and that negation commutes with multiplication.

Theorem
Let R be a ring and let x be any element of R. Then 0×x=x×0=0.
Proof
We'll prove that 0×x=0; the other case is symmetric. The proof uses both the distributive law and the existence of 1, the multiplicative identity. Compute 0x+x = 0x+1x = (0+1)x = 1x = x = 0+x. But since addition in a ring is a group, if 0x+x = 0+x then 0x = 0.

The familiar rule that (-x)y = -(xy), (-x)(-y) = xy, etc. follows immediately from the preceding theorem:

Corollary
Let R be a ring and x and y any elements of R. Then (-x)y = x(-y) = -(xy).
Proof
Again we will prove only the first case (-x)y = -(xy). It is enough for this to show that (-x)y + xy = 0. Compute (-x)y + xy = (-x + x)y = 0y = 0. Here the first step uses the distributive law, the second uses the definition of the additive inverse, and the last uses the theorem---just proved---that 0x = 0 for all x.

Invertible elements and zero divisors

An element of a ring is invertible if it has a multiplicative inverse. The element 1 is always invertible, since 1*1=1, and so is -1 (although -1 may be the same element as 1, as in ℤ2). The set of invertible elements is closed under multiplication: if a is invertible and b is invertible, then so is ab (its inverse is b-1a-1). Since it also contains the multiplicative identity and inverses, it's a group, called the multiplicative group of the ring. So the multiplicative groups ℤ*m of integers relatively prime to m are just a special case of this definition.

One way to show that an element is not invertible is to show that it's a zero divisor. An element x is a zero divisor if it divides 0, i.e. if there is some element y≠0 such that xy = 0 or yx=0. For example, in ℤ6, the element 2 is a zero divisor (2*3=0). A zero divisor can't be invertible because if xy=0 and x-1 existed, then y=x-1xy=x-10=0. However, in many rings there are elements that are not invertible without being zero divisors, e.g. 2 in ℤ or x in ℤ[x].

Subrings

A subring S of a ring R is just a subset of R that is also a ring. The main thing we need to prove in order to show that this is true is that S is closed under all the ring operations.

Examples:

Ideals and quotient rings

An ideal of a ring R is a subring S such that whenever x is in S, xy and yx are in S for any y in R. An example is the even integers, which are an ideal of the ring of integers. An example of a subring that is not and ideal is the dyadics; (1/2)(1/3) = (1/6), which is not a dyadic.

The main reason for looking at ideals is that they have the same role in rings that normal subgroups have in groups: the additive cosets in R of an ideal I yield a quotient ring R/I. This follows from the fact that as an additive subgroup, an ideal is always a normal subgroup (since addition is commutative), so (I+a)+(I+b)=I+(a+b) is well-defined; and that when we multiply (I+a)(I+b), we get II + Ib + aI + ab = I + ab since the first three terms all collapse into I. One way to remember the required properties of an ideal is that the ideal must act like 0 in the quotient ring it yields: it's an additive identity (normal subgroup) but a multiplicative annihilator (drags products into itself).

Some examples of quotient rings:

Ring homomorphisms

A ring homomorphism is a function that preserves all the ring operations (this is just a special case of the usual definition of an algebra homomorphism). So in particular, given a function f:R→S, we need f(0R) = 0S, f(1R) = 1S, f(a+b) = f(a)+f(b), and f(ab) = f(a)f(b). These are pretty strong requirements, so ring homomorphisms are much rarer than, say, group homomorphisms.

Some examples of ring homomorphisms:

Semirings

A semiring is like a ring, except that the addition group is replaced by a commutative monoid; alternatively, we can think of a semiring as a ring without additive inverses. The theorem in rings that 0*x=x*0=0 becomes an additional axiom (the annihilation axiom) for semirings—without inverses it is impossible to prove it from the other axioms.

An example of a semiring is the tropical semiring or max-plus semiring on ℝ∪{-∞} where addition is max, multiplication is +, the additive identity is -∞, and the multiplicative identity is 0. This gives a commutative monoid for both addition (max) and multiplication (+), since max and + both commute and max(x,-∞) = x, but neither operation yields a group (for the multiplication operation +, -∞ has no inverse). It also satisfies the distributive law: x+max(y,z) = max(x,y) + max(x,z); and the annihilation axiom: -∞ + x = -∞. The tropical semiring is useful for studying scheduling problems: for example, if it takes 2 years to graduate after taking CPSC 202 and 3 more years after taking MATH 112, then my graudation time is at least max(t202 + 2, t112 + 3), and the ability to manipulate complicated expressions involving max and + may help me plan out more difficult tasks.

