A bottom-up approach to Hilbert’s Basis Theorem
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 , 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 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 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 in two variables (or any number of variables actually), the set of points such that for all can be written as the intersection of finitely many sets of points , where such that for some polynomial . 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, . 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 . 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 is a set with two binary operations addition and multiplication, denoted by and respectively. These operations have the following properties:
-
(1)
Addition is associative and commutative; there exists an additive identity (denoted by ); each element in the ring has a (unique) additive inverse.
-
(2)
Multiplication is associative; there exists a multiplicative identity (denoted by ).
-
(3)
Multiplication is distributive over addition. That is, and hold for .
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, can be equal to . In this case, it is easy to see that the ring has only one element, (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 where 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 (a certain subset of this ring). This set has some interesting properties.
First, if you add any two elements divisible by , you end up with an integer divisible by . Second, if you multiply an element in the set with any other integer, you end up with an integer divisible by . Ideals are defined by these two properties.
Definition 2.3.
An ideal is a subset of a ring that, inheriting the addition and multiplication operations from , satisfies the following properties:
-
(1)
Closure with respect to addition (if then ); associativity with respect to addition; additive identity is in the ideal (); each element’s additive inverse is also in the ideal (if , ).
-
(2)
If and , .
-
(3)
If and , .
Remark 2.4.
Since we work with commutative rings, it is only necessary to prove either or . In non-commutative rings, an ideal that satisfies but not necessarily is a left-ideal, and an ideal that satisfies but not necessarily 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 (or as discussed) is sufficient.
-
(a)
Associativity is inherited from the ring definition.
-
(b)
For , .
An ideal cannot be empty because it must contain .
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 and are ideals.
But are these the only ideals? Consider first the integers.
Proposition 2.5.
For all , define . is an ideal. These are the only ideals of .
Proof.
The proof that is an ideal follows from my previous remark. So then suppose there exists an ideal that cannot be written as an . because then , so there exists at least one nonzero integer in . Consider the set of nonzero positive integers in . This set must be nonempty; an empty set would imply that the nonzero integer we found earlier is negative, but —a contradiction. By the Well-Ordering Principle, find the smallest positive nonzero integer .
Because , there must exist an element that is not in the form for any . Suppose it is positive (if it is negative multiply it by and it will still be in the ideal). We know that because is the smallest positive integer. Now, observe that is still in the ideal and also will not be in the form . Repeat until the result is less than . The result will be in , positive and nonzero (because it is not in the form ). Thus, we have found a smaller positive nonzero integer in that is smaller than —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 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 are the field itself and the zero ideal. Let be an arbitrary ideal.
-
Case 1:
contains no non-zero elements. Then, .
-
Case 2:
contains a non-zero element. Let be this element. Then, because is a field, there exists a multiplicative inverse of , denoted by . By the property of ideals, . Any element is in the ideal now because , so then .
Second, we prove that if the only ideals of a ring are the ring and the zero ideal, the ring is a field. Let be a non-zero element in our commutative ring . Construct a set . It is easy to check that this is an ideal. Because this ideal contains a nonzero element , it cannot be the zero ideal, so by our given, the ideal must be the entire ring. Then, there exists an element such that since is in our ideal. Thus, has an inverse. ∎
From this proof, we can also see the following statement:
Corollary 2.7.
Let be an ideal of . Then, .
2.3. Another definition of an ideal
Consider the following definition:
Definition 2.8.
A ring homomorphism is a map from one ring to another ring with the following properties:
-
(1)
,
-
(2)
, and
-
(3)
, where are the multiplicative identities of and respectively.
We introduce an equivalent definition of the ideal using ring homomorphisms:
Definition 2.9.
An ideal is a subset of a ring that is the kernel of a ring homomorphism. That is, a set is an ideal if there exists a ring homomorphism , where is some other ring, such that .
Proposition 2.10.
Proof.
First, we show Definition 2.9 from Definition 2.3. We need to find a ring and a function as described. We need such that for all , . Furthermore, we need such that adding to any will not change the computed result (i.e., ).
Let be an element in . Define . Define .
We now define an equivalence relation between elements in . We say that for if and only if . It follows from properties of an ideal that is an equivalence relation. Notice that, under this definition, .
