Introduction To Ring Theory: Sachi Hashimoto
Introduction To Ring Theory: Sachi Hashimoto
Introduction To Ring Theory: Sachi Hashimoto
Sachi Hashimoto
Mathcamp
Summer 2015
1 Day 1
1.1 What are we talking about?
Broadly speaking, a ring is a set of objects which we can do two things with: add and
multiply. In many ways it will look like our familiar notions of addition and multiplication,
but sometimes it won’t. We have to decide what properties of addition and multiplication
we are going to require to be true, and which ones we can do without. Before we give these
properties, let’s go through a few familiar examples.
Example 1. Our basic example of a ring will be the integers, which we will write as Z. As
a set, this is just the numbers {0, 1, −1, 2, −2, . . . } and addition and multiplication work
“as usual”.
Example 2. Another key example that will come up again and again are the polynomials,
which I will denote as Z[x]. Here, my objects are polynomials an xn + an−1 xn−1 + · · · +
a1 x + a0 , where the coefficients ai are integers. We could also take polynomials whose
coefficients are in R or Q or C, in which case we would write our ring as R[x] or Q[x] or
C[x]. Polynomials come equipped with a usual notion of how to multiply them and how
to add them.
These two basic examples in some sense represent all of the examples we will talk about
in this class. Roughly speaking, we usually think of rings as either ‘functions’ or ‘numbers’.
(Actually, we could generalize even further, and think of numbers as constant functions,
but this is quite silly, and it’s better to just think of numbers like Z or Q.) Let’s look at
three more examples.
Example 3. The functions f : R → R form a ring which we will (nonstandardly) name
Fun(R, R). When we add two functions f and g, we get a new function (f + g) : R → R
which takes r ∈ R to f (r) + g(r). When we multiply two functions, f and g, we produce
a new function (f · g) : R → R where (f · g)(r) = f (r)g(r). We call this pointwise addition
and pointwise multiplication, because the rule to add or multiply two functions is to add
or multiply them at each point.
1
Example 4. The Gaussian integers are the subset of the complex numbers consisting of
all complex numbers C with integer coefficients {a + bi|a, b ∈ Z}. We write them as Z[i].
The usual addition and multiplication of complex numbers apply.
Example 5. Modular arithmetic, Z/nZ, is a ring with the usual addition and multiplica-
tion.
Example 6. The two by two matrices with entries in the real numbers R form a ring,
under matrix multiplication and addition.
Definition. A ring, R, is a set of objects along with two binary operations, · multiplication
and + addition, with the following properties:
7. Distributivity: for any a, b, c ∈ R we have that a(b+c) = ab+ac and (b+c)a = ba+ca.
One thing that might seem funny about these examples is that while we require addition
to be commutative and have inverses, we don’t require multiplication to be commutative
or have inverses. We want to be able to work with rings that have a looser multiplicative
structure: for example, we want our rings to model sets of functions, and one thing we
notice about functions is that they don’t always commute. The two by two matrices over
R are are good simple example of a noncommutative ring. Also, in the case of rings of
numbers, like our example Z, many numbers don’t have multiplicative inverses. The only
integers which have multiplicative inverses are 1 and −1.
Remark. In this class, we will only work with commutative rings (rings where multipli-
cation is commutative), but for the purpose of proving things in more generality, all of the
proofs we do today will work whether or not we require our rings to be commutative.
2
1.3 Basic Properties of Rings
From the definition, we can deduce a few basic propositions. The goal in proving these is
to say some facts about rings, but also to give you an example for how basic proofs in ring
theory go.
Proof. Suppose R is a ring with two multiplicative identities, 1 and 10 . Then both of them
satisfy the property that for all r ∈ R, 1r = r1 = r and 10 r = r10 = r.
In particular, we can let r = 10 in the first equation and then we get the identity
11 = 10 1 = 10 but in the second equation letting r = 1 we get 10 1 = 110 = 1. This proves
0
that 1 = 10 .
Proof. First consider (−a)b. Then we know that −a is the element such that a + (−a) = 0.
So using distributivity we get that (a + (−a))b = ab + (−a)b and also (a + (−a))b = 0b = 0
by the previous proposition. Therefore ab + (−a)b = 0. Let’s add −(ab) to both sides:
we get ab + (−a)b + (−(ab)) = −(ab) and we can commute the things on the left side
to get 0 + (−a)b = −(ab) and so (−a)b = −(ab). You will show on the homework that
a(−b) = −(ab) to complete the proof.
