Processing math: 5%

Tuesday, May 28, 2013

Alternate Proof of Hilbert's Basis Theorem and Gröbner Bases (9.6.43-44)

Dummit and Foote Abstract Algebra, section 9.6, exercises 43-44:

MathJax TeX Test Page 43. Let S={N | N not a finitely generated monomial ideal} and assume by way of contradiction that S.

(a) Prove S has a maximal element.
(b) Show there exist monomials x,yM such that xyM.
(c) Show there exists M0M with M0 a finitely generated monomial ideal such that M0+(x)=M+(x) and M=M0+(x)(M:(x)). Deduce M is finitely generated, forcing S=.

44. If IF[x1,..,xn] is an ideal, prove LT(I) is finitely generated. Conclude I has a Gröbner basis and deduce Hilbert's Basis Theorem.

Proof: (a) Let C be a typical chain totally ordered by inclusion. Since M=NCN is a monomial ideal by 9.6.39, by Zorn's Lemma it suffices to show it is infinitely generated; assume the contrary, so that M=(n1,...,nr). Let ni be maximal with regard to the ordering niNinjNj:⇔NjNi, and observe NiM=(n1,...,nr)Ni so that Ni=M is finitely generated, a contradiction.

(b) By 9.6.42(a), the only prime monomial ideals are finitely generated. Therefore there exist polynomials x,yF[x1,...,xn] such that x,yM and xyM. By 9.6.10, there exist monomial terms in x,y not divisible by any mi for M=(m1,...). Showing xy=mM and expanding x to subtract the products of y and monomial terms of x divisible by monomials of M (which are elements of M), we obtain xyM for x composed of terms not divisible by monomials of M. Similarly, xyM. Now LT(xy)=LT(x)LT(y)LT(M)M (the latter being due again to 9.6.10) so that LT(x) and LT(y) are monomials not within M whose product is within M.

(c) Fix a monomial ordering . First note that mmmd=m so that, since d1, we have md=mm, i.e. mmmm.

Let x be the minimal monomial by not within M such that there exists another monomial y not within M fulfilling xyM. Revisit M=(m1,...) and let M=(m1,...,x) as before. Since MM by the maximality of M we must have M=(T) for some finite subset of monomials T={q1,...,qt}. Reformat this generating subset so that it is minimal, by excluding a generator n if M=(T{n}). This process is seen to terminate, and in part we are left with T={n1,...,ns} where ni for all i≠j. Assume x∉T: Then by 9.6.10 since x∈M^* we have n_i \mid x so n_id=x for some i, and since n_i∈M⇒x∈M, yet also n_i∉M⇒n_i < x \land n_i(dy)∈M with dy∉M since d < x as well, we inevitably violate the minimality of x, and we must instead have x∈T. Rewrite T=\{n_1,...,n_{s-1},x\}.

Now let M_0=(T \setminus \{x\}). Assume n_j∈T \setminus \{x\} yet n_j∉M: Since the elements of T are from M^* generated by \{m_1,...,x\} we must have either m_i \mid n_j for some i or x \mid n_j; the latter is excluded by the minimalization of T and the former implies n_j∈M. Therefore M_0 \subseteq M, and evidently M_0+(x)=M+(x)=M^*.

Show M=M_0+(x)(M : (x)): (\supseteq) By the definition of (M : (x)), the direction is clear. (\subseteq) We have m_i∈M \subseteq M_0+(x), so that either n_k \mid m_i ⇒ m_i∈M_0 or x \mid m_i ⇒ m_i ∈ (x)(M : (x)).

Finally, we see that M_0 is finitely generated, and also y∉M with y∈(M : (x)) so that (M : (x)) is finitely generated by the maximality of M, and also (x)(M : (x)) and now M_0+(x)(M : (x))=M is finitely generated, a contradiction.~\square

44. Since LT(I) is a monomial ideal, by the previous it is finitely generated. In the same way as in 9.6.1, we may generate each of the elements of the finite generating set of LT(I) by a finite number of finite combinations of LT(g_i) for g_i elements of I, so that by proposition 24 these g_i form a Gröbner basis and as well a finite generating set for the arbitrary ideal I, which is Hilbert's Basis Theorem.~\square

Monday, May 27, 2013

Prime and Maximal Monomial Ideals (9.6.42)

Dummit and Foote Abstract Algebra, section 9.6, exercise 42:

MathJax TeX Test Page (a) Show M is a monomial prime ideal if and only if M=(S) for some S \subseteq \{x_1,...,x_n\}.
(b) Show (x_1,...,x_n) is the only maximal monomial ideal.

