<<

in Integral I

Throughout these notes, R denotes an integral unless other- wise noted.

1 Unique factorization domains and principal domains

Definition 1.1. For r, s ∈ R, we say that r divides s (written r|s) if there exists a t ∈ R such that s = tr. An element u ∈ R is a unit if it has a multiplicative inverse, i.e. if there exists an element v ∈ R such that uv = 1. The (multiplicative) of units is denoted R∗. If r, s ∈ R, then r and s are associates if there exists a unit u ∈ R∗ such that r = us. In this case, s = u−1r, and indeed the relation that r and s are associates is an equivalence relation (HW). Another easy result is (this was on the midterm): Lemma 1.2. For an R, if r, s ∈ R, then r and s are associates ⇐⇒ (r) = (s). Proof. =⇒ : If r = us, then r ∈ (s), hence (r) ⊆ (s). Since s = u−1r, s ∈ (r), and hence (s) ⊆ (r). Thus (r) = (s). ⇐= : If r = 0, then (r) = (0) = {0}, so if (s) = (r), then s ∈ {0}, hence s = 0. Thus r = s = 0, and in particular r and s are associates, since for example s = 1 · r. If r 6= 0, then (s) = (r) =⇒ s ∈ (r) =⇒ s = tr for some t ∈ R. Likewise r ∈ (s), so r = us for some u ∈ R. Thus

r = us = utr.

By assumption, r 6= 0, and since R is an integral domain, we can cancel r. Thus 1 = ut. So u is a unit and r = us, so that r and s are associates.

∗ Example 1.3. 1) R = Z. The units Z = ±1. Two n and m are associates ⇐⇒ m = ±n. 2) R = F [x], F a field. The units in F [x] are: (F [x])∗ = F ∗, the set of constant nonzero . Hence, if F is infinite, there are an infinite

1 number of units. Two polynomials f and g are associates ⇐⇒ there exists a c ∈ F ∗ with g = cf. ∗ 3) R = Z[i], the Gaussian integers. The units (Z[i]) = {±1, ±i}. Two elements α, β √∈ Z[i] are associates ⇐⇒ α = ±β or α = ±iβ√. 4) R = Z[ 2]. As we√ have seen on the homework, 1 + 2 is a unit of ∗ ∼ infinite order.√ In fact, (Z[ 2]) = Z × (Z/2Z). √ ∗ 5) R = Z[ −2]. An easy calculation shows that (Z[ −2]) = {±1}. We say that r ∈ R is irreducible if r 6= 0, r is not a unit, and, for all s ∈ R, if s divides r then either s is a unit or s is an associate of r. In other words, if r = st for some t ∈ R, then one of s or t is a unit (and hence the other is an associate of r). If r ∈ R with r 6= 0 and r is not a unit, then r is reducible if it is not irreducible. Definition 1.4. Let r, s ∈ R, not both 0. A (gcd) of r and s is an element d ∈ R such that d|r, d|s, and if e ∈ R and e|r, e|s, then e|d. Lemma 1.5. Let r, s ∈ R, not both 0. If a gcd of r and s exists, it is unique up to a unit (i.e. any two gcd’s of r and s are associates). Proof. If d and d0 are both a gcd of r and s, then by definition d|d0 and d0|d. Hence there exist u, v ∈ R such that d0 = ud and d = vd0. Thus d = uvd. Since d|r and d|s and at least one of r, s is nonzero, d 6= 0. Hence uv = 1, so that u and v are units. Thus d and d0 are associates.

Definition 1.6. Let r, s ∈ R, not both 0. The elements r and s are relatively prime if 1 is a gcd of r and s; equivalently, if d ∈ R and d|r, d|s, then d is a unit. Lemma 1.7. Let r, s ∈ R, with r irreducible.Then a gcd of r and s exists, and is either a unit or an associate of r. Hence, if r does not divide s, then r and s are relatively prime. Proof. Suppose that r divides s. Then r is a gcd of r and s, since r divides both r and s and any e which divides r and s is either a unit or an associate of r, hence divides r. If r does not divide s, then any e which divides r and s cannot be an associate of r, hence must be a unit. It follows that 1 is a gcd of r and s, and hence r and s are relatively prime.

Definition 1.8. R is a unique factorization domain (UFD) if

(i) for every r ∈ R not 0 or a unit, there exist irreducibles p1, . . . , pn ∈ R such that r = p1 ··· pn, and

2 (ii) if pi, 1 ≤ i ≤ n and qj, 1 ≤ j ≤ m are irreducibles such that p1 ··· pn = q1 ··· qm, then n = m and, after reordering, pi and qi are associates. Note that two separate issues are involved: (i) the existence of some factorization of r into irreducibles and (ii) the uniqueness of a factorization. As we shall see, these two questions are in general unrelated, in the sense that one can hold but not the other. Given an element r in a UFD, not 0 or a unit, it is often more natural to factor r by grouping together all of the associated irreducibles (after making some choices). Hence, such an r can always be written as

a1 an r = up1 ··· pn , where u is a unit, the pi are irreducibles, ai > 0, and, for i 6= j, pi and pj are not associates, and such a product is essentially unique in the following sense: if also b1 bm r = vq1 ··· qm , where v is a unit, the qj are irreducibles, bj > 0, and, for k 6= `, qk and q` are not associates, then n = m and, after reordering, pi and qi are associates and ai = bi. Definition 1.9. R is a domain (PID) if every ideal I of R is principal, i.e. for every ideal I of R, there exists r ∈ R such that I = (r).