We’re not going to be this explicit about associativity, zero, commutativity, and so on
in the future, because it does get a little tedious. However, it’s good to have a grounding
in the basics.
3
2 Homework Day 1
2.1 The Basics
1. Prove that the additive identity 0 is unique. That is, if 0 and 00 are two elements of
a ring R such that for all r ∈ R, r + 0 = 0 + r = r and 00 + r = r + 00 = r then 0 = 00 .
3. Convince yourself that each of the examples we talked about in class are actually
rings. That is, go through the definition of ring, and verify any property that you
aren’t sure you believe. (For example: what is the identity in the ring of functions
from R to R?)
4. Show that additive identities are unique: that is, show that if a + b = 0 then b = −a.
2. Which elements in Z[x] have multiplicative inverses? What about Q[x]? What about
Fun(R, R)?
3. Let X be a set and let P (X) be the power set of X, that is, the set of all subsets of X.
(For example, if X = {a, b, c} then P (X) = {∅, {a}, {b}, {c}, {a, b}, {a, c}, {b, c}, {a, b, c}}).
Show that we can make P (X) into a ring where if Y, Z ∈ P (X) then Y + Z =
(Y − Z) ∪ (Z − Y ) and Y · Z = Y ∩ Z. What is the identity? What is Y · Y for any
Y ∈ P (X)?
2.3 To ponder
1. Consider the power series ring Q[[x]] that consists of elements of the form ∞ n
P
n=0 an x .
What elements have multiplicative inverses? What is the inverse of the power series
1 + x? Note that Q[x], the polynomials in x with rational coefficients, sits inside
Q[[x]] and 1 + x ∈ Q[x] but 1 + x is not a unit in Q[x]. Deduce that if S is a subring
of R then all of the units of S are units of R but not all of the units of R contained
in S are also units of S.
4
3 Day 2
Consider the rings Z[x] = {an xn +· · ·+a1 x+a0 |ai ∈ Z} and Z[y 2 ] = {b2n y 2n +b2n−2 y 2n−2 +
· · · + b2 y 2 + b0 |b2i ∈ Z}. These two rings act and seem very much the same. In fact, we can
make a one to one correspondence of objects in Z[x] with objects in Z[y 2 ] just by taking
replacing each x with y 2 , or vice verse. How do we quantify this notion of sameness? How
can we in general talk about rings being the same?
Definition. We say two rings R and S are isomorphic if there is a bijective function
f : R → S such that f (a +R b) = f (a) +S f (b), f (a ·R b) = f (a) ·S f (b), and f (1R ) = 1S .
More generally, we can make analogies between rings by considering other functions
from one ring to another.
Here, we’re just relaxing the idea that it has to be a bijection. If our function is injective,
then f gives us a way to situate R in S as a smaller ring, and if our function is surjective,
we will see that f gives a way of grouping elements of R into “symmetry classes” so that
S is R mod some symmetries. In math, we often look at homomorphisms to tell us more
about the structure of a ring, rather than looking at the individual ring itself.
Example 7. Consider the map f : Z[x] → Z given by p(x) 7→ p(0). This just plucks out
the constant term of the polynomial p(x). This map is surjective but not injective. All the
polynomials with no constant term get sent to 0.
Notice that if p and q are polynomials that get sent to 0 under f then p + q and rp get
sent to zero for any r ∈ Z[x].
Example 8. More generally, if we have a ring of functions like Fun(R, R), we can create
a ring homomorphism which is evaluation of the functions at a point. For example, the
ring of functions from R → R has a homomorphism fp : Fun(R, R) → R for each point
p ∈ R, sending φ : R → R to φ(p). We can check that if φ : R → R and ψ : R → R
are elements of Fun(R, R), then fp (φ + ψ) = (φ + ψ)(p) = φ(p) + ψ(p) = fp (φ) + fp (ψ).
Also fp (φ · ψ) = (φ · ψ)(p) = φ(p)ψ(p) = fp (φ)fp (ψ). Finally since the constant function
c1 : R → R is the multiplicative identity in Fun(R, R) we check fp (c1 ) = c1 (p) = 1.
Example 11. There are injective maps Z → Z[x] and Z → Q realizing Z as a smaller ring
inside these rings.
Definition. More formally: a subring of a ring is a subset of a ring which is also a ring.