Proof: (a) (\Leftarrow) Letting I be such that S=\{x_i~|~i∈I\}, we have F[x_1,...,x_n]/M ≅ F[x_j~|~j∉I] which is again a polynomial ring in several variables over a field and is thus in particular an integral domain. () Since M is a monomial ideal, let M=(m_1,...) with the monomials countably ordered by m_1 ≤ m_2 if \text{deg}(m_2) < \text{deg}(m_1) and decided by lexicographic ordering if \text{deg}(m_2)=\text{deg}(m_1). Trim this generating list by a process of eliminating m_i if it is not generatable by linear combinations of previous monomials. Choose m∉\{x_1,...,x_n\} in the generating set. letting x_i be a variable dividing m, we must have x_i∉M else by exercise 10 x_i is divisible by a term in the generating list, forcing x_i in the generating list so that by the ordering x_i appears before m and now m could not have been included. Since x_i was arbitrary, we have the product of the variables dividing m violating the primality of M. Therefore m does not exist, and this generating subset is a subset of \{x_1,...,x_n\}.

(b) Since maximal ideals are prime ideals, if M=(S) with S≠\{x_1,...,x_n\}, by part (a) we have M \subset (x_1,...,x_n) \subset F[x_1,...,x_n], a contradiction.~\square

Saturday, May 25, 2013

Minimal Gröbner Bases' Commonalities (9.6.15)

Dummit and Foote Abstract Algebra, section 9.6, exercise 15:

MathJax TeX Test Page Fix a monomial ordering on R=F[x_1,...,x_n].
(a) Prove \{g_1,...,g_n\} \subseteq I \subseteq R for I an ideal and monic g_i is a minimal Gröbner basis for I if and only if \{LT(g_1),...,LT(g_n)\} is a minimal generating set for LT(I).
(b) Prove the leading terms and order of a minimal Gröbner basis of I is uniquely determined.

Proof: (a) By the definition of a minimal Gröbner basis, we have (LT(g_1),...,LT(g_n))=LT(I) and LT(g_i) \not \mid LT(g_j) for any i≠j. We can see that (\{LT(g_1),...,LT(g_n)\} \setminus \{LT(g_i)\}) \subset I since for there to be equality would be, by exercise 10, to presume an element in \{LT(g_1),...,LT(g_n)\} \setminus \{LT(g_i)\} divides LT(g_i). \Leftarrow By proposition 24 we have \{g_1,...,g_n\} is a Gröbner basis of I and since their leading terms is a minimal generating set for I we must have LT(g_i) \not \mid LT(g_j) for i≠j.

(b) Since LT(I) is a monomial ideal, by the previous exercise its minimal generating set of monomials is uniquely determined. By part (a) the leading terms of a minimal Gröbner basis's elements must be uniquely determined, as well as the number of leading terms (which is the order of the basis).~\square

Uniqueness of Minimal Generating Monomial Subsets of Ideals (9.6.14)

Dummit and Foote Abstract Algebra, section 9.6, exercise 14:

MathJax TeX Test Page Suppose I is a monomial ideal in F[x_1,...,x_n] minimally generated by M=\{m_1,...,m_k\} for m_i a monomial. Prove this generating subset is unique.

Proof: Note we must necessarily have m_i \not \mid m_j for all i≠j or else it is not a minimally generating set. Now, assume another minimally generating set N=\{n_1,...,n_l\}. Since (M)=(N)=I we have by exercise 10 that LT(n_1)=n_1 is divisible by, say, m_1. Similarly, m_1 is divisible by a monomial from N. Continue in this fashion so that we must necessarily reach a point in the chain of division such that m_i repeats, so that m_i \mid ... \mid n_j \mid ... \mid m_i and now m_i=n_j. Starting again with an element m in M besides m_i, we can follow the chain once more which will reveal another equality between elements, which won't be the one between m_i and n_j as then m \mid m_i, a contradiction. In this fashion, every element of the minimally-ordered generating set will be paired up with its equal in the opposite set, i.e. will be a subset of the opposite set. By the nature of minimally generated sets, this inclusion cannot be proper, and now M=N.~\square

Tuesday, May 21, 2013

Multiplicative and Additive Independence (9.5.7)

Dummit and Foote Abstract Algebra, section 9.5, exercise 7:

MathJax TeX Test Page For F a field, prove F^\times \not ≅ F^+.