Example 1.10. The rings Z and F [x], where F is a field, are PID’s. We shall prove later: A is a unique factorization domain. However, there are many examples of UFD’s which are not PID’s. For example, if n ≥ 2, then the F [x1, . . . , xn] is a UFD but not a PID. Likewise, Z[x] is a UFD but not a PID, as is Z[x1, . . . , xn] for all n ≥ 1.

Proposition 1.11. If R is a UFD, then the gcd of two elements r, s ∈ R, not both 0, exists.

Proof. If say r = 0, then the gcd of r and s exists and is s. If r is a unit, then the gcd of r and s exists and is a unit. So we may clearly assume that r is neither 0 nor a unit, and likewise that s is neither 0 nor a unit. Then we can factor both r and s as in the comments after the definition of a UFD. In fact, it is clear that we can write

a1 ak b1 bk r = up1 ··· pk , s = vp1 ··· pk

3 where u and v are units, the pi are irreducibles, ai, bi ≥ 0, and, for i 6= j, pi and pj are not associates. (Here, we set ai = 0 if pi is not a factor of r, and similarly for bi.) Then set c1 ck d = p1 ··· pk , where ci = min{ai, bi}. We claim that d is a gcd of r and s. Clearly d|r and d|s. If now e|r and e|s and q is an irreducible factor of e, then q is an associate of pi for some i, and so we can assume that q = pi. If mi is mi mi mi the largest such that pi |e, then since pi |r and pi |s, mi ≤ ai and mi ≤ bi. Hence mi ≤ ci. It then follows by taking the factorization of e into powers of the pi times a unit that e|d. Hence d is a gcd of r and s. Lemma 1.12. If R is a UFD and p, r, s ∈ R are such that p is an irreducible and p|rs, then either p|r or p|s. More generally, if t and r are relatively prime and t|rs then t|s. Proof. To see the first statement, write rs = pt and factor r, s, t into irre- ducibles. Then p must be an associate of some irreducible factor of either r or s, hence p divides either r or s. The second statement can be proved along similar but slightly more complicated lines.

As a consequence, we have: Proposition 1.13. Let R be a UFD and let r ∈ R, where r 6= 0. Then (r) is a ⇐⇒ r is irreducible. Proof. =⇒ : If (r) is a prime ideal, then r is not a unit, and r 6= 0 by assumption. If r = st, then one of s, t ∈ (r), say s ∈ (r), hence s = ru. Then r = rut so that ut = 1 and t is a unit. Hence r is irreducible. (Note: this part did not use the fact that R was a UFD, and holds in every integral domain.) ⇐= : If r is irreducible, then it is not a unit and hence (r) 6= R. Suppose that st ∈ (r). Then r|st. By the remark above, either r|s or r|t, i.e. either s ∈ (r) or t ∈ (r). Hence (r) is prime.

Remark 1.14. (i) In case R is not a UFD, there will in general exist irre- ducibles r such that (r) is not a prime ideal. (ii) In general, suppose that R is an integral domain and that r ∈ R, r 6= 0. If (r) is a prime ideal, then for all s, t ∈ R, if r divides st, then either r divides s or r divides t. We turn now to the study of a PID, with a view toward showing even- tually that a PID is a UFD.

4 Theorem 1.15. Let R be a PID, and let r, s ∈ R, not both 0. Then a gcd d of r and s exists. Moreover, d is a linear combination of r and s: there exist a, b ∈ R such that d = ar + bs.

Note: For a general UFD, the gcd of two elements r and s will not in general be a linear combination of r and s. For example, in F [x, y], the elements x and y are relatively prime, hence their gcd is 1, but 1 is not a linear combination of x and y, since if f = xa + yb is any linear combination of x and y, where a, b ∈ F [x, y], then ev0,0(f) = f(0, 0) = 0, but ev0,0(1) = 1. Proof. This argument is very similar to the corresponding argument for F [x], or for Z. Given r, s ∈ R, not both 0, consider the ideal

(r, s) = {ar + bs : a, b ∈ R} = (r) + (s).

Then (r, s) is easily checked to be an ideal, hence there exists a d ∈ R with (r, s) = (d). By construction d = ar + bs for some r, s ∈ R. Since r = 1 · r + 0 · s ∈ (r, s) = (d), this says that d|r. Similarly d|s. Finally, if e|r and e|s, then e|(ar + bs) = d.

Corollary 1.16. If R is a PID, r, s ∈ R are relatively prime and r|st, then r|t.

Proof. Write 1 = ar+bs for some a, b ∈ R. Then t = tar+tbs = r(at)+b(st). By assumption r|st and clearly r|r(at). Hence r|t.

Corollary 1.17. If R is a PID, and r ∈ R is an irreducible, then for all s, t ∈ R, if r|st, then either r|s or r|t.

Proof. By Lemma 1.7, if r does not divide s, then r and s are relatively prime. Then, by the previous corollary, r|t. Hence either r|s or r|t.

