This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

A bottom-up approach to Hilbert’s Basis Theorem

Marc Maliar
Abstract.

In this expositional paper, we discuss commutative algebra—a study inspired by the properties of integers, rational numbers, and real numbers. In particular, we investigate rings and ideals, and their various properties. After, we introduce the polynomial ring and the fundamental relationship between polynomials and sets of points. We prove some results in algebraic geometry, notably Hilbert’s Basis Theorem.

1. Introduction

Mathematics often begins with concrete problems, for example finding the length of the hypotenuse of a triangle. But as a mathematician investigates such problems, they begin to abstract away various amounts of details, uncovering patterns and defining properties.

In this expositional paper, we take this approach to commutative algebra.

In Section 22, we attempt to find the patterns and defining properties hidden in algebra. In particular, we investigate sets of numbers such as the integers, rational numbers and real numbers. We define a ring, the generalization of the properties of such sets of numbers. We realize that there exist special sets in the integers called ideals—named in this way because they behave in an “ideal” way. Finally, we explore properties of these rings and ideals.

But we realize that polynomials behave as integers do: we can add and multiply polynomials, identify “prime” polynomials, and find ideals. So we begin Section 33 by defining the polynomial ring. Unlike the integers, though, polynomials have an interesting relationship with sets of points in a field. And so the rest of section 33 is devoted to applying commutative algebra to this relationship between sets of points and polynomials. For example, we establish that, given a set of polynomials SS in two variables (or any number of variables actually), the set of points X=(x,y)X=(x,y) such that f(x,y)=0f(x,y)=0 for all fSf\in S can be written as the intersection of finitely many sets of points X1,,XnX_{1},\dots,X_{n}, where Xi=(x,y)X_{i}=(x,y) such that fi(x,y)=0f_{i}(x,y)=0 for some polynomial fif_{i}. This result is a corollary of Hilbert’s Basis Theorem (Theorem 3.11).

By the end, we are able to use abstractions to prove nontrivial theorems about sets of points and polynomials.

2. Rings and ideals

We begin this section by exploring rings. What is a ring? Consider the set of integers, \mathbb{Z}. Recall that this fundamental set comes equipped with two binary operations (the usual addition and multiplication). We generalize the properties of this set and arrive at the definition of a ring.

We then consider special subsets of a ring, called ideals. We motivate this study by again considering \mathbb{Z}. In particular, we look at sets of integers that are divisible by a certain integer, and we note that they possess interesting properties. We then define ideals by these interesting properties. In the rest of the section, we discover other interesting properties of ideals and investigate related definitions.

2.1. Definitions of a ring

Definition 2.1.

A ring RR is a set with two binary operations addition and multiplication, denoted by ++ and * respectively. These operations have the following properties:

  1. (1)

    Addition is associative and commutative; there exists an additive identity (denoted by 0); each element in the ring has a (unique) additive inverse.

  2. (2)

    Multiplication is associative; there exists a multiplicative identity (denoted by 11).

  3. (3)

    Multiplication is distributive over addition. That is, r1(r2+r3)=r1r2+r1r3r_{1}*(r_{2}+r_{3})=r_{1}*r_{2}+r_{1}*r_{3} and (r1+r2)r3=r1r3+r2r3(r_{1}+r_{2})*r_{3}=r_{1}*r_{3}+r_{2}*r_{3} hold for r1,r2,r3Rr_{1},r_{2},r_{3}\in R.

Why such a definition? The ring definition comes from trying to generalize numbers (integers, rational numbers, and real numbers) and their various properties. Notice that the aforementioned sets of numbers with their usual operations satisfy these properties.

We consider commutative rings: rings in which multiplication is commutative. From now on, we consider only commutative rings; any ring mentioned is commutative. Notice that the integers, rational numbers, and real numbers are commutative rings.

While each element in a ring has an additive inverse, it does not necessarily have a multiplicative inverse. In addition, 11 can be equal to 0. In this case, it is easy to see that the ring has only one element, 0 (and that this set of one element is a ring).

Recall the classical definition of a field: a set with addition and multiplication satisfying the usual properties. Fields turn out to be valuable objects in algebraic geometry, so I mention them here. Also, in the next section, we prove some results about commutative rings that are fields, so I introduce the following (equivalent) definition:

Definition 2.2.

A field is a commutative ring kk where 101\neq 0 and every element has a multiplicative inverse.

2.2. Definition of an ideal

Consider the ring of integers; now, consider the set of all elements divisible by 33 (a certain subset of this ring). This set has some interesting properties.

First, if you add any two elements divisible by 33, you end up with an integer divisible by 33. Second, if you multiply an element in the set with any other integer, you end up with an integer divisible by 33. Ideals are defined by these two properties.

Definition 2.3.

An ideal II is a subset of a ring RR that, inheriting the addition and multiplication operations from RR, satisfies the following properties:

  1. (1)

    Closure with respect to addition (if i1,i2Ii_{1},i_{2}\in I then i1+i2Ii_{1}+i_{2}\in I); associativity with respect to addition; additive identity is in the ideal (0I0\in I); each element’s additive inverse is also in the ideal (if iIi\in I, iI-i\in I).

  2. (2)

    If rRr\in R and iIi\in I, riIr*i\in I.

  3. (3)

    If rRr\in R and iIi\in I, irIi*r\in I.

Remark 2.4.

Since we work with commutative rings, it is only necessary to prove either (2)(2) or (3)(3). In non-commutative rings, an ideal that satisfies (2)(2) but not necessarily (3)(3) is a left-ideal, and an ideal that satisfies (3)(3) but not necessarily (2)(2) is a right-ideal.

Note that we do not need to explicitly prove all of these statements. Indeed, proving closure with respect to addition, existence of additive inverse, and (2)(2) (or (3)(3) as discussed) is sufficient.

  1. (a)

    Associativity is inherited from the ring definition.

  2. (b)

    For iIi\in I, i(1)=iIi*(-1)=-i\in I.

An ideal cannot be empty because it must contain 0.