Define addition and multiplication as follows: for , let and . Both of these are well-defined. Consider, for example, addition. Let be elements in with and . Let be elements in such that and . Then,
We conclude that equivalent inputs implies equivalent outputs under our addition operation. Multiplication follows similarly as well.
Under these operations, forms a ring. To finish the proof, define and as defined above. By our definition of , we already know that and . It suffices to show that where is the multiplicative identity of . The proof follows from the definition.
Consider the ideals of we have found. What is the ring and ring homomorphism that we should be able to find, according to our new definition of an ideal?
-
(1)
: we use ( is the set of integers modulo —a ring) and where . The set of elements that map to is .
-
(2)
: we use and where . The only element that maps to is .
-
(3)
: we use and where . All elements map to .
2.4. What is in an ideal?
Consider an ideal of ring . 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 be an ideal and let be a subset of indexed by . Consider the expression where is an element of the ideal and is an element in . Can we write any element in in this form for some ’s? If so, we denote the ideal by
and say that it is generated by . For example, (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 , we can keep multiplying and finding larger and larger elements.
In addition, note that just because we can write an element as a sum of ’s doesn’t mean that for any two elements such that , either or in . An ideal in which this is true for every is called a prime ideal:
Definition 2.11.
A prime ideal of ring is an ideal that satisfies the following properties:
-
(1)
If and , then either or .
-
(2)
is not equal to the ring .
Consider the ideal of integers divisible by in the ring of integers: this ideal is a prime ideal because we can write every element as multiplied by an integer, and is in our ideal. However, the ideal of integers divisible by does not have this property; for example but neither nor are divisible by . The generalization of these statements corresponds to the following statement:
Proposition 2.12.
In , ideals generated by a prime number are prime. The zero ideal is also prime. All other ideals are not prime.
Proof.
Let be an ideal generated by a prime natural number . Suppose is not prime. Then, there exists an element and such that but and . This means that is divisible by but neither nor are divisible by , which is impossible.
The zero ideal is prime because no integer divides .
We already showed that all ideals of are generated by one element, and so it is only necessary to look at ideals generated by nonprime elements. Let be an ideal generated by a nonprime natural number . Because is not prime, for some not divisible by , so and are not in ; then, is not prime. The ideal generated by is the same as the one generated by 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 of a ring is the kernel of a ring homomorphism , where is a ring that is not equal to and has no zero-divisors (no elements such that for some , ).
Proposition 2.14.
Proof.
Let 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 and such that is the kernel of . Then, take and that Definition 2.13 refers to. Consider two cases:
-
Case 1:
. is not equal to . Then, there exists an element in that is not in , and .
-
Case 2:
has a zero-divisor. Let be a zero-divisor that is not equal to with a corresponding that is not equal to such that is equal to . By construction, is surjective, so there exist such that and . Then,
so . However, —otherwise (and similar argument for ). Then, is not prime according to Definition 2.11, so we have arrived at a contradiction.
For sake of contradiction, let be an ideal that is not prime by Definition 2.11. By Definition 2.9, there exists a set and such that is the kernel of . Consider two cases:
-
Case 1:
. Then, every element maps to , so .
-
Case 2:
There exist elements such that is in but neither nor is in . Then,
but neither nor are equal to because neither nor are in . Then, both and are zero-divisors of 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 that generate ?
Note that such a set must exist. If we choose then is generated by . Let be an element in . Set for and for .
But can we find other sets? Specifically, can we find a finite set that generates ? Such an ideal is said to be finitely generated. Furthermore, for a ring , is every ideal finitely generated? Such a ring is said to be a Noetherian ring.
Consider the ring of integers. Since every ideal of consists of all multiples of some element, every ideal is finitely generated and 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:
-
i
For every ideal , there exist such that .
-
ii
Every chain of ideals eventually terminates such that .
-
iii
Every nonempty set of ideals of has an ideal that is not contained in any other.
Proof.
(i) (ii): Take a chain of ideals, and let be the union of them all. For , there exist that generate . All of these 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 , we can find an for some that finally contains all of them. It is clear that because they are generated by the same elements. To prove the chain terminates, let be a natural number bigger than . is a subset of and therefore is a subset of . is a subset of because we have a chain of ideals. Then, , so the chain terminates.