5
Definition. The image of a ring homomorphism f : R → S is the set of elements s ∈ S
such that there exists r ∈ R with f (r) = s.
Let’s go back and take a look at the image and kernel of the previous examples.
Notice that we can characterize being injective as having the kernel equal to 0, and
surjective as having image equal to the target.
2. If r, s ∈ ker(f ) then so is r + s.
3. 0 ∈ ker(f ).
Proof. We’ll start with 0 ∈ ker(f ). Note that f (0) = f (0 + 0) = f (0) + f (0), so subtracting
f (0) from both sides in S we see that 0 = f (0).
Then, consider r ∈ ker(f ) and suppose c is any element of R. Then f (cr) = f (c)f (r)
because f is a homomorphism, and f (r) = 0 because r is in the kernel of f . Thus f (cr) =
f (c)f (r) = f (c)0 = 0.
If r, s ∈ ker(f ) then f (r + s) = f (r) + f (s) = 0 + 0 = 0 so r + s ∈ ker(f ).
2. If r, s ∈ I then so is r + s.
3. 0 ∈ I.
an ideal of R.
Example 14. Let p ∈ R and consider Mp = {f ∈ Fun(R, R)|f (p) = 0}. Then Mp is an
ideal of R and there are no other ideals other than Fun(R, R) containing Mp .
6
Definition. We say M ⊂ R is a maximal ideal of R if it is not contained in any proper
ideals of R.
Example 15. The ideal (x) in Z[x] is all polynomials divisible by x. Similarly, for any
element p(x) we can create the ideal generated by p(x) which will consist of all multiples of
p(x) and be denoted (p(x)). Another way of thinking of this notation is that (a, b, c) ⊂ R
is the smallest ideal containing the elements a, b, c. So it must contain all multiples as well
as differences and sums.
7
4 Homework Day 2
4.1 The Basics
1. Let R be a commutative ring. A field is a commutative ring where every nonzero
element has a multiplicative inverse. Show that {0} is a maximal ideal of R if and
only if R is a field. Hint: An ideal is a proper ideal if and only if it does not contain
1. Consider the ideals (a) for a ∈ R.
2. Find all ring homomorphisms from Z to Z/30Z. Describe the kernel and image.
3. What is the ideal (4, 6) generated by the elements 4 and 6 in Z? What about (2, 3)?
What about (m, n)?
4.3 To ponder
1. Show that if I1 ⊂ I2 ⊂ . . . are ideals of a ring R then so is ∪∞n=1 In . Come up with
an example of an inclusion of ideals I1 ⊂ I2 ⊂ I3 ⊂ . . . such that In 6= In+1 for all n.
(You will need to think of rings that we haven’t talked about yet.)
8
2. Let R be a commutative ring. Then we denote N(R) to be the nilradical of R, where
N(R) = {r ∈ R|rn = 0, for some n ∈ N} is the set of all nilpotent elements. Show
that N(R) is an ideal.
9
5 Day 3
Yesterday, we defined ideals and talked about how they arose from the kernel of ring
homomorphisms. Today, we will talk about how, given an ideal I of R we can find a ring
S and a ring homomorphism φ : R → S where I = ker(φ). To do this, we will construct S
as a quotient ring R/I.
On the homework, for r ∈ R and I an ideal of R, we constructed the set r + I =
{r + a|a ∈ I}. For example, if R = Z and I = 5Z we computed that we have 0 + 5Z, 1 + 5Z,
2 + 5Z, 3 + 5Z, and 4 + 5Z. If n is any other number then n + 5Z = (n mod 5) + 5Z. For
example, 7 + 5Z = 2 + 5Z.
Recall from the homework that we have the following proposition:
Proof. We showed this by noting that if r−s ∈ I then r+I = r−s+s+I = {r−s+s+a|a ∈
I} = {s + a|a ∈ I} since r − s + a ∈ I because I is closed under addition. Conversely, if
the two sets are equal then 0 ∈ I implies r ∈ r + I so r ∈ s + I so r = s + a for some a ∈ I
and so r − s = a ∈ I.
Proposition 6. Addition and multiplication defined in this way on R/I are well defined.
Proof. Suppose r, r0 are two elements of [r] = [r0 ] and s, s0 are both elements of [s] = [s0 ].