Let us find other ideals of the rings we have mentioned (integers, rational numbers, and real numbers). To begin, note that, for any ring, both {0}\{0\} and RR are ideals.

But are these the only ideals? Consider first the integers.

Proposition 2.5.

For all zz\in\mathbb{Z}, define Iz={zrr}I_{z}=\{z*r\mid r\in\mathbb{Z}\}. IzI_{z} is an ideal. These are the only ideals of \mathbb{Z}.

Proof.

The proof that IzI_{z} is an ideal follows from my previous remark. So then suppose there exists an ideal II that cannot be written as an IzI_{z}. I{0}I\neq\{0\} because then I=I0I=I_{0}, so there exists at least one nonzero integer in II. Consider the set of nonzero positive integers in II. This set must be nonempty; an empty set would imply that the nonzero integer zz we found earlier is negative, but z(1)Iz*(-1)\in I—a contradiction. By the Well-Ordering Principle, find the smallest positive nonzero integer zIz\in I.

Because IIzI\neq I_{z}, there must exist an element zIz^{\prime}\in I that is not in the form zrz*r for any rr\in\mathbb{Z}. Suppose it is positive (if it is negative multiply it by 1-1 and it will still be in the ideal). We know that z>zz^{\prime}>z because zz is the smallest positive integer. Now, observe that z+(z)z^{\prime}+(-z) is still in the ideal and also will not be in the form zrz*r. Repeat until the result is less than zz. The result will be in II, positive and nonzero (because it is not in the form zrz*r). Thus, we have found a smaller positive nonzero integer in II that is smaller than zz—a contradiction. ∎

In other words, the only ideals of the ring of integers are multiples of an integer. Now, consider rational numbers and real numbers. It turns out that we can make a statement about fields in general:

Proposition 2.6.

The only ideals of a field kk are the field itself and the zero ideal. The converse is also true: if a commutative ring has exactly two ideals, the zero ideal and the entire ring, then the commutative ring is a field.

Proof.

First, we prove that that the only ideals of kk are the field itself and the zero ideal. Let II be an arbitrary ideal.

  1. Case 1:

    II contains no non-zero elements. Then, I={0}I=\{0\}.

  2. Case 2:

    II contains a non-zero element. Let ii be this element. Then, because kk is a field, there exists a multiplicative inverse of ii, denoted by i1i^{-1}. By the property of ideals, ii1=1Ii*i^{-1}=1\in I. Any element jkj\in k is in the ideal now because 1j=jI1*j=j\in I, so then I=kI=k.

Second, we prove that if the only ideals of a ring RR are the ring and the zero ideal, the ring is a field. Let rr be a non-zero element in our commutative ring RR. Construct a set I={rqqR}I=\{rq\mid q\in R\}. It is easy to check that this is an ideal. Because this ideal contains a nonzero element rr, it cannot be the zero ideal, so by our given, the ideal must be the entire ring. Then, there exists an element rr^{\prime} such that r(r)=1r(r^{\prime})=1 since 11 is in our ideal. Thus, rr has an inverse. ∎

From this proof, we can also see the following statement:

Corollary 2.7.

Let II be an ideal of RR. Then, 1II=R1\in I\iff I=R.

2.3. Another definition of an ideal

Consider the following definition:

Definition 2.8.

A ring homomorphism is a map ϕ\phi from one ring RR to another ring SS with the following properties:

  1. (1)

    ϕ(r1+r2)=ϕ(r1)+ϕ(r2)\phi(r_{1}+r_{2})=\phi(r_{1})+\phi(r_{2}),

  2. (2)

    ϕ(r1r2)=ϕ(r1)ϕ(r2)\phi(r_{1}*r_{2})=\phi(r_{1})*\phi(r_{2}), and

  3. (3)

    ϕ(1R)=1S\phi(1_{R})=1_{S}, where 1R,1S1_{R},1_{S} are the multiplicative identities of RR and SS respectively.

We introduce an equivalent definition of the ideal using ring homomorphisms:

Definition 2.9.

An ideal II is a subset of a ring RR that is the kernel of a ring homomorphism. That is, a set IRI\subset R is an ideal if there exists a ring homomorphism ϕ:RS\phi\colon R\to S, where SS is some other ring, such that ϕ1(0)=I\phi^{-1}(0)=I.

Proposition 2.10.

Definition 2.3 is equivalent to Definition 2.9.

Proof.

First, we show Definition 2.9 from Definition 2.3. We need to find a ring SS and a function ϕ\phi as described. We need ϕ\phi such that for all iIi\in I, ϕ(i)=0\phi(i)=0. Furthermore, we need ϕ\phi such that adding iIi\in I to any rRr\in R will not change the computed result (i.e., ϕ(r)=ϕ(r+i)\phi(r)=\phi(r+i)).

Let rr be an element in RR. Define ϕ:rXr:={r+i:iI}\phi\colon r\mapsto X_{r}:=\{r+i\ :\ i\in I\}. Define R/I={XrrR}R/I=\{X_{r}\mid r\in R\}.

We now define an equivalence relation between elements in RR. We say that rqr\sim q for r,qRr,q\in R if and only if ϕ(r)=ϕ(q)\phi(r)=\phi(q). It follows from properties of an ideal that \sim is an equivalence relation. Notice that, under this definition, ϕ(i)=ϕ(0)iI\phi(i)=\phi(0)\;\forall\;i\in I.

Define addition and multiplication as follows: for ϕ(r),ϕ(q)R/I\phi(r),\phi(q)\in R/I, let ϕ(r)+ϕ(q)=ϕ(r+q)\phi(r)+\phi(q)=\phi(r+q) and ϕ(r)ϕ(q)=ϕ(rq)\phi(r)*\phi(q)=\phi(r*q). Both of these are well-defined. Consider, for example, addition. Let r1,r2,q1,q2r_{1},r_{2},q_{1},q_{2} be elements in RR with r1r2r_{1}\sim r_{2} and q1q2q_{1}\sim q_{2}. Let i,ji,j be elements in II such that r1=r2+ir_{1}=r_{2}+i and q1=q2+jq_{1}=q_{2}+j. Then,

ϕ(r1)+ϕ(q1)=ϕ(r1+q1)=ϕ(r2+q2+i+j)=ϕ(r2+q2)=ϕ(r2)+ϕ(q2).\phi(r_{1})+\phi(q_{1})=\phi(r_{1}+q_{1})=\phi(r_{2}+q_{2}+i+j)=\phi(r_{2}+q_{2})=\phi(r_{2})+\phi(q_{2}).