The following proves the uniqueness half of the assertion that a PID is a UFD:

Corollary 1.18. If R is a PID, then uniqueness of factorization holds in R: if pi, 1 ≤ i ≤ n and qj, 1 ≤ j ≤ m are irreducibles such that p1 ··· pn = q1 ··· qm, then n = m and, after reordering, pi and qj are associates. Proof. This is proved in exactly the same way as the argument we gave for F [x] (or, in Modern Algebra I, for Z). Theorem 1.19. A PID is a UFD.

5 Proof. We have already seen that, if an irreducible factorization exists, it is unique up to order and associates. Thus the remaining point is to show that, if R is a PID, then every element r ∈ R, not 0 or a unit, admits some factorization into a product of irreducibles. The proof will be in several steps. Lemma 1.20. Let R be an integral domain with the property that, if

(r1) ⊆ (r2) ⊆ · · · ⊆ (rn) ⊆ (rn+1) ⊆ · · · is an increasing sequence of principal ideals, then the sequence is eventually constant, i.e. there exists an N such that, for all n ≥ N, (rn) = (rN ). Then every nonzero r ∈ R which is not a unit factors into a product of irreducibles. We can paraphrase the hypothesis of the lemma by saying that R satisfies the ascending chain condition (a.c.c) on principal ideals.

Proof of the lemma. Suppose by contradiction that r ∈ R is an element, not zero or a unit, which does not factor into a product of irreducibles. In particular, r itself is not irreducible, so that r = r1s1 where neither r1 nor s1 is a unit. Thus (r) is properly contained in (r1) and in (s1), by Lemma 1.2. Clearly, we can assume that at least one of r1, s1, say r1, does not factor into irreducibles (if both so factor, so does the product). By applying the above to r1, we see that (r1) is strictly contained in a principal ideal (r2), where r2 does not factor into a product of irreducibles. Continuing in this way, we can produce a strictly increasing infinite chain of principal ideals (r1) ⊂ (r2) ⊂ · · · , i.e. each (ri+1) properly contains the previous ideal (ri), contradicting the hypothesis on R.

To complete the proof of the theorem that a PID is a UFD, it suffices to show that a PID R satisfies the hypotheses of the above lemma. First suppose that (r1) ⊆ (r2) ⊆ · · · is an increasing sequence of ideals of R. It S is easy to check that I = i(ri) is again an ideal. More generally, we have the following: Claim 1.21. Let R be a ring, not necessarily an integral domain, and let S I1 ⊆ I2 ⊆ · · · be an increasing sequence of ideals of R. If I = n In, then I is an ideal of R.

Proof. It suffices to show that I is nonempty, closed under addition and has the absorbing property. Since In ⊆ I for all n and In 6= ∅, I 6= ∅. Given a, b ∈ I, there exists a j such that a ∈ Ij and there exists a k such that b ∈ Ik. Setting ` = max{j, k}, we have a ∈ Ij ⊆ I` and b ∈ Ik ⊆ I`. Hence

6 a, b ∈ I`, and since I` is an ideal, a + b ∈ I` ⊆ I. Thus I is closed under addition. Similarly, if a ∈ I, then a ∈ Ij for some j. Hence, for all r ∈ R, ra ∈ Ij ⊆ I. Thus I is an ideal. Returning to the proof of the theorem, given the increasing sequence of S ideals (r1) ⊆ (r2) ⊆ · · · , the claim implies that I = i(ri) is again an ideal of R. Since R is a PID, I = (r) for some r ∈ R. Necessarily r ∈ (rN ) for some S N. But then (r) ⊆ (rN ) ⊆ (rN+1) · · · ⊆ i(ri) = (r). Thus all inclusions are equalities, and (rn) = (rN ) for all n ≥ N, i.e. the sequence is eventually constant. Hence R satisfies the hypotheses of the previous lemma, so that every r ∈ R, not 0 or a unit, factors into a product of irreducibles.

The ascending chain condition and the arguments we have just given are so fundamental that we generalize them as follows:

Proposition 1.22. For a ring R, not necessarily an integral domain, the following two conditions are equivalent:

(i) Every ideal I of R is finitely generated: if I is an ideal of R, then I = (r1, . . . , rn) for some ri ∈ R. (ii) Every increasing sequence of ideals is eventually constant, in other words if I1 ⊆ I2 ⊆ · · · ⊆ In ⊆ In+1 ⊆ · · · ,

where the In are ideals of R, then there exists an N ∈ N such that for all k ≥ N, Ik = IN . If the ring R satisfies either of the equivalent conditions above, then R is called a .