Then we want to show that [r + s] = [r0 + s0 ] On the homework, we showed that r − r0 ∈ I
and s − s0 ∈ I. Let’s call these a and b. Then (r + s) − (r0 + s0 ) = a + b ∈ I since I is closed
under addition. Therefore by the some homework problem, we see that [r + s] = [r0 + s0 ].
Now consider [rs] and [r0 s0 ]. We would like to show that rs − r0 s0 ∈ I. Let’s replace r0
with r − a and s0 with s − b. Then rs − r0 s0 = rs − (r − a)(s − b) = rs − rs + as + bs + ab =
as + bs + ab ∈ I, since anything times a or b is in I, as a, b ∈ I. Therefore [rs] = [r0 s0 ].
10
This proposition showed us that it doesn’t matter what name we give for these sets
r + I, the operations on R/I are well defined. In fact, these operations turn R/I into a
ring (if you don’t believe it, take some time on the homework to test out each of the ring
axioms.)
Note that there is a nice way to take elements of R to elements of R/I, by taking r to
r + I.
Proposition 7. The map φ : R → R/I given by r 7→ r + I is a surjective ring homomor-
phism with kernel I.
Proof. Clearly this map is surjective. Note that we actually defined R/I to make this into
a ring homomorphism: suppose r, s ∈ R. Then φ(r + s) = [r + s] = [r] + [s] = φ(r) + φ(s)
and similarly φ(rs) = [rs] = [r][s] = φ(r)φ(s). Check for yourself that φ(1) = [1] is the
identity in R/I.
What’s the kernel of φ? It’s the set of r ∈ R such that [r] = [0], i.e. the set of r ∈ R
such that r − 0 ∈ I or r ∈ I.
Thus, every ideal is the kernel of a homomorphism and every kernel is an ideal, and we
know exactly how that correspondence works.
Example 16. Let’s use the quotient ring construction in one particular case, Z[x]/(x2 −2).
Suppose p(x) is any polynomial in Z[x]. We can use polynomial long division (since x2 − 2
has leading coefficient 1) to write p(x) = r(x)+(x2 −2)q(x) for some polynomials q(x), r(x)
in Z[x] with the degree of r(x) less than 2. Then we can see that [p(x)] = [r(x)] since
p(x) − r(x) ∈ (x2 − 2) is a multiple of (x2 − 2). Conversely, for any two polynomials p(x)
and r(x) with degree less than 2, then [p(x)] 6= [r(x)] because p(x) − r(x) has degree less
than 2, and cannot be in the ideal (x2 − 2) unless p(x) − r(x) = 0, as 0 is the only element
of (x2 − 2) of degree less than 2. Therefore p(x) = r(x).
So we can find equivalence class representatives by taking all of the elements of Z[x]
less than degree 2: Z[x]/(x2 − 2) = {[a + bx]|a, b ∈ Z}.
Let’s do a few computations: [x][x] = [x2 ] and we can write x2 = 2 + x2 − 2 using
division. Therefore [x]2 = [2]. √
This ring is actually the same as the ring Z[ 2] because x acts as a root of the poly-
nomial x2 − 2. In general, taking polynomial rings mod an ideal is a nice way to construct
new rings when we want to force x to act a certain way.
11
6 Homework Day 3
6.1 The Basics
1. Prove the proposition from class: if φ : R → S is a homomorphism of rings then
R/ ker(φ) is isomorphic to φ(R).
3. Describe the elements of Z[x]/(x). What ring homomorphism that we have discussed
does this correspond to? What about Z[x]/(x − 3)?
2. Write out the multiplication table for (Z/3Z)[x]/(x2 + 1). How many elements does
this have? Note that this satisfies the field axioms, and is an example of a “finite
field”.
3. Suppose I want to construct a ring such that there is some element which is not zero
but whose square is zero. What quotient of Z[x] would I take?
6.3 To ponder
1. What is the map Fun(R, R) → Fun(R, R)/Mp where Mp is defined as before as the
maximal ideal of functions f : R → R which are zero at p?
12
7 Day 4
7.1 Generating Ideals
Up until now, we’ve been playing a little fast and loose with the notation that we have
been using for ideals. We have been writing ideals using their generators: for example, the
ideal {2n|n ∈ Z} ⊂ Z we write as (2), because it’s in some sense the ideal generated by
the element 2 ∈ Z. How does that work? Well, it’s the smallest ideal that could possibly
contain the element 2, because in order for an ideal of Z to contain 2, it would need to
contain every Z-multiple of 2, and in turn, the set of multiples is closed under addition and
multiplication.