We conclude that equivalent inputs implies equivalent outputs under our addition operation. Multiplication follows similarly as well.

Under these operations, R/IR/I forms a ring. To finish the proof, define S=R/IS=R/I and ϕ\phi as defined above. By our definition of ϕ\phi, we already know that ϕ(r1+r2)=ϕ(r1)+ϕ(r2)\phi(r_{1}+r_{2})=\phi(r_{1})+\phi(r_{2}) and ϕ(r1r2)=ϕ(r1)ϕ(r2)\phi(r_{1}*r_{2})=\phi(r_{1})*\phi(r_{2}). It suffices to show that ϕ(1R)=1S\phi(1_{R})=1_{S} where 1S1_{S} is the multiplicative identity of R/IR/I. The proof follows from the definition.

Second, we show Definition 2.3 from Definition 2.9. As discussed, we must prove only three statements:

  1. (a)

    Closure with respect to addition: let i1,i2Ii_{1},i_{2}\in I. We know ϕ(i1+i2)=ϕ(i1)+ϕ(i2)=0+0=0\phi(i_{1}+i_{2})=\phi(i_{1})+\phi(i_{2})=0+0=0. Then, i1+i2Ii_{1}+i_{2}\in I because ϕ(i1+i2)=0\phi(i_{1}+i_{2})=0.

  2. (b)

    Existence of additive inverse: ϕ(0)=0\phi(0)=0 so 0I0\in I.

  3. (c)

    If iIi\in I and rRr\in R then irIi*r\in I: let iI,rRi\in I,r\in R. ϕ(ir)=ϕ(i)ϕ(r)=0ϕ(r)=0\phi(i*r)=\phi(i)*\phi(r)=0*\phi(r)=0. Then, irIi*r\in I because ϕ(ir)=0\phi(i*r)=0.

Consider the ideals of \mathbb{Z} we have found. What is the ring SS and ring homomorphism ϕ\phi that we should be able to find, according to our new definition of an ideal?

  1. (1)

    Iz={nzz}I_{z}=\{nz\mid z\in\mathbb{Z}\}: we use S=/nS=\mathbb{Z}/n\mathbb{Z} (SS is the set of integers modulo nn—a ring) and ϕ\phi where ϕ(r)=rmodn\phi(r)=r\mod n. The set of elements that map to 0 is {nzz}\{nz\mid z\in\mathbb{Z}\}.

  2. (2)

    I={0}I=\{0\}: we use S=RS=R and ϕ\phi where ϕ(r)=r\phi(r)=r. The only element that maps to 0 is 0.

  3. (3)

    I==RI=\mathbb{Z}=R: we use S={0}S=\{0\} and ϕ\phi where ϕ(r)=0\phi(r)=0. All elements map to 0.

2.4. What is in an ideal?

Consider an ideal II of ring RR. How can we imagine what is in this ideal?

Let us try to come up with an expression that can represent any element in the ideal. Let II be an ideal and let {iμ}\{i_{\mu}\} be a subset of II indexed by μ\mu. Consider the expression μiμrμ\sum_{\mu}i_{\mu}*r_{\mu} where iμi_{\mu} is an element of the ideal and rμr_{\mu} is an element in RR. Can we write any element in II in this form for some rμr_{\mu}’s? If so, we denote the ideal II by

I=({iμ})I=(\{i_{\mu}\})

and say that it is generated by {iμ}\{i_{\mu}\}. For example, I=R=(1)I=R=(1) (as we saw in Corollary 2.7). So, now we have come up with a way to imagine elements in our ideal.

Some things to emphasize: in rings that extend outwards infinitely (like the integers and the ring of polynomials introduced later), non-zero ideals grow large up to arbitrary finite size. For example, with the ideal of integers divisible by 33, we can keep multiplying and finding larger and larger elements.

In addition, note that just because we can write an element iIi\in I as a sum of iμrμi_{\mu}*r_{\mu}’s doesn’t mean that for any two elements a,bRa,b\in R such that ab=ia*b=i, either aa or bb in II. An ideal in which this is true for every iIi\in I is called a prime ideal:

Definition 2.11.

A prime ideal II of ring RR is an ideal that satisfies the following properties:

  1. (1)

    If kIk\in I and ij=ki*j=k, then either iIi\in I or jIj\in I.

  2. (2)

    II is not equal to the ring RR.

Consider the ideal of integers divisible by 33 in the ring of integers: this ideal is a prime ideal because we can write every element as 33 multiplied by an integer, and 33 is in our ideal. However, the ideal of integers divisible by 66 does not have this property; for example 6=236=2*3 but neither 22 nor 33 are divisible by 66. The generalization of these statements corresponds to the following statement:

Proposition 2.12.

In \mathbb{Z}, ideals generated by a prime number are prime. The zero ideal is also prime. All other ideals are not prime.

Proof.

Let II be an ideal generated by a prime natural number pp. Suppose II is not prime. Then, there exists an element kIk\in I and i,jIi,j\in I such that ij=ki*j=k but iIi\not\in I and jIj\not\in I. This means that kk is divisible by pp but neither ii nor jj are divisible by pp, which is impossible.

The zero ideal is prime because no integer divides 0.

We already showed that all ideals of \mathbb{Z} are generated by one element, and so it is only necessary to look at ideals generated by nonprime elements. Let II be an ideal generated by a nonprime natural number nn. Because nn is not prime, n=mn=\ell*m for some ,m\ell,m not divisible by nn, so \ell and mm are not in II; then, II is not prime. The ideal II generated by n-n is the same as the one generated by nn so the same logic applies. ∎

We have an alternate definition of prime ideal as well. This definition relates to Definition 2.11 similarly to how Definition 2.9 relates to Definition 2.3.