Proof. (i) =⇒ (ii): given an increasing sequence of ideals I1 ⊆ I2 ⊆ · · · , let S I = n In. Then by the claim above, I is an ideal, and hence I = (r1, . . . , rn) for some ri ∈ R. Thus ri ∈ Ini for some ni. If N = maxi ni, then ri ∈ IN for every i. Hence, for all k ≥ N, I = (r1, . . . , rn) ⊆ IN ⊆ Ik ⊆ I. It follows that Ik = IN = I for all k ≥ N. (ii) =⇒ (i): Let I be an ideal of R and choose an arbitrary r1 ∈ I (for example, r1 could be 0). Set I1 = (r1). If I = I1, stop. Otherwise there exists an r2 ∈ I − I1. Set I2 = (r1, r2), and note that I2 strictly contains I1. If I = I2, stop, otherwise there exists an r3 ∈ I − I2. Inductively suppose that we have found Ik = (r1, . . . , rk) with Ik ⊆ I. If I = Ik we are done, otherwise there exists rk+1 ∈ I − Ik and we set Ik+1 = (r1, . . . , rk+1). So if I is not finitely generated, we have constructed a strictly increasing

7 sequence I1 ⊂ I2 ⊂ · · · , contradicting the assumption on R. Thus I is finitely generated.

Example 1.23. Clearly, a PID satisfies (i) and hence a PID is Noetherian. But there are many examples of Noetherian rings which are not PIDs. For example, the Hilbert basis theorem states that, if R is Noetherian, then so is R[x1, . . . , xn]. Thus Z[x1, . . . , xn] and F [x1, . . . , xn] are Noetherian (where F is a field). It is also easy to see that R Noetherian =⇒ R/I is Noetherian for every ideal I ⊆ R. Thus rings of the form Z[x1, . . . , xn]/I and F [x1, . . . , xn]/I, which are the important rings arising in algebraic geometry and number theory, are Noetherian.

By analyzing the proofs of the the above results, one can show:

Theorem 1.24. Suppose that R is a Noetherian integral domain. Then every element r ∈ R, not 0 or a unit, factors into a product of irreducibles. Moreover, the following are equivalent:

(i) R is a UFD.

(ii) If r ∈ R is irreducible, then (r) is a prime ideal.

2 Euclidean domains

We turn now to finding new examples of PID’s.

Definition 2.1. Let R be an integral domain. A Euclidean norm on R is a function n: R − {0} → Z satisfying: (i) For all r ∈ R − {0}, n(r) ≥ 0.

(ii) For all a, b ∈ R with a 6= 0, there exist q, r ∈ R with b = aq + r and either r = 0 or n(r) < n(a).

An integral domain R such that there exists a Euclidean norm on R is called a .

Example 2.2. (1) R = Z, n(a) = |a|; (2) R = F [x], F a field, and n(f) = deg f, defined for f 6= 0; (3) R = F is a field, and n(a) = 0 for all a ∈ F − {0}. Here (3) is an easy exercise, and, for (1) and (2), Condition (i) of the definition is clear and (ii) is the statement of long in Z or in F [x].

8 Remark 2.3. In the definition of a Euclidean norm, we do not require that the q, r ∈ R are unique. In fact, uniqueness even fails in Z with the usual norm if we allow q and r to be negative. For example, with a = 3, b = 11, we can write 11 = 3 · 3 + 2 = 3 · 4 + (−1).

Proposition 2.4. If R is a Euclidean domain, then R is a PID.

Proof. This argument should be very familiar. Let I be an ideal of R. If I = {0}, then I = (0) is principal. Otherwise, consider the nonempty set A of nonnegative integers {n(r): r ∈ I − {0}}. By the well-ordering principle, there exists an a ∈ I − {0} such that n(a) is a smallest element of A. We claim that I = (a). Clearly (a) ⊆ I since a ∈ I. Conversely, if b ∈ I, then there exist q, r ∈ R such that b = aq + r with either r = 0 or n(r) < n(a). As b, aq ∈ I, r = b − aq ∈ I. Hence n(r) < n(a) is impossible by the choice of a, so that r = 0 and b = aq ∈ (a). Thus (a) ⊆ I and hence (a) = I.

Some remarks on submultiplicative Euclidean norms: We describe some additional assumptions we can make for Euclidean norms (since many authors add these to the definition).

Definition 2.5. The Euclidean norm n is submultiplicative if in addition n satisfies: For all a, b ∈ R − {0}, n(a) ≤ n(ab). It is multiplicative if n satisfies: For all a, b ∈ R − {0}, n(ab) = n(a)n(b). For R = Z, n(a) = |a|, or R = F a field, and n(a) = 0 for all a ∈ F − {0}, n is multiplicative, whereas deg: F [x] → Z is submultiplicative but not multiplicative. Remark 2.6. In fact, a multiplicative norm is always submultiplicative. If n is multiplicative and n(a) > 0 for all a ∈ R − {0}, then n is clearly submultiplicative. Suppose that n(a) = 0 for some a. Then 1 = aq + r with either n(r) < n(a) or r = 0. Since n(r) < 0 is impossible, r = 0, hence 1 = aq and a is a unit. For every b ∈ R − {0}, n(b) = n(a)n(a−1b) = 0, and then the above argument shows that n(b) = 0 as well. Thus n is identically zero, and hence is submultiplicative in this case as well. Note that, if n is multiplicative and n(a) = 0 for some a ∈ R − {0}, then n is identically 0 and R is a field.

Lemma 2.7. Let R be an integral domain and let n be a submultiplicative Euclidean norm on R. For all b ∈ R − {0}, exactly one of the following holds:

(i) b is not a unit, n(b) > n(1), and n(a) < n(ab) for all a ∈ R − {0}.

9 (ii) b is a unit, n(b) = n(1), and n(a) = n(ab) for all a ∈ R − {0}.