What about the ideal (4, 6) in Z, generated by the two elements 4 and 6? This means
we want to find the smallest ideal containing 4 and 6. Certainly that ring should contain
{4n|n ∈ Z} and {6n|n ∈ Z}. But it also has to contain other elements, like 2, since we can
write 2 = 6 − 4, and thus if 6 and 4 are in my ideal, then so is 2. But then it must also
contain all multiples of 2, that is, (2). However, no sum or product or sum of products of
4 and 6 will ever contain an odd number, so we can conclude (4, 6) = (2). In this case,
we started with an ideal generated by two elements, and did some work to show that it
actually was generated by one element.
Warning! Not every ideal in every ring can be generated by the multiples of just one
element of the ring. This is a little tricky, because in some rings like Z and R[x], it is true
that every ideal is generated by one element, but it is easy to construct a ring in which
this doesn’t hold.
Example 17. Consider the ring R[x, y] polynomials in two variables with coefficients in
R. The ideal (x, y) is not generated by one element.
First of all, what is this ideal? It should be sums of multiples of x and y. But that’s
everything that you can write without a constant term. So (x, y) = {f (x, y)|f (0, 0) = 0}.
Suppose that we did have a way of writing (x, y) = (g) for some g ∈ R[x, y]. Then x ∈ (g)
and y ∈ (g) so we would have to have that both x and y were multiples of g. That would
be impossible, unless g were an element of R, but we know that (x, y) does not contain any
element of R other than 0.
There are plenty more examples exactly like this: we could create ideals like (x−2, y−3)
and (x − 1, y − 5) that are not generated by one element for exactly the same reason.
One more thing to note about generating ideals is that if you have two elements that
differ by a unit (an invertible element of your ring) then they generate the same ideal. So,
for example (−2) = (2) in Z because −1 is a unit, with inverse −1. Or, in R[x], (3x) = (x),
since 3 is a unit with inverse 1/3. In general, if R is a ring, a ∈ R and u ∈ R a unit, then
(ua) = (a) because u−1 (ua) ∈ (ua) and so a ∈ (ua), but clearly also ua ∈ a.
13
7.2 Maximal and Prime Ideals
Let’s switch gears to talk about maximal and prime ideals. Recall from the last homework:
Definition. A maximal ideal M of R is an ideal which is not contained in any proper
ideal.
Definition. A prime ideal P of R is an ideal such that whenever ab ∈ P , with a, b ∈ R,
then a ∈ P or b ∈ P (or both.)
Example 18. Consider the ring R[x]. What are the maximal ideals in this ring? Suppose
we have an ideal generated by one element, f (x) ∈ R[x]. Then (f (x)) is the ideal of
all multiples of f (x), that is, anything we can write as m(x)f (x) for some m(x) ∈ R[x].
When is this contained in another ideal? If f (x) factors, say f (x) = a(x)b(x) then the
ideal (a(x)) contains f (x), because f (x) is a multiple of a(x). That is, if m(x)f (x) is
an element of (f (x)), we can rewrite it as m(x)a(x)b(x), to make it more obvious that
m(x)f (x) ∈ (a(x)). This shows that (f (x)) ⊂ (a(x)) and therefore is not maximal. This
both very easily generalizes to a proof that all maximal ideals are prime, and also suggests
that if our ideal (f (x)) is going to be maximal, then f (x) can’t factor.
Example 19. What about the ring C[x]? If you’re familiar with the fundamental theorem
of algebra, you might know that every polynomial factors if you’re allowed to have coef-
ficients over the complex numbers, so that if f (x) is a degree n polynomial, √ we can write
√
it as A(x − r1 )(x − r2 ) . . . (x − rn ) for A, ri ∈ C, like 3x2 + 5 = 3(x + i5/ 3)(x − i5/ 3).
Thus the only things that generate maximal ideals are polynomials of degree 1, which can’t
factor anymore. (Elements of C will generate the whole ring.)
The maximal ideals of C[x] are of the form (x − a) for a ∈ C (since if we have a degree
one polynomial like (bx − c) we can divide by b and get (x − c/b) which makes the same
ideal.) So we get a bijection between maximal ideals of C[x] and points of the complex
plane.
This is actually a very deep observation, and the connection between points and max-
imal ideals, as well as rings and functions, form the basis of modern algebraic geometry
and radically transformed the ways that we do geometry around the end of the 1800s and
into the beginning of the 1900s. For the remainder of the class, we will explore a little bit
of this connection between ring theory and geometry, as a means to show one reason why
mathematicians have developed a theory of rings.