Definition 2.13.

A prime ideal II of a ring RR is the kernel of a ring homomorphism ϕ:RS\phi\colon R\rightarrow S, where SS is a ring that is not equal to {X0}\{X_{0}\} and has no zero-divisors (no elements ss such that for some tSt\in S, st=0s*t=0).

Proposition 2.14.

Definition 2.11 is equivalent to Definition 2.13.

Proof.

First, we prove that Definition 2.11 is equivalent to Definition 2.13.

Let II be a prime ideal by Definition 2.11. Suppose that Definition 2.13 does not hold. By Definition 2.9, there must exist a set SS and ϕ\phi such that II is the kernel of ϕ\phi. Then, take SS and ϕ\phi that Definition 2.13 refers to. Consider two cases:

  1. Case 1:

    S={X0}S=\{X_{0}\}. II is not equal to RR. Then, there exists an element rr in RR that is not in II, and ϕ(r)X0\phi(r)\neq X_{0}.

  2. Case 2:

    SS has a zero-divisor. Let ss be a zero-divisor that is not equal to 0 with a corresponding tt that is not equal to 0 such that sts*t is equal to 0. By construction, ϕ\phi is surjective, so there exist r,qRr,q\in R such that ϕ(r)=s\phi(r)=s and ϕ(q)=t\phi(q)=t. Then,

    0=st=ϕ(r)ϕ(q)=ϕ(rq),0=s*t=\phi(r)*\phi(q)=\phi(r*q),

    so rqIr*q\in I. However, rIr\not\in I—otherwise ϕ(r)=0\phi(r)=0 (and similar argument for ss). Then, II is not prime according to Definition 2.11, so we have arrived at a contradiction.

Second, we prove that Definition 2.13 is equivalent to Definition 2.11.

For sake of contradiction, let II be an ideal that is not prime by Definition 2.11. By Definition 2.9, there exists a set SS and ϕ\phi such that II is the kernel of ϕ\phi. Consider two cases:

  1. Case 1:

    I=RI=R. Then, every element iIi\in I maps to 0, so S={X0}S=\{X_{0}\}.

  2. Case 2:

    There exist elements r,qRr,q\in R such that rqr*q is in II but neither rr nor qq is in II. Then,

    0=ϕ(rq)=ϕ(r)ϕ(q),0=\phi(r*q)=\phi(r)*\phi(q),

    but neither ϕ(r)\phi(r) nor ϕ(q)\phi(q) are equal to 0 because neither rr nor qq are in II. Then, both ϕ(r)\phi(r) and ϕ(q)\phi(q) are zero-divisors of SS so we have arrived at a contradiction.

There are many other interesting ways to characterize rings and ideals. See [4, Chapter 1].

2.5. Finitely generated ideals

Let us go back to our equation for elements in an ideal. Can we make some claim about the sets iμi_{\mu} that generate II?

Note that such a set must exist. If we choose {iμ}=I\{i_{\mu}\}=I then II is generated by {iμ}\{i_{\mu}\}. Let iλi_{\lambda} be an element in II. Set rμ=1r_{\mu}=1 for μ=λ\mu=\lambda and rμ=0r_{\mu}=0 for μλ\mu\neq\lambda.

But can we find other sets? Specifically, can we find a finite set {iμ}\{i_{\mu}\} that generates II? Such an ideal is said to be finitely generated. Furthermore, for a ring RR, is every ideal IRI\subset R finitely generated? Such a ring is said to be a Noetherian ring.

Consider the ring of integers. Since every ideal of \mathbb{Z} consists of all multiples of some element, every ideal II is finitely generated and \mathbb{Z} is Noetherian.

Later, I will give an example of a ring that is not Noetherian.

There are two other alternate definitions of Noetherian rings:

Proposition 2.15.

The following are equivalent:

  1. i

    For every ideal IRI\subset R, there exist x1,,xnIx_{1},\dots,x_{n}\in I such that I=(x1,,xn)I=(x_{1},\dots,x_{n}).

  2. ii

    Every chain of ideals I1IkI_{1}\subset\dots\subset I_{k}\dots eventually terminates such that I=I+1=I_{\ell}=I_{\ell+1}=\dots.

  3. iii

    Every nonempty set of ideals of RR has an ideal that is not contained in any other.

Proof.

(i) \Rightarrow (ii): Take a chain of ideals, and let II be the union of them all. For II, there exist x1,,xnx_{1},\dots,x_{n} that generate II. All of these xix_{i} should appear at some point in the chain of ideals. When they appear, they stay in the chain because each ideal contains all previous ideals. Because there are finitely many xix_{i}, we can find an II_{\ell} for some \ell that finally contains all of them. It is clear that I=II_{\ell}=I because they are generated by the same elements. To prove the chain terminates, let kk be a natural number bigger than ll. IlI_{l} is a subset of II and therefore is a subset of IkI_{k}. IkI_{k} is a subset of IlI_{l} because we have a chain of ideals. Then, Il=IkI_{l}=I_{k}, so the chain terminates.
  (ii) \Rightarrow (iii): At the ideal at which the chain terminates, that ideal is not contained in any other.
  (iii) \Rightarrow (i): Let II be an ideal and Σ={JIJ is a finitely generated ideal}\Sigma=\{J\subset I\mid J\text{ is a finitely generated ideal}\}. By our given, JJ has an element J0J_{0} that is not contained in any other. Suppose that J0IJ_{0}\neq I. Then, there exists xI,xJ0x\in I,x\not\in J_{0}. Let J0J_{0}^{\prime} be the ideal generated by all the finitely many generators in J0J_{0} and also xx. J0J_{0}^{\prime} is still finitely generated and contains J0J_{0}, contradicting our assumption that J0J_{0} is not contained in any other. Then, I=J0I=J_{0} is finitely generated. ∎

3. Algebraic varieties and the polynomial ring

We begin our study of a polynomial ring, points, and the relationship between them. We motivate our study first by introducing the plane curve, a mathematical object studied heavily in the 1919th century. Next, we formally define polynomials and the fundamental functions “taking the variety of” and “taking the ideal of” that map polynomials to points and vice-versa, respectively. After, we investigate how these two functions serve as inverses to each other. We then attempt to understand the nature of plane curves (and algebraic sets in general). In particular, we prove Hilbert’s Basis Theorem. Finally, we investigate bijective correspondences between sets of polynomials and points.