Proof. We always have n(a) ≤ n(ab), and begin by showing that n(a) = n(ab) ⇐⇒ b is a unit. First, if b is a unit, then n(a) ≤ n(ab) and n(ab) ≤ n(abb−1) = n(a), so that n(a) = n(ab). Taking a = 1, we see that n(b) = n(1). Conversely, suppose that n(a) = n(ab) for some a ∈ R − {0}. Applying long division of ab into a, we se that a = (ab)q + r, with either r = 0 or n(r) < n(ab) = n(a). We claim that r must be 0, since otherwise r = a − abq = a(1 − bq) with 1 − bq 6= 0, and hence

n(a) ≤ n(a(1 − bq)) = n(r) < n(a), a contradiction. Thus r = 0, so that a = abq and thus bq = 1, i.e. b is a unit.

Corollary 2.8. Let R be an integral domain and let n be a submultiplicative Euclidean norm on R. If r ∈ R − {0} and r = ab with neither a nor b a unit, then n(a) < n(r) and n(b) < n(r).

For a Euclidean domain with a submultiplicative Euclidean norm, we can give a direct argument for the existence of a factorization into irreducibles (which we proved in Theorem 1.19 by a somewhat complicated argument):

Proposition 2.9. If R is a Euclidean domain with a submultiplicative Eu- clidean norm n nd r ∈ R is not 0 or a unit, then r is a product of irreducibles.

Proof. Given r, not 0 or a unit, if r is irreducible we are done. Otherwise, r = r1r2, with neither r1 nor r2 a unit. Hence n(ri) < n(r), i = 1, 2. If ri is irreducible for i = 1, 2, we are done. Otherwise at least one of r1, r2 factors into factors: say r1 = ab, with n(a) < n(r1) < n(r) and n(b) < n(r1) < n(r). Clearly this process cannot continue indefinitely. A more formal way to give this argument is as follows: if there exists an r ∈ R, not 0 or a unit, which is not a product of irreducibles, then there exists an r such that n(r) is minimal among all such, i.e. if s ∈ R is not 0, a unit, or a product of irreducibles, then n(r) ≤ n(s), by the well-ordering principle. But such an r cannot be irreducible (since a single irreducible is by convention a product of one irreducible). So r = r1r2, with neither r1 nor r2 a unit, and so n(ri) < n(r), i = 1, 2. But at least one of r1 and r2 is not a product of irreducibles, since if both r1 and r2 were a product of irreducibles, then r1r2 = r would also be a product of irreducibles. Say r1 is not a product of irreducibles. Then by the choice of r, n(r) ≤ n(r1). This contradicts n(r1) < n(r). Hence no such r can exist.

10 The Euclidean in a Euclidean domain: Let R be a Eu- clidean domain with Euclidean norm n. Begin with a, b ∈ R, with b 6= 0. Write a = bq1 + r1, with q1, r1 ∈ R, and either r1 = 0 or n(r1) < n(b). Note that r1 = a + b(−q1) is a linear combination of a and b. If r1 = 0, stop, otherwise repeat this process with b and r1 instead of a and b, so that b = r1q2 + r2, with r2 = 0 or n(r2) < n(b) If r2 = 0, stop, otherwise repeat again to find r1, . . . , rk with n(r1) > n(r2) > n(r3) > ··· > n(rk) ≥ 0, with rk−1 = rkqk+1 + rk+1. Since the integers n(ri) decrease, and they are all nonnegative, eventually this procedure must stop with an rn such that rn+1 = 0, and hence rn−1 = rnqn+1. The procedure looks as follows:

a = bq1 + r1

b = r1q2 + r2

r1 = r2q3 + r3 . .

rn−2 = rn−1qn + rn

rn−1 = rnqn+1.

Then rn is a gcd of a, b and tracing back through the steps shows how to write it as a linear combination of a and b.

3 Factorization in the Gaussian integers

We now consider factorization in the Gaussian integers

Z[i] = {a + bi : a, b ∈ Z}.

We shall show that Z[i] is a new example of a Euclidean domain, hence Z[i] is a PID and a UFD. As a result, we shall describe which positive integers are a sum of two integer squares. Consider the function N : Z[i] → Z defined by N(α) = αα¯, where if α = a + bi, thenα ¯ = a − bi (i.e. N(a + bi) = a2 + b2). Since α = α, clearly N(α) = N(¯α). Note that, given n ∈ Z, n = N(α) for some α ∈ Z[i] ⇐⇒ n is a sum of two integer squares.

Lemma 3.1. The function N satisfies:

(a) N(α) ≥ 0 for all α ∈ Z[i], and N(α) = 0 ⇐⇒ α = 0.

11 (b) For all α, β ∈ Z[i], N(αβ) = N(α)N(β) (N is multiplicative). Hence, if n1 and n2 are two integers which are each a sum of two integer squares, then n1n2 is a sum of two integer squares. (c) N(α) = 1 ⇐⇒ α is a unit.

(d) There is an extension of N to a function Q(i) → Q, which we continue to denote by N, defined by N(α) = αα¯, and it satisfies (a) and (b).