For more examples of semirings see Semiring.

Fields

If the nonzero elements of a commutative ring form an abelian group, the ring is a field. An equivalent definition is that 1≠0 (so the group is nonempty) and every nonzero element is invertible. The most common examples in ordinary mathematics are ℚ, ℝ, and ℂ, with ℤp showing up in NumberTheory. A more exotic example is the field of rational functions F(x1,...xn,): functions of the form p/q where p and q are both polynomials with variables x1 through xn and coefficients in F, with q≠0.

For ComputerScience, the most useful fields are the finite fields, since (being finite) they fit inside computers. There is a unique (up to isomorphism) finite field of size pk for every prime p and every positive integer k. These fields are known as the Galois fields after Evariste_Galois (the guy who died from staying up all night doing his homework) and are written as GF(pk). When k=1, they have a particularly simple construction; for larger values of k the Galois fields are more complicated.

The field GF(p) is just ℤp with multiplication defined mod p. We've already seen (in NumberTheory) that every nonzero element of ℤp has a multiplicative inverse; everything else follows from ℤp being a ring.

For n > 0, there is a standard construction of GF(pn) based on polynomials. The idea is to start with all polynomials over ℤp, and then take remainders mod some particular polynomial to get rid of any high-degree terms. We'll do GF(22) as an example.

Step 1: Consider all polynomials in ℤ2[x]: 0, 1, x, 1+x, 1+x+x3, etc. Note that the only coefficients are 0 and 1, since those are the only elements we have in ℤ2.

Step 2: Choose an irreducible polynomial r, one for which the equation r(x) = 0 has no solutions in ℤ2. The degree (highest exponent) of this polynomial will determine the n in GF(pn). We'll pick r(x) = 1+x+x2. It is easy to check that r(0) = 1+0+0 = 1 and r(1) = 1+1+1 = 1, so it's irreducible.

Step 3: Let the elements of GF(pn) be the residue classes mod r, i.e. the equivalence classes of polynomials where p1~p2 if there is some polynomial q such that p1 = p2 + qr. In GF(22) this means that whenever we multiply two polynomials together and get a polynomial with an x2 in it, we subtract r to get rid of the x2. In this way we take as representatives of the classes the polynomials that have degree at most n-1; in GF(22) this will be the four polynomials 0, 1, x, and 1+x.

Addition of elements of GF(22) is done term by term, e.g. (1+x) + (x) = 1. (The x's cancel since 1+1=0 in ℤ2). Multiplication is more complicated. Here's the full multiplication table:

0

1

x

1+x

0

0

0

0

0

1

0

1

x

1+x

x

0

x

1+x

1

1+x

0

1+x

1

x

Most of the entries in the table are straightforward (we know, for example, that 0*y = 0 for all y). The tricky ones are in the bottom right corner. Let's look at x*x first. This is x2, which is too big, so we'll take the remainder mod x2+x+1, which in this case means computing x2-(x2+x+1)=x+1 (remember we are in ℤ2 where 0-1=1). Similarly we can compute (1+x)*(1+x) = 1 + 2x + x2 = 1+x2 = 1+x2-(1+x+x2) = x. For x(1+x) we have x(1+x) = x+x2 = x+x2-(1+x+x2) = 1.

This construction is a special case of a more general technique called field extension, where one makes an existing field (ℤ2 in this case) bigger by tacking on an extra member (x) that satisfies some polynomial equation not satisfied by any element of the old field (1+x+x2=0). Another example of the same technique is the construction of the complex numbers from the reals by taking ℝ and adding a new element i that satisfies 1+i2=0. Our operation of replacing x2 with x+1 in GF(22) is the equivalent in that field of replacing i2 with -1 in the complex numbers. The fact that we can get away with this in each case can be justified formally by sufficient handwaving about free rings and quotients.

Note that the structure of GF(22) is very different from ℤ4, which isn't a field (since 2*2 = 0 mod 4). Even the additive group of GF(22) is not the same as ℤ4; there is no single element, for example, that generates the entire additive group. This is typical of GF(pn), whose additive group in general will look like the product of n copies of ℤp, and whose multiplicative group will be terribly tangled up.