3.1. Introducing plane curves

Definition 3.1.

Let kk be a field and f:k2kf\colon k^{2}\rightarrow k be a polynomial. A plane curve XX is defined as the set of all points {(x,y)}k2\{(x,y)\}\in k^{2} that satisfy f(x,y)=0f(x,y)=0 (such a function is said to “vanish” on XX).

Figure 1. The plane curve defined by k=k=\mathbb{R} and f(x,y)=y2x2(x+1)f(x,y)=y^{2}-x^{2}(x+1).
Refer to caption

We are usually familiar with finding the zeros of a polynomial f(x)f(x) by setting f(x)=0f(x)=0, and we know that there are finite number of such xx (over a field). Now, we have f(x,y)f(x,y), and there can be an infinite number of (x,y)(x,y) that satisfy f(x,y)=0f(x,y)=0 (as seen in Figure 1). Algebraic geometry is concerned with studying the relationship between these polynomial functions and sets of points.

3.2. The polynomial ring

We introduce machinery for understanding how to work with these polynomial equations.

Definition 3.2.

Let kk be a field, nn be a natural number, and AA be n\mathbb{N}^{n}. The polynomial ring in nn variables k[X1,,Xn]k[X_{1},\dots,X_{n}] is the set of all expressions of the form {αAaαXα}\{\sum_{\alpha}^{A}a_{\alpha}X^{\alpha}\} where aαka_{\alpha}\in k for all αA\alpha\in A and Xα=X1α1XnαnX^{\alpha}=X_{1}^{\alpha_{1}}\dots X_{n}^{\alpha_{n}} where αi\alpha_{i} is the iith component in α\alpha. A polynomial in nn variables is an element of the polynomial ring. Polynomials are finite sums so we impose the additional restriction that aα=0a_{\alpha}=0 for all but finitely many αA\alpha\in A. Let {αAaαXα}\{\sum_{\alpha}^{A}a_{\alpha}X^{\alpha}\} and {αAbαXα}\{\sum_{\alpha}^{A}b_{\alpha}X^{\alpha}\} be two polynomials. Addition of two polynomials is defined as {αA(aα+bα)Xα}\{\sum_{\alpha}^{A}(a_{\alpha}+b_{\alpha})X^{\alpha}\}. Multiplication is defined as {αAΠβA(aαbα)Xαβ}\{\sum_{\alpha}^{A}\Pi_{\beta}^{A}(a_{\alpha}*b_{\alpha})X^{\alpha*\beta}\} where Xαβ=X1α1β1XnαnβnX^{\alpha*\beta}=X_{1}^{\alpha_{1}*\beta_{1}}\dots X_{n}^{\alpha_{n}*\beta_{n}}. One can prove that the polynomial ring is a ring by Definition 2.1.

Note that all concepts we defined about rings and ideals in general apply now to polynomial rings. For example, the polynomial rings has ideals, ring homomorphisms, generating sets, and prime ideals.

3.3. The relationship between points and polynomials

In the beginning of this section, we defined a plane curve. In the last subsection, we generalized the algebraic aspect of plane curves by defining polynomials in nn variables. Now, we generalize the “set of points” aspect of plane curves.

Definition 3.3.

Let V:k[X1,,Xn]knV\colon k[X_{1},\dots,X_{n}]\to k^{n} be a function where V(S)={xknf(x)=0fS}V(S)=\{x\in k^{n}\mid f(x)=0\;\forall\;f\in S\}. V(S)V(S) is known as the variety of SS. XknX\subset k^{n} is an algebraic set if X=V(S)X=V(S) for some Sk[X1,,Xn]S\subset k[X_{1},\dots,X_{n}].

Remark 3.4.

Note that we are evaluating fk[X1,,Xn]f\in k[X_{1},\ldots,X_{n}] on a point xknx\in k^{n}. This is done using the field operations of kk.

Under this definition, a plane curve is just an algebraic set created from only one (ff) two-dimensional (variables x,yx,y) polynomial.

We introduce the following proposition that follows from the definition:

Proposition 3.5.

=V({k[X1,,Xn]})\emptyset=V(\{k[X_{1},\dots,X_{n}]\}) and kn=V({0})k^{n}=V(\{0\}) are algebraic sets.

When one takes the variety, one maps polynomials to points. Now, we consider a map which takes points to polynomials:

Definition 3.6.

Let XX be a subset of knk^{n} (not necessarily an algebraic set). The ideal of XX, denoted by I(X)I(X), is the set of polynomials in nn variables that vanish on XX.

Proposition 3.7.

I(X)I(X) is an ideal.

Proof.

Follows from field operations. ∎

It is clear that the variety of a set of polynomials and the ideal of set of points are somehow fundamentally linked. In which way are these two functions “inverses” of each other?

Proposition 3.8.

The correspondences VV and II satisfy the following properties for any field kk:

  1. (1)

    Let X,YX,Y be subsets of knk^{n}. If XYX\subset Y, I(Y)I(X).I(Y)\subset I(X).

  2. (2)

    Let S,TS,T be subsets of k[X1,,Xn]k[X_{1},\dots,X_{n}]. If STS\subset T, V(T)V(S)V(T)\subset V(S).

  3. (3)

    Let XX be a subset of knk^{n}. Then, XV(I(X))X\subset V(I(X)). X=V(I(X))X=V(I(X)) if and only if XX is an algebraic set.

  4. (4)

    Let SS be a subset of k[X1,,Xn]k[X_{1},\dots,X_{n}]. Then, SI(V(S))S\subset I(V(S)).

Proof.

(1)(1) and (2)(2) follow from the definition.