Proof. (a) Clear. (b) N(αβ) = (αβ)(αβ) = (αβ)(¯αβ¯) = ααβ¯ β¯ = N(α)N(β). (c) While this follows from (c) of Lemma 2.7, we can see this either directly (N(α) = 1 ⇐⇒ α = ±1 or α = ±i) or as follows: if N(α) = 1, then αα¯ = 1 and hence α is a unit with α−1 =α ¯. Conversely, if α is a unit, then αβ = 1 for some β ∈ Z[i], hence N(αβ) = 1 = N(α)N(β). Thus N(α) is a positive integer dividing 1, so N(α) = 1. (d) Clear.

Proposition 3.2. In the integral domain Z[i], the function N(α) = αα¯ is a multiplicative Euclidean norm.

Proof. Given α, β ∈ Z[i] with α 6= 0, we must show that we can find ξ, ρ ∈ Z[i] with β = αξ + ρ and ρ = 0 or N(ρ) < N(α). Consider the quotient β/α ∈ Q[i]. Write β/α = r + si with r, s ∈ Q. Then there exist integers 1 1 n, m ∈ Z with |r − n| ≤ 2 and |s − m| ≤ 2 . Set ξ = n + mi and γ = β/α − ξ. Then β = αξ + αγ = αξ + ρ, say, where ρ = αγ. Since ρ = β − αξ, ρ ∈ Z[i]. Moreover,

12 12 1 1 1 N(γ) = N(β/α−ξ) = (r−n)2 +(s−m)2 ≤ + = + = < 1. 2 2 4 4 2 Then β = αξ + ρ with either ρ = 0 or

N(ρ) = N(αγ) = N(α)N(γ) < N(α).

Hence N is a Euclidean norm, and it is multiplicative by (b) of Lemma 3.1.

Corollary 3.3. Z[i] is a PID and a UFD.

Now that we know that Z[i] is a UFD, we can try to describe all of the irreducibles in Z[i] and relate the answer to the problem of deciding when a p is a sum of two squares. The first step is:

Lemma 3.4. (i) If N(α) = p, where p is a prime number, then α is irreducible.

12 (ii) If p is a prime number, then p is not irreducible in Z[i] ⇐⇒ p = N(α) for some α ∈ Z[i] ⇐⇒ p is a sum of two integer squares. In this case, if α divides p and α is not a unit or an associate of p, then p = N(α) and hence α is irreducible.

Proof. (i) If α = βγ, then p = N(α) = N(βγ) = N(β)N(γ), and so one of N(β), N(γ) is 1. Hence either β or γ is a unit, so that α is irreducible. (ii) If p is not irreducible, then p = αβ where neither α nor β is a unit, hence N(α) and N(β) are both greater than 1. Then p2 = N(p) = N(α)N(β), so that N(α) = N(β) = p, and α is irreducible by (i). Con- versely, if p = N(α), then p = αα¯ with N(α) = N(¯α) = p, so that neither α norα ¯ is a unit. Hence p is not irreducible in Z[i].

Lemma 3.5. If π is an irreducible element of Z[i], then there exists a prime number p such that π divides p in Z[i]. If the prime number p is also ir- reducible in Z[i], then π and p are associates, so that π = ±p or ±ip. If the prime number p is not irreducible in Z[i], then p = N(π) and every irreducible factor of p is either an associate of π or an associate of π¯.

Proof. Consider N(π) ∈ Z. Since π is not a unit, N(π) > 1, and hence N(π) is a product of prime numbers p1 ··· pr (not necessarily distinct). Since Z[i] is a UFD and π is an irreducible dividing the product p1 ··· pr, there must exist an i such that π divides pi, and we take p = pi. If p is also irreducible, then π and p are associates, and hence π = ±p or ±ip. If p is not irreducible, then we have seen that p = αα¯ for some α ∈ Z[i], where both α andα ¯ are irreducible. Hence π divides p = αα¯. It follows that π divides either α orα ¯, say π divides α, and hence that π is an associate of α since α is irreducible. Since units have norm 1, it follows that N(π) = N(α) = p.

Note that 2 is not irreducible in Z[i], and in fact 2 = N(1 + i). The irreducible factors of 2 are ±1±i, and they are all associates, since −1−i = (−1)(1 + i), 1 − i = (−i)(1 + i), and −1 + i = (i)(1 + i). Up to a unit, 2 2 is a square since 2 = (−i)(1 + i) (but it is not an actual square in Z[i]). For other primes p of the form N(α) = αα¯, this does not happen: For any α = a + bi, the associates of α are ±(a + bi) and ±i(a + bi) = ±(−b + ai). Hence, if a, b 6= 0,α ¯ = a − bi is an associate of α ⇐⇒ a = ±b. If moreover α is irreducible, then since a|(a ± ai), a = ±1 and p = 2. We may now describe the irreducibles in Z[i] as follows:

Theorem 3.6. The irreducible elements in Z[i] are:

13 (i) 1 + i and its associates ±1 ± i;

(ii) Ordinary prime numbers p ∈ Z ⊆ Z[i] congruent to 3 mod 4 and their associates ±p, ±ip;

(iii) Gaussian integers α = a + bi such that N(α) = a2 + b2 = p, where p is a prime number congruent to 1 mod 4. Moreover, for every prime number p congruent to 1 mod 4, there exists an α = a + bi such that N(α) = a2 + b2 = p.