Proof: Assume F is finite. Then |F^\times|=|F^+|-1, so they cannot be in bijection. Therefore F is infinite: Assume -1=1; then 2=0 implying a+a=0 for all a∈F^+ so that u^2=1 for all u∈F^\times, so that there are an infinite number of roots to the equation x^2-1, an impossibility. Therefore 1≠-1 so that 2a=a+a≠0 for all nonzero a∈F^+, so that for all nonidentity u∈F^\times we have u^2≠1, a contradiction by (-1)^2=1.~\square

Saturday, May 18, 2013

Polynomial Factorization Practice #2 (9.4.1-4)

Dummit and Foote Abstract Algebra, section 9.4, exercise- 1-4:

MathJax TeX Test Page 1. Factor the following polynomials:
(a) x^2+x+1∈\mathbb{F}_2[x]
(b) x^3+x+1∈\mathbb{F}_3[x]
(c) x^4+1∈\mathbb{F}_5[x]
(d) x^4+10x+1∈\mathbb{Z}[x]
2. Prove the following polynomials are irreducible in \mathbb{Z}[x]:
(a) x^4-4x^4+6
(b) x^6+30x^5-15x^3+6x-120
(c) x^4+4x^3+6x^2+2x+1
(d) \dfrac{(x+2)^p-2^p}{x} for p an odd prime.
3. Show that f_n(x)=(x-1)(x-2)...(x-n)-1 is irreducible over \mathbb{Z} for all n \geq 1.
4. Show that f_n(x)=(x-1)(x-2)...(x-n)+1 is irreducible over \mathbb{Z} for all n \geq 1 and n≠4.

Proof: (1) (a) There are no zeros to this equation, so it is irreducible.
(b) By discovering f(1)=0, we obtain x^3+x+1=(x-1)(x^2+x+2), where the latter has no zeros.
(c) x^4+1=x^4-4=(x^2-2)(x^2+2).
(d) Reduce modulo 5. If f(x)=a(x)b(x), then due to monicity of f(x) we have a(x) and b(x) are monic and not collapsing to units so that \overline{a(x)} is associate to \overline{x^2+2} or \overline{x^2-2}, say the former. Since \overline{a(x)} is monic we have \overline{a(x)}=\overline{x^2+2} which forces the constant term of a(x) to not be \pm 1, a contradiction.

Alternatively, since f(x) clearly has no zeros in \mathbb{Z}, it would have to factor as the product of two quadratics such that the following hold:f(x)=a(x)b(x)a(x)=x^2+a_1x+a_0b(x)=x^2+b_1x+b_0a_0b_0=1a_1b_0+a_0b_1=0a_1b_1+a_0+b_0=10a_1+b_1=0Solving for b_1 as a typical system of equations reveals\dfrac{1}{4}b_1^4+15b_1^2+24=0which is impossible.

(2) (a) Eisenstein's criterion for p=2 applies.
(b) Eisenstein applies for p=3.
(c) We generalize the technique employed in example 2. Lemma 1: Let R be an integral domain with f(x)=a_0+a_1x+...+a_nx^n∈R[x] and g(x)=f(r(x)) for r(x) a nonconstant polynomial. If g(x) is irreducible, then f(x) is irreducible. Proof: If f(x)=a(x)b(x) for some nonunit polynomials, then g(x)=a(r(x))b(r(x)) where necessarily \text{deg}(a(r(x)))=\text{deg}(a(x)) \cdot \text{deg}(r(x)) \geq 1 if \text{deg}(a(x)) \geq 1 and a(r(x))=a(x) if a(x) is a constant and similarly for b(r(x)), a contradiction.~\square

Let g(x)=f(x-1)=x^4-2x+2, where Eisenstein applies for p=2.
(d) \dfrac{(x+2)^p-2^p}{x}=x^p+...+2p, where by the binomial theorem the omitted terms have coefficients of multiples of p. Eisenstein applies for p.

(3) When n=1 we clearly have f_n(x) is irreducible, so assume n > 1. Assuming f_n(x) has a zero in \mathbb{Z}, then 1=(x-1)(x-2)...(x-n) so that in particular x-1=\pm 1 and x-2=\pm 1, a contradiction. Therefore if f_n(x)=a(x)b(x) then the degrees of a(x) and b(x) are greater than one.

We demonstrate a basic algebraic result, first. If R is an integral domain and f(x)∈R[x] of degree n \geq 1, then there are at most n zeros of f(x). If not, then f(x)=(x-\alpha_1)(x-\alpha_2)...(x-\alpha_{n+1})f'(x) entails \text{deg}(f(x)) > n, a contradiction.

Since f_n(m)=-1 for all integers 1≤m≤n we have a(m)= \pm 1 and b(m)= \mp 1 for at least n values. We thus have a(x)+b(x) has at least n zeros, so that necessarily a(x)=-b(x). But now the largest term's coefficient of a(x)b(x) is -1, a contradiction.

(4) Similar as in the last problem, f_1(x) is irreducible and f_n(x) has no zeros for n > 1, so that necessarily a(