For (3)(3), note that I(X)I(X) is defined as the set of all polynomials that vanish on XX, so these polynomials will vanish on xXx\in X. For the second part of (3)(3), we know that X=V(I(X))X=V(I(X)); then, X=V(S)X=V(S) for S=I(X)S=I(X), meaning that XX is an algebraic set. Next, let us prove that XX is an algebraic set implies that X=V(I(X))X=V(I(X)). X=V(S)X=V(S) for some SS. Suppose, for sake of contradiction, that I(X)I(X) does not contain SS. Then, there exists a polynomial fS,fI(X)f\in S,f\not\in I(X). We know that f(x)=0f(x)=0 for all xXx\in X by definition, so then fI(X)f\in I(X)—a contradiction. Because SI(X)S\subset I(X), V(I(X))V(S)=XV(I(X))\subset V(S)=X so V(I(X))=XV(I(X))=X (we already know XV(I(X))X\subset V(I(X)) because of the first part of (3)(3)).

For (4)(4), let SS be such a set and let fSf\in S. By definition, we know that ff vanishes at the points at which SS vanishes. ∎

We figured out under what conditions X=V(I(X))X=V(I(X)) for (3)(3) of the proposition. But we did not figure out under what conditions S=I(V(S))S=I(V(S)). We discuss it later; it is the motivation behind Hilbert’s Nullstellensatz.

Finally, we introduce some operations on elements in the polynomial ring, and what the resulting algebraic set. We hope that these examples provide some intuition on “taking the variety”.

Proposition 3.9.

Let f1,f2k[X1,,Xn]f_{1},f_{2}\in k[X_{1},\dots,X_{n}]. Then,

  1. (1)

    V(f1f2)=V(f1)V(f2)V(f_{1}*f_{2})=V(f_{1})\cup V(f_{2}) and

  2. (2)

    V((f1,f2))=V(f1)V(f2)V((f_{1},f_{2}))=V(f_{1})\cap V(f_{2}).

Proof.

Both follow from the definition. ∎

3.4. Other properties of algebraic sets

Before continuing with relationships between points and polynomials, let us come back briefly to algebraic sets. First, consider these two statements:

Proposition 3.10.

Algebraic sets have the following properties:

  1. (1)

    The union of finitely many algebraic sets is an algebraic set.

  2. (2)

    The intersection of any collection of algebraic sets is an algebraic set.

Proof.
  1. (1)

    Let X=V(S)X=V(S) and Y=V(T)Y=V(T) be algebraic sets for some sets of polynomials SS and TT. Indeed, XY=V(ST)X\cup Y=V(ST), where STST denotes the set of all products of an element of SS by an element of TT. Let xx be an element in XYX\cup Y. Then, xXx\in X or xYx\in Y. Without loss of generality, suppose xXx\in X. Then, f(x)=0f(x)=0 for all fSf\in S so every function in STST vanishes on xx. Now, let xV(ST)x\in V(ST). Without loss of generality, suppose that xXx\not\in X. Then, there is an fSf\in S such that f(s)0f(s)\neq 0. All polynomials fgSTfg\in ST, where gTg\in T, vanish on xx, so then g(t)=0g(t)=0 for all gTg\in T and xx vanishes on YXYY\subset X\cup Y.

    By induction, the union of finitely many algebraic sets is an algebraic set as well.

  2. (2)

    Let {Xα=V(Sα)}\{X_{\alpha}=V(S_{\alpha})\} be a collection of algebraic sets indexed by α\alpha. Then, αXα=V(αXα)\bigcap_{\alpha}X_{\alpha}=V(\bigcup_{\alpha}X_{\alpha}). Let xαXαx\in\bigcap_{\alpha}X_{\alpha}. Then, xXαx\in X_{\alpha} for all α\alpha so xV(Sα)x\in V(S_{\alpha}) for all α\alpha. Now, let xV(αXα)x\in V(\bigcup_{\alpha}X_{\alpha}). By definition, xV(Xα)x\in V(X_{\alpha}) for every α\alpha so xXαx\in X_{\alpha} for every α\alpha. Then, xαXαx\in\bigcap_{\alpha}X_{\alpha}.

These two statements work for a polynomial in any finite amount of variables, but let us turn our attention back to plane curves. The union of finitely many plane curves is a plane curve, and the intersection of any set of plane curves is a plane curve.

We now return to the Noetherian property. Is the ring of polynomials in nn variables Noetherian? First, we prove the following statement:

Theorem 3.11.

Let RR be a ring. If RR is Noetherian, R[x]R[x] is Noetherian.

Proof.

Since we work with R[x]R[x], any polynomial mentioned in this proof is a polynomial in one dimension.

First, I introduce some definitions: the degree of a polynomial is the smallest possible natural number ii such that aα=0a_{\alpha}=0 for all α>i\alpha>i. In addition, the leading coefficient of a polynomial of degree ii is aia_{i}.

Let II be an ideal of R[x]R[x]. Define JJ to be the set of leading coefficients of all polynomials in II. JJ is an ideal in RR:

  1. (a)

    Closure with respect to addition: let j,kj,k be elements in JJ. They correspond to two polynomials f,gIf,g\in I. Since II is closed under addition, the polynomial f+gIf+g\in I. The polynomial f+gf+g has leading coefficient j+kj+k so j+kJj+k\in J.

  2. (b)

    Existence of additive inverse: 0I0\in I so 0J0\in J.

  3. (c)

    If iJii\in J_{i} and rRr\in R then irIi*r\in I: let iJ,rRi\in J,r\in R. The element ii corresponds to a polynomial fIf\in I. II is an ideal so the polynomial frIf*r\in I. The polynomial frf*r has leading coefficient iri*r so irJi*r\in J.

JJ is finitely generated because RR is Noetherian. Let A={aμ}A=\{a_{\mu}\} be the finite set of leading coefficients that generate JJ. For each aμAa_{\mu}\in A, pick gμIg_{\mu}\in I with leading coefficient aμa_{\mu}. Let G={gμ}G=\{g_{\mu}\}. By the Well-Ordering Principle, GG has a polynomial with largest degree: let kk be this degree.