Proof. Let π be an irreducible in Z[i]. We have seen that either π is an associate of a prime p which is irreducible in Z[i], or N(π) = p is a prime number and that the irreducible factors of p are exactly the associates of π orπ ¯. Moreover, 2 is not irreducible and the only irreducibles dividing 2 are 1+i and its associates. If p is an odd prime, p is not irreducible in Z[i] ⇐⇒ 2 2 p = a + b , where a, b ∈ Z. Since p is odd, a and b cannot be both odd or both even, so one of them, say a, is odd and the other, say b, is even. Then a2 ≡ 1 mod 4 and b2 ≡ 0 mod 4, so that p = a2 + b2 ≡ 1 mod 4. In other words, if p is an odd prime which is not irreducible in Z[i], then p ≡ 1 mod 4. Hence, if p is an odd prime with p ≡ 3 mod 4, then p is irreducible in Z[i] and its irreducible factors are its associates ±p, ±ip. Thus we will be done if we show that every odd prime number congruent to 1 mod 4 is not irreducible in Z[i], for then the remaining irreducibles of Z[i] will be the nontrivial factors of p for such primes p, which are necessarily irreducible and of norm p. To see this statement, we use the following: 2 Lemma 3.7. If p ≡ 1 mod 4, then there exists a k ∈ Z such that k ≡ −1 mod p.

Proof. The assumption p ≡ 1 mod 4 is exactly the statement that 4|p − 1. ∗ Now we know that (Z/pZ) is a cyclic group of order p−1. By known results ∗ on cyclic groups, there exists an element k of (Z/pZ) of order 4. In other 4 ∗ 2 ∗ 2 words, k = 1 in (Z/pZ) but k 6= 1 in (Z/pZ) . Since k is then a root 2 of the polynomial x − 1 = (x + 1)(x − 1) in the field Z/pZ, we must have k2 = ±1, and since by assumption k2 6= 1, k2 = −1. This says that there is an integer k such that k2 ≡ −1 mod p.

To complete the proof of the theorem, if p ≡ 1 mod 4, then we shall show 2 that p is not irreducible in Z[i]. Let k ∈ Z be such that k ≡ −1 mod p, so 2 2 that p divides k + 1. In Z[i], we can factor k + 1 = (k + i)(k − i). If p were

14 an irreducible, then since p divides k2 + 1 = (k + i)(k − i), p would divide one of the factors k ± i. But k ± i k 1 = ± i. p p p

Since ±1/p is not an integer, the quotient (k ± i)/p does not lie in Z[i]. Hence p does not divide either factor k ± i of k2 + 1, and so cannot be an irreducible.

It follows that the prime numbers which are sums of two integer squares are exactly the primes 2 and the odd primes ≡ 1 mod 4. The following describes all positive integers which are sums of two integer squares:

a1 ar Corollary 3.8. Let n ∈ N, n > 1, and write n = p1 ··· pr , where the pi are distinct prime numbers and ai ∈ N. Then n is a sum of two integer squares if and only, for every prime factor pi of n such that pi ≡ 3 mod 4, ai is even.

Proof. ⇐= : If n is as described, then every prime factor pi of n which is ai either 2 or ≡ 1 mod p is a sum of two squares, hence so is pi for an arbitrary ai positive power ai. If pi ≡ 3 mod 4, then, if ai is even, pi is also a square a1 ar since it is an even power. Thus n = p1 ··· pr is a sum of two squares since it is a product of factors, each of which is a sum of two squares. =⇒ : Suppose that n is a sum of two squares. Then n = N(α) for some α ∈ Z[i], not 0 or a unit. Factor α into a product of irreducibles: b1 bs α = uπ1 ··· πs , where u is a unit, the bi are positive integers, and πi is not an associate . If πi is not an associate of a prime pi ≡ 3 mod 4, then N(πi) is either 2 or a prime ≡ 1 mod 4. If πi is an associate of a prime pi ≡ 3 mod 4, 2 bi 2bi then N(πi) = pi and thus N(πi ) = pi . Hence

b1 bs n = N(α) = (N(π1)) ··· (N(πs)) is a product of prime powers with the property that all of the primes ≡ 3 mod 4 occur to even powers. It follows that the prime factorization of n is as claimed.

4 Examples where unique factorization fails

One can try to extend the above arguments√ to more general classes of rings. One natural ring to consider is Z[ −d], where d ∈ N. We usually assume that d has no squared prime factors, in other words that either d = 1 or

15 √ √ 2 d = p1 √··· pk is a product of distinct primes,√ since −a e = a −e. Note that Z[ −d] is a of the field Q( −d),√ which is called an imaginary quadratic field. Similarly, we could look at Z[ √d], where d ∈ N, d > 1, and d has√ no squared prime factors. In this case Z[ d] is a subring of the field Q( d), which is called a real quadratic field. √ There√ is a natural√ N : Z[ −d] → Z defined by, if α = a + b −d ∈ Z[ −d],

N(α) = αα¯ = a2 + db2.