14
Definition. A curve is the set of points in R2 satisfying the equation f (x, y) = 0 for some
polynomial f (x, y) ∈ R[x, y].
Let’s start with the hyperbola. The equation for the standard hyperbola is y = 1/x or,
to avoid dividing by x, we can write it as xy − 1 = 0. Already we have a familiar object
living in a ring like R[x, y]. What happens when we evaluate functions in R[x, y] on points
of xy − 1? Consider the function x + 2y + 4 ∈ R[x, y]. Then this takes a point (2, 1/2)
on the hyperbola, and adds two times the y-coordinate plus 4 to the x-coordinate to get
2 + 2/2 + 4 = 7. Let’s also consider the function x + 2y + xy + 3 and apply these to some
points:
Hyperbola Point x + 2y + 4 x + 2y + xy + 3
(2, 1/2) 7 7
(3, 1/3) 7.6 7.6
(1/2, 2) 8.5 8.5
(−1, −1) 1 1
Why are these two different functions giving us the same values when we apply them
to points on the hyperbola? We can write x + 2y + xy + 3 = x + 2y + 4 + (xy − 1). For every
point on the hyperbola, xy − 1 = 0 so these have to be the same. So, when two functions
in R[x, y] differ by a multiple of (xy − 1) we can see that they will give the same values on
the the hyperobola.
Therefore, if we only care about functions on our hyperbola, taking the curve to R,
then the functions are R[x, y]/(xy − 1), the equivalence classes of functions, up to adding
a multiple of (xy − 1), which we know doesn’t change the value of the function. Suppose
f (x, y) is the equation for a curve. Then we would like the functions on f (x, y) correspond
to the ring R[x, y]/(f (x, y)).
This is almost right, but there is a slight problem. What if we consider the curve
2
x = 0, which looks like just a copy of the y-axis? If we evaluate functions on the points
of this curve, then the functions x + y and x2 + y take on the same values (namely, just y)
since x2 = 0 implies x = 0, but in the ring R[x, y]/(x2 ) these are two different equivalence
classes. So, instead of taking the ideal (x2 ) we should have taken the ideal (x), which really
represents all of the 2
p functions which are zero on the points where x = 0. In general, we
will need to take (f (x, y)) the radical of the ideal f (x, y), to make sure we include all
functions which are zero on the points of f (x, y) = 0.
Definition. We say an ideal I ⊂ R is radical√ if whenever an ∈ I for n ∈ N, then a ∈ I. If
I is an ideal, then the radical of I, denoted I, is the smallest radical ideal containing I,
which is the set {r|r ∈ R and rn ∈ I for some n ∈ N}.
With this we can define the coordinate ring.
Definition. Let f (x, y) ∈ R[x, y] and consider the curve f (x, y) = 0. Thenpthe functions
on this curve to R form a ring, called the coordinate ring, which is R[x, y]/ (f (x, y)).
15
8 Homework Day 4
8.1 The Basics
√
1. Show that if I ⊂ R is an ideal than what we defined as the radical of I, I = {r|r ∈
R and rn ∈ I for some n ∈ N}, is actually an ideal of R containing I.
2. Exhibit an isomorphism between the coordinate ring of the parabola y = x2 and the
coordinate ring of the x-axis.
4. For the sake of not giving you a very complicated definition, our notion of curve is a
little sillier than most. For example, we allow a “curve” which consists of the points
in the x or y-axis, i.e. {(x, y)|x = 0 or y = 0}. What is the equation of this curve?
Can you think of a curve that only has two points in it?
(To avoid this non-intuitive definition, some people define curve to be the set of pairs
of complex numbers satisfying a polynomial which doesn’t factor, or the zeroes in R2
of a polynomial which doesn’t factor and has infinitely many points in R2 . Both of
these will lead to more familiar shapes.)
5. If you missed this problem on the previous homework, use your improved under-
standing of generators and ideals to compute the elements of the ideal (8, 6) ⊂ Z and
write it as an ideal with one generator. Do this also with (10, 15) and predict the
answer with (m, n).