For more on finite fields see http://mathworld.wolfram.com/FiniteField.html or Chapter 23 of BiggsBook.

Exercise
Which of the examples of rings given previously are also fields?

Subfields and homomorphisms

Subfields are fairly common. For example, ℚ is a subfield of ℝ which is in turn a subfield of ℂ. Among finite fields, GF(p) is a subfield of GF(pn) for any n (consider the set of polynomials with only a constant term).

Field homomorphisms are much less common; indeed, we can prove that all homomorphisms between fields are injective (giving an embedding—a bijection between the domain and some subfield of the codomain). Proof: Let f:X→Y be a field homomorphism and suppose that f is not injective; i.e., that f(x)=f(y) for some x≠y. Then f(x-y) = f(x)-f(y) = 0. But since x≠y, x-y has an inverse (x-y)-1, and so f(1) = f((x-y)(x-y)-1) = f(x-y)f((x-y)-1) = 0⋅f((x-y)-1) = 0 ≠ 1, contradicting the requirement that f maps 1 to 1. It follows that any homomorphic f is injective.

Vector spaces

A vector space is not, strictly speaking, an algebraic structure of the sort we have defined previously, as it involves two classes of objects: scalars, which are elements of some field, and vectors, which are elements of some abelian group, written additively. In addition to their own field and group operations, scalars and vectors are connected by an operation called (rather confusingly) scalar multiplication that multiplies a scalar by a vector to produce another vector. Scalar multiplication must satisfy the following axioms, where normal letters represent scalars and boldfaced letters represent vectors:

The distributive laws imply further properties like 0x = 0 or (-1)x = -x; the proofs are essentially the same as in a field. Note that the two zeroes and the two negations in these formulas are different; on the left-hand side we are in the scalar field (or its additive group), and on the right-hand side we are in the vector space (thinking of it as a group).

When writing about vector spaces, the vectors have the starring role, and we think of a vector space V as consisting of the set of vectors, with the field of scalars often assumed to be some standard field like the reals or the complex numbers. (If we want to emphasize that the scalar field for some particular vector space V is F, we say that V is a vector space over F.) The original motivation for vectors was directions in (real, Euclidean, physical) space: in pirate terms, vector x might representing walking one pace North, vector y might represent walking three paces East, and vector 9x+12y might represent walking 9 paces north then 12*3=36 paces East in order to find the buried treasure. The assumption that vector addition is abelian says that it doesn't matter if we walk the 9x vector or the 12y vector first---this assumption may or may not world in the real world of treasure islands full of traps and pitfalls, but we should get to the same place either way in an idealized space without obstacles. The properties of vector multiplication follow from the intuition that multiplying by a scalar is "scaling" the vector---stretching it or shrinking it in its original direction by some factor determined by the particular scalar we choose.

If we replace the scalar field with a scalar ring, we get a more general structure called a module. We won't talk about these much.

For calculational purposes, we often think of vectors as sequences of coordinates, where each coordinate is an element of the field the vector space is defined over. Adding two vectors involves adding their coordinates componentwise, e.g. (1,2,3) + (30,20,10) = (31,22,13), and scalar multiplication also multiplies componentwise, e.g. 3(1,2,3) = (3,6,9). The length of each sequence is the dimension of the vector space. This approach can be taken in any vector space that has a finite basis: details are given in AbstractLinearAlgebra.

Some other examples of vector spaces:

Homomorphisms of vector spaces

A homomorphism between two vector spaces over the same field is a homomorphism between the groups of vectors that also preserves scalar multiplication. In particular, this means that

Such homomorphisms are called linear transformations or linear maps. Linear transformations between finite-dimensional vector spaces can be represented by matrices (more on this in AbstractLinearAlgebra).

Subspaces

A subspace of a vector space is just a subgroup of the group of vectors that is closed under scalar multiplication, i.e. a set V'⊆V such that x+y, -x, and ax are in V' whenever x and y are, where a is any scalar. Subspaces also have an important role in AbstractLinearAlgebra and will be discussed further there.


CategoryMathNotes

  1. Also called pseudo-rings. You do not need to remember this definition. (1)


2014-06-17 11:57