Just as in the case d = 1, N is multiplicative,√ i.e. N(αβ) = N(α)N(β), and N extends to a function from√Q( −d) to Q which is a homomorphism of ∗ ∗ multiplicative groups from Q( −d) to Q . Adapting the arguments in the preceding section for Z[i], it is not hard to show: √ Proposition 4.1. In the integral domain Z[ −2], the function N(α) = αα¯ is a multiplicative Euclidean norm. √ However, this fails for every d >√2: If d > 2, the norm N : Z[ −d] → Z fails to be Euclidean, and in fact Z[ −d] is not a UFD. √ Example√ 4.2. The integral domain Z[ −3] is not a UFD. In fact, in Z[ −3], √ √ 4 = 2 · 2 = (1 + −3)(1 − −3). √ We will show that√ 2 and 1 ± −3 are all irreducible, and that 2 is not an associate√ of 1 ± −3. First, arguing as for Z[i], it is easy to check that α √∈ Z[ −3] is a unit ⇐⇒ N(α) = 1. Now suppose that 2 factors in Z[ −3]: say 2 = αβ. Then N(α)N(β) = N(2) = 4. If neither α nor β is a unit, then√ N(α) > 1 and N(β) > 1, hence N(α) = N(β) = 2. But if say 2 2 2 α = a + b −3 with a, b ∈ Z, then a + 3b = 2, hence b = 0√ and a = 2, which is impossible. Thus 2 is irreducible,√ and since N(1 ± −3) = 4 as well,√ a similar argument shows that 1 ± −3 is irreducible. Finally, 2√ and 1 + √−3 are not associates,√ since if they were,√ then 2√ would divide 1 + −3 in Z√[ −3]. But (1 + −3)/2 = 1√/2 + (1/2) −3 ∈/√Z[ −3]. Likewise, 2 and 1 − −3 are not associates in Z[ −3]. Hence Z[ −3] is not a UFD. √ This example is slightly misleading, because Z[ −3] is a subring of a somewhat more natural ring which is in fact a UFD: Let ω = e2πi/3 = 1 1 √ − 2 + 2 −3 be a cube . Note that ω is a root of the monic 2 3 3 2 polynomial x +x+1, since ω is a root of x −1 and√x −1 = (x−1)(x +x+1). 3 2 −1 2 Note that, since ω = 1, ω = ω =ω ¯. Hence −3 = ω − ω ∈ Z[ω], so

16 √ that Z[ −3] is a subring of Z[ω]. More generally, we say that an α ∈ C is an if α is a root of a monic polynomial with integer coefficients, i.e. f(α) = 0, where f ∈ Z[x] is monic. (It is easy to see that every algebraic number is a root of a polynomial f ∈ Z[x], but f is not necessarily monic.) Them if E ≤ C is an algebraic extension of Q, one can show that the set of algebraic integers in E is a subring of E whose quotient field is E, and this ring plays the role of the subring Z of Q. For E = Q(i√), for example, the subring of algebraic integers is just Z[i], but for E = Q( −3), the subring of algebraic integers is Z[ω]. In this particular example, Z[ω] is in fact a PID and hence a UFD. In fact, one can use this to give a fairly straightforward solution of Fermat’s Last Theorem for the case p = 3: The only solutions in integers to the equation x3 + y3 = x3 are the trivial solutions, where one of x, y, z is 0. However, unique√ factorization still turns out to be a rare occurrence. For example,√ Z[ −5] turns out to be the full subring of algebraic integers in Q( −5), but it is easy to check that √ √ 6 = 2 · 3 = (1 + −5)(1 − −5) gives a factorization of 6√ into a product of irreducibles in two essentially different ways. Hence Z[ −5] is not a UFD, and hence it is not a PID. More generally, a famous theorem due to Heegner-Stark says that there is a finite (and relatively short) list of imaginary quadratic fields whose rings of integers are UFD’s. It turns out that, for such a ring, being a UFD is equivalent to being a PID. However, not every such PID is Euclidean. Hence, the implications R Euclidean =⇒ R is a PID =⇒ R is a UFD are all strict, in the sense that none of them is an ⇐⇒ statement. In other words, there exist rings R which are a UFD but which are not a PID and there exist rings R which are a PID but which are not Euclidean. Much of the√ above discussion carries over to real quadratic√ fields. For example, for Z[ 2], we have a multiplicative function N : Z[ 2] → Z defined by √ N(a + b 2) = |a2 − 2b2|. One can check that, at least in this case, N is a Euclidean norm. For general real quadratic fields, one can define an analogous multiplicative function N, which will usually not however be a Euclidean norm. It is unknown if there are finitely or infinitely many real quadratic fields whose rings of integers are UFD’s. More general rings along these lines are studied in algebraic number the- ory. Historically, one major motivation was the attempt to prove Fermat’s

17 Last Theorem (which would be quite straightforward to prove if unique fac- torization always held in such rings). Ideals were originally introduced in this special context. If R is a UFD, then r is irreducible ⇐⇒ (r) is prime, and (r) uniquely determines r up to associates. In general, nonzero prime ideals can function to a certain extent as a substitute for irreducible ele- ments. In fact, a basic theorem in states that, if R is a ring of algebraic integers (correctly defined), and I is an ideal of R, with I 6= {0},R, then I can be uniquely factored into a product of prime ideals (where ideal products were defined in the HW).

18