(ii) (iii): At the ideal at which the chain terminates, that ideal is not contained in any other.
(iii) (i): Let be an ideal and . By our given, has an element that is not contained in any other. Suppose that . Then, there exists . Let be the ideal generated by all the finitely many generators in and also . is still finitely generated and contains , contradicting our assumption that is not contained in any other. Then, 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 th 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 be a field and be a polynomial. A plane curve is defined as the set of all points that satisfy (such a function is said to “vanish” on ).

We are usually familiar with finding the zeros of a polynomial by setting , and we know that there are finite number of such (over a field). Now, we have , and there can be an infinite number of that satisfy (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 be a field, be a natural number, and be . The polynomial ring in variables is the set of all expressions of the form where for all and where is the th component in . A polynomial in variables is an element of the polynomial ring. Polynomials are finite sums so we impose the additional restriction that for all but finitely many . Let and be two polynomials. Addition of two polynomials is defined as . Multiplication is defined as where . 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 variables. Now, we generalize the “set of points” aspect of plane curves.
Definition 3.3.
Let be a function where . is known as the variety of . is an algebraic set if for some .
Remark 3.4.
Note that we are evaluating on a point . This is done using the field operations of .
Under this definition, a plane curve is just an algebraic set created from only one () two-dimensional (variables ) polynomial.
We introduce the following proposition that follows from the definition:
Proposition 3.5.
and 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 be a subset of (not necessarily an algebraic set). The ideal of , denoted by , is the set of polynomials in variables that vanish on .
Proposition 3.7.
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 and satisfy the following properties for any field :
-
(1)
Let be subsets of . If ,
-
(2)
Let be subsets of . If , .
-
(3)
Let be a subset of . Then, . if and only if is an algebraic set.
-
(4)
Let be a subset of . Then, .
Proof.
and follow from the definition.
For , note that is defined as the set of all polynomials that vanish on , so these polynomials will vanish on . For the second part of , we know that ; then, for , meaning that is an algebraic set. Next, let us prove that is an algebraic set implies that . for some . Suppose, for sake of contradiction, that does not contain . Then, there exists a polynomial . We know that for all by definition, so then —a contradiction. Because , so (we already know because of the first part of ).
For , let be such a set and let . By definition, we know that vanishes at the points at which vanishes. ∎
We figured out under what conditions for of the proposition. But we did not figure out under what conditions . 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 . Then,
-
(1)
and
-
(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)
The union of finitely many algebraic sets is an algebraic set.
-
(2)
The intersection of any collection of algebraic sets is an algebraic set.
Proof.
-
(1)
Let and be algebraic sets for some sets of polynomials and . Indeed, , where denotes the set of all products of an element of by an element of . Let be an element in . Then, or . Without loss of generality, suppose . Then, for all so every function in vanishes on . Now, let . Without loss of generality, suppose that . Then, there is an such that . All polynomials , where , vanish on , so then for all and vanishes on .
By induction, the union of finitely many algebraic sets is an algebraic set as well.
-
(2)
Let be a collection of algebraic sets indexed by . Then, . Let . Then, for all so for all . Now, let . By definition, for every so for every . Then, .
∎
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 variables Noetherian? First, we prove the following statement:
Theorem 3.11.
Let be a ring. If is Noetherian, is Noetherian.
Proof.
Since we work with , 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 such that for all . In addition, the leading coefficient of a polynomial of degree is .
Let be an ideal of . Define to be the set of leading coefficients of all polynomials in . is an ideal in :
-
(a)
Closure with respect to addition: let be elements in . They correspond to two polynomials . Since is closed under addition, the polynomial . The polynomial has leading coefficient so .
-
(b)
Existence of additive inverse: so .
-
(c)
If and then : let . The element corresponds to a polynomial . is an ideal so the polynomial . The polynomial has leading coefficient so .
is finitely generated because is Noetherian. Let be the finite set of leading coefficients that generate . For each , pick with leading coefficient . Let . By the Well-Ordering Principle, has a polynomial with largest degree: let be this degree.