8.3 To ponder
1. Some ideals aren’t finitely generated, for example, the ring of polynomials R[x1 , x2 , x3 , . . . ]
in infinitely many variables xi , i ∈ N has the ideal (x1 , x2 , x3 , . . . ) which is not finitely
generated. A ring in which every ideal is finitely generated is called Noetherian. Show
that if R is a Noetherian ring then any chain of ideals I1 ( I2 ( I3 ( . . . must be
finite.
16
9 Day 5
Yesterday, we left off by defining the coordinate ring ofp a curve, which we said was the ring
of
p distinct polynomial functions on the curve R[x, y]/ (f (x, y) where we are thinking of
(f (x, y) as the ideal of functions which are zero on the curve f (x, y) = 0.
So, we have found a way to associate a ring to a curve, but this would be more useful
if we knew the ring encoded other information about the curve. It’s hopeless to think that
the ring completely defines the curve and vice verse: on the homework, we saw that even
though the parabola and the x-axis are different curves, they have isomorphic coordinate
rings.
In fact, we can make ring-like objects which completely encapsulate the geometric
information about the curves we care about. This was a major innovation of 20th century
algebraic geometry pioneered by Grothendieck. He showed that something called “locally
ringed spaces” can be assigned to geometric objects which contain almost all of the data
that we care about. The basic idea that for any geometric object, like a curve, or a
hypersurface, or something which we have glued together which is locally the zero-set of
polynomials in n-dimensions, we can assign a topology, or notion of closeness/far away-
ness, and to every neighborhood of the space we can assign a ring which consists of the
functions of the geometric object on that neighborhood. At each point, we assign a special
kind of ring, called a local ring, which only has one maximal ideal, corresponding to the
functions that are zero at that point.
Today, though, we’re going to try to show the correspondence between algebra and
geometry in the two most important properties of a curve: points on the curve (which will
correspond to maximal ideals) and maps between curves (which will correspond to maps
between rings.)
17
Note that (x−2)(y−1/2) = xy+1−1/2x−2y, and xy+1−1/2x−2y+1/2(x−2) = xy−2y
and finally xy − 2y + 2(y − 1/2) = xy − 1 so the ideal (xy − 1) ⊂ (x − 2, y − 1/2). But, we
should have predicted that because (xy − 1) is the ideal of polynomial functions which are
zero on the whole curve xy − 1, while (x − 2, y − 1/2) is the ideal of functions which are
zero on the point (2, 1/2), and since (2, 1/2) is a point on the curve xy − 1, any function
in the first ideal must be in the second. However, it’s nice to see this algebraically.
So we have a series of maps where we take
first, sending all the functions which are zero on xy − 1 to zero, then sending all functions
which are zero on the point (2, 1/2) to zero. We had to check that this made sense, that the
first ideal was contained in the second, so that it makes sense to take successive quotients.
In general, if (a, b) is a point on our curve f (x, y) = 0,p
we can look at the maximal
ideal M(a,b) = (x − a, x − b) in the coordinate ring R[x, y]/ (f (x, y)) and we get a way
of turning points of our curve into maximal ideals. (Note that this p correspondence isn’t
exact: in general, there will be some maximal ideals of R[x, y]/ (f (x, y)) which do not
correspond to points, and this comes from the fact that we are working over R2 and that
means we’re missing some points that we can’t see because they lie in C2 . In the case
where we look at C instead of R, we get an exact correspondence between maximal ideals
and points.)
18
Remark. We’re sweeping a lot under the rug here! How do we know that g ◦ π is even
a polynomial? Well, π takes (a, b) to (a, 0) and then g is a polynomial in x and y. So
when we compose, g we plug (a, 0) into g instead of (a, b), so it is the same as taking g and
letting y = 0 everywhere.
Let’s do an example using our hyperbola projecting to the x-axis map. Consider the
function g(x, y) ∈ R[x, y]/(y) which is g(x, y) = 2x + y − 1. Where does this go under the
map π? Well we just precompose with π: so g ◦ π(x, y) = g(x, 0) = 2x − 1. Now if (2, 1/2)
is a point on the hyperbola, it goes to 4 − 1 = 3, or, instead, if we first mapped it to the
x-axis point (2, 0), then we could apply g to get 4 + 0 − 1 = 3.
Where do the functions x and y in R[x, y]/(y) go? Well, x 7→ x and y 7→ 0 according
to our calculations. So if we want to describe the image of the map π ∗ we can see that
π ∗ : R[x, y]/(y) → R[x, y]/(xy − 1) has image which is just all polynomials which are only
in x.
19