Then, for every iki\leq k, define JiJ^{i} to be the set of leading coefficients of all polynomials in II with degree smaller than or equal to ii. JiJ^{i} is an ideal in RR (by a similar argument as given above). JiJ^{i} is therefore also finitely generated. Let Ai={aμii}A^{i}=\{a_{\mu^{i}}^{i}\} be the finite set of leading coefficients that generate JiJ^{i}. For each aμiAa_{\mu}^{i}\in A, pick gμiIg_{\mu}^{i}\in I with leading coefficient aμia_{\mu}^{i}. Let Gi={gμi}G^{i}=\{g_{\mu}^{i}\}.

Now, let I0I_{0} be the ideal generated by all gig^{i} in all GiG^{i} and all gg in GG. There are finite number of gig^{i} and gg so I0I_{0} is finitely generated. I claim now that I=I0I=I_{0}.

Let a polynomial fI0f\in I_{0}. ff is generated by some gg’s and gig^{i}’s, which are all elements of II, so fIf\in I.

Now, for sake of contradiction, suppose that II0I\not\subset I_{0}. Find a polynomial ff in II that is not in I0I_{0} with smallest degree. We introduce two cases:

  1. Case 1:

    The degree \ell of ff is bigger than kk. Let aa be the leading coefficient of ff. Then, aJa\in J. This means that a=μAaμrμa=\sum_{\mu}^{A}a_{\mu}*r_{\mu}. Let f0=μArμXdeg(f)deg(gμ)gμf_{0}=\sum_{\mu}^{A}r_{\mu}X^{\text{deg}(f)-\text{deg}(g_{\mu})}g_{\mu}. The polynomial f0f_{0} has the same leading coefficient as ff and the same degree as ff, so deg(ff0)<deg(f)\text{deg}(f-f_{0})<\text{deg}(f). Also, ff0f-f_{0} is not in I0I_{0} because otherwise by the definition of an ideal, ff is in I0I_{0}. Then, we have found an element with smaller degree in II that is not in I0I_{0}—a contradiction.

  2. Case 2:

    The degree \ell of ff is smaller than kk. Let aa be the leading coefficient of ff. Then, aJa\in J_{\ell}. This means that a=μiAaμirμia=\sum_{\mu^{i}}^{A_{\ell}}a_{\mu^{i}}*r_{\mu^{i}}. Let f0=μiArμiXdeg(f)deg(gμi)gμif_{0}=\sum_{\mu^{i}}^{A_{\ell}}r_{\mu^{i}}X^{\text{deg}(f)-\text{deg}(g_{\mu^{i}})}g_{\mu^{i}}. We get a similar contradiction as the one in Case 11.

Note that we must separate out the proof into two cases in this way because we do not want the degree of XX in f0f_{0} to be negative. ∎

Corollary 3.12.

k[X1,,Xn]k[X_{1},\dots,X_{n}] is Noetherian.

Proof.

kk is Noetherian, so k[X1]k[X_{1}] is Noetherian. Now, consider k[X1,X2]k[X_{1},X_{2}]. Define k[X1][X2]k[X_{1}][X_{2}] to be the ring of one variable X2X_{2} with coefficients in k[X1]k[X_{1}]. k[X1]k[X_{1}] is Noetherian so k[X1][X2]k[X_{1}][X_{2}] is Noetherian. The ring k[X1][X2]k[X_{1}][X_{2}] is equivalent (isomorphic) to the ring k[X1,X2]k[X_{1},X_{2}]. By induction, k[X1,,Xn]k[X_{1},\dots,X_{n}] is Noetherian. ∎

Theorem 3.11 is known as Hilbert’s Basis Theorem. What can we use it for?

Proposition 3.13.

Let XX be an algebraic set. XX is the intersection of the varieties of finitely many polynomials.

Proof.

X=V(I(X))X=V(I(X)) by Proposition 3.8. I(X)I(X) is finitely generated by Corollary 3.12. So, we have X=V((f1,,fn))X=V((f_{1},\dots,f_{n})). V((f1,,fn))=V(f1)V(fn)V((f_{1},\dots,f_{n}))=V(f_{1})\cap\dots\cap V(f_{n}) follows from the definition of the variety of a set of polynomials. ∎

This fact is not only interesting when considering algebraic sets (and plane curves), but also useful. In particular, Hilbert used it to prove an important result in invariant theory.

We use Hilbert’s Basis Theorem again. First, we introduce a new definition:

Definition 3.14.

An algebraic set XknX\subset k^{n} is irreducible if there do not exist X1,X2X_{1},X_{2} such that X1X2=XX_{1}\cup X_{2}=X with X1,X2XX_{1},X_{2}\subsetneq X.

We then prove an intermediate result:

Proposition 3.15.

Every descending chain of algebraic sets X1X2X_{1}\supset X_{2}\supset\dots eventually terminates.

Proof.

Suppose that the chain does not terminate. Then, I(X1)I(X2)I(X_{1})\subset I(X_{2})\subset\dots does not terminate—a contradiction because k[X1,,Xn]k[X_{1},\dots,X_{n}] is Noetherian. ∎

Finally, we reach the proposition:

Proposition 3.16.

Let XX be an algebraic set. XX can be written as a finite union of irreducible algebraic sets X1,,XnX_{1},\dots,X_{n} where XiXjX_{i}\not\subset X_{j} for all iji\neq j (let this be property *).

Proof.

Let Σ\Sigma be the set of all algebraic sets in knk^{n} that do not have property *. If Σ\Sigma is empty, our proof is done. Suppose that Σ\Sigma is not empty. Then, there exists an algebraic set XX that does not contain any other algebraic set in Σ\Sigma (otherwise we have a descending chain of algebraic sets that does not terminate). XX is not irreducible by our given. Then, let X1,X2X_{1},X_{2} be algebraic subsets such that X1,X2XX_{1},X_{2}\subsetneq X and X=X1X2X=X_{1}\cup X_{2}. One of X1,X2X_{1},X_{2} cannot have property *; otherwise XX has property * (by unioning all finite irreducible polynomials making up X1X_{1} and X2X_{2}) so then XΣX\not\in\Sigma. Thus, at least one of X1,X2X_{1},X_{2} is in Σ\Sigma, so there does exist an algebraic set in Σ\Sigma that is contained in XX—a contradiction. ∎

Thus, an algebraic set can be written as the intersection of finitely many varieties of polynomials. In addition, an algebraic set can be written as the union of finitely many irreducible algebraic sets.