Then, for every , define to be the set of leading coefficients of all polynomials in with degree smaller than or equal to . is an ideal in (by a similar argument as given above). is therefore also finitely generated. Let be the finite set of leading coefficients that generate . For each , pick with leading coefficient . Let .
Now, let be the ideal generated by all in all and all in . There are finite number of and so is finitely generated. I claim now that .
Let a polynomial . is generated by some ’s and ’s, which are all elements of , so .
Now, for sake of contradiction, suppose that . Find a polynomial in that is not in with smallest degree. We introduce two cases:
-
Case 1:
The degree of is bigger than . Let be the leading coefficient of . Then, . This means that . Let . The polynomial has the same leading coefficient as and the same degree as , so . Also, is not in because otherwise by the definition of an ideal, is in . Then, we have found an element with smaller degree in that is not in —a contradiction.
-
Case 2:
The degree of is smaller than . Let be the leading coefficient of . Then, . This means that . Let . We get a similar contradiction as the one in Case .
Note that we must separate out the proof into two cases in this way because we do not want the degree of in to be negative. ∎
Corollary 3.12.
is Noetherian.
Proof.
is Noetherian, so is Noetherian. Now, consider . Define to be the ring of one variable with coefficients in . is Noetherian so is Noetherian. The ring is equivalent (isomorphic) to the ring . By induction, is Noetherian. ∎
Theorem 3.11 is known as Hilbert’s Basis Theorem. What can we use it for?
Proposition 3.13.
Let be an algebraic set. is the intersection of the varieties of finitely many 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 is irreducible if there do not exist such that with .
We then prove an intermediate result:
Proposition 3.15.
Every descending chain of algebraic sets eventually terminates.
Proof.
Suppose that the chain does not terminate. Then, does not terminate—a contradiction because is Noetherian. ∎
Finally, we reach the proposition:
Proposition 3.16.
Let be an algebraic set. can be written as a finite union of irreducible algebraic sets where for all (let this be property ).
Proof.
Let be the set of all algebraic sets in that do not have property . If is empty, our proof is done. Suppose that is not empty. Then, there exists an algebraic set that does not contain any other algebraic set in (otherwise we have a descending chain of algebraic sets that does not terminate). is not irreducible by our given. Then, let be algebraic subsets such that and . One of cannot have property ; otherwise has property (by unioning all finite irreducible polynomials making up and ) so then . Thus, at least one of is in , so there does exist an algebraic set in that is contained in —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 , we did not know of any rings that were not Noetherian. We now give the following example:
Proposition 3.17.
Let be a field. is not Noetherian.
Proof.
The chain of ideals 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 be an algebraic set. is irreducible is prime.
Proof.
Let us prove that if is irreducible, is prime. Suppose that is not prime. Then, there exist polynomials such that . Let be the ideal generated by and all generators of . so because is an algebraic set. The same logic applies for . So we have that . Let . We know that , so either or , so or . Then, so is reducible.
Let us prove now that if is prime, is irreducible. Suppose that is not irreducible. Let such that . Because , , there exists such that but (and a similar for ). vanishes at all points in so , so is not prime. ∎
Now we return to Proposition 3.8 to see if we can find conditions such that . If we succeed, we can define a bijective correspondence between these sets and algebraic sets . Why are we not guaranteed equality?
Instead of considering all sets of polynomials , we consider ideals of the polynomial ring only in this problem. Recall that if a set of polynomials and an ideal both vanish on (i.e. ), . As we are trying to achieve , using would mean that —not what we want.
First, consider the polynomial , where . Let be the ideal generated by this polynomial. , so . As we can see, the solution just doesn’t exist in . If we were to use , though, would not be empty (actually ). So, in order for equality to hold, we need to have some special property.
Second, consider the polynomial , where . Let be the ideal generated by this polynomial. , and . 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 , for every in , must be in .
We define two new definitions based on these two observations:
Definition 3.19.
A field is algebraically closed if every polynomial in has a root.
Definition 3.20.
Let be an ideal. The radical of , denoted by , is the set . is an ideal.
Hilbert’s Nullstellensatz proves the bijective correspondence between radical ideals and algebraic sets:
Theorem 3.21.
Let be an algebraically closed field and be an ideal with coefficients in . Then, .
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.