In Section 22, we did not know of any rings that were not Noetherian. We now give the following example:

Proposition 3.17.

Let kk be a field. k[X1,X2,]k[X_{1},X_{2},\dots] is not Noetherian.

Proof.

The chain of ideals (X1)(X1,X2)(X_{1})\subset(X_{1},X_{2})\subset\dots never terminates. ∎

3.5. Correspondences between points and polynomials

We introduce a bijective correspondonce from irreducible algebraic sets to prime ideals:

Proposition 3.18.

Let XknX\subset k^{n} be an algebraic set. XX is irreducible \iff I(X)I(X) is prime.

Proof.

Let us prove that if XX is irreducible, I(X)I(X) is prime. Suppose that I(X)I(X) is not prime. Then, there exist polynomials f1,f2I(X)f_{1},f_{2}\not\in I(X) such that f1f2I(X)f_{1}f_{2}\in I(X). Let I1I_{1} be the ideal generated by ff and all generators of II. I(X)I1I(X)\subsetneq I_{1} so V(I1)V(I(X))=XV(I_{1})\subsetneq V(I(X))=X because XX is an algebraic set. The same logic applies for f2f_{2}. So we have that V(I1),V(I2)XV(I_{1}),V(I_{2})\subsetneq X. Let xXx\in X. We know that f1f2(x)=0f_{1}f_{2}(x)=0, so either f1(x)=0f_{1}(x)=0 or f2(x)=0f_{2}(x)=0, so xV(I1)x\in V(I_{1}) or xV(I2)x\in V(I_{2}). Then, XV(I1)V(I2)X\subset V(I_{1})\cup V(I_{2}) so XX is reducible.

Let us prove now that if I(X)I(X) is prime, XX is irreducible. Suppose that XX is not irreducible. Let X1,X2XX_{1},X_{2}\subsetneq X such that X=X1X2X=X_{1}\cup X_{2}. Because X1XX_{1}\subsetneq X, I(X)I(X1)I(X)\subsetneq I(X_{1}), there exists f1f_{1} such that f1I(X1)f_{1}\in I(X_{1}) but f1I(X)f_{1}\not\in I(X) (and a similar f2f_{2} for I(X2)I(X_{2})). f1f2f_{1}f_{2} vanishes at all points in XX so f1f2I(X)f_{1}f_{2}\in I(X), so I(X)I(X) is not prime. ∎

Now we return to Proposition 3.8 (4)(4) to see if we can find conditions such that S=I(V(S))S=I(V(S)). If we succeed, we can define a bijective correspondence between these sets SS and algebraic sets V(S)V(S). Why are we not guaranteed equality?

Instead of considering all sets of polynomials SS, we consider ideals of the polynomial ring JJ only in this problem. Recall that if a set of polynomials SS and an ideal JJ both vanish on XX (i.e. V(S)=V(J)V(S)=V(J)), SJS\subset J. As we are trying to achieve S=I(V(S))S=I(V(S)), using SJS\neq J would mean that SI(V(S))S\subset I(V(S))—not what we want.

First, consider the polynomial x2+1k[x]x^{2}+1\in k[x], where k=k=\mathbb{R}. Let JJ be the ideal generated by this polynomial. V(J)=V(J)=\emptyset, so I(V(J))=knI(V(J))=k^{n}. As we can see, the solution just doesn’t exist in \mathbb{R}. If we were to use k=k=\mathbb{C}, though, V(J)V(J) would not be empty (actually |V(J)|=2|V(J)|=2). So, in order for equality to hold, we need kk to have some special property.

Second, consider the polynomial x2k[x]x^{2}\in k[x], where k=k=\mathbb{R}. Let JJ be the ideal generated by this polynomial. V(J)={0}V(J)=\{0\}, and I(V(I))=(x)I(V(I))=(x). As we can see, the original ideal did not contain all polynomials that vanish on its vanishing set, so we ended up with inequality. Then, if we want J=I(V(J))J=I(V(J)), for every fnf^{n} in JJ, ff must be in JJ.

We define two new definitions based on these two observations:

Definition 3.19.

A field kk is algebraically closed if every polynomial in k[x]k[x] has a root.

Definition 3.20.

Let JJ be an ideal. The radical of JJ, denoted by J\sqrt{J}, is the set {ffnJ for some n>0}\{f\mid f^{n}\in J\text{ for some }n>0\}. J\sqrt{J} is an ideal.

Hilbert’s Nullstellensatz proves the bijective correspondence between radical ideals and algebraic sets:

Theorem 3.21.

Let kk be an algebraically closed field and JJ be an ideal with coefficients in kk. Then, J=I(V(J))\sqrt{J}=I(V(J)).

Proof of Hilbert’s Nullstellensatz requires more definitions; we do not prove it here. We hope though that the reader has noted how useful such statements are to understanding the polynomial ring and these sets of points.

Acknowledgements

I am very grateful to my mentor Michael Neaton and to Dr. May for their useful comments and suggestions. All errors are mine.

References

  • [1] R. Fulton. Algebraic Curves. Addision-Wesley. 1969.
  • [2] I. R. Shafarevich. Basic Algebraic Geometry 1. Springer. 1974.
  • [3] K. E. Smith, L. Kahanpaa, P. Kekalainen, W. Traves. An Invitation to Algebraic Geometry. Springer. 2000.
  • [4] M. F. Atiyah, I. G. MacDonald. Introduction to Commutative Algebra. Addison-Wesley. 1969.
  • [5] M. Reid. Undergraduate Algebraic Geometry. Cambridge University Press. 1989.
  • [6] R. Hartshorne. Algebraic Geometry. Springer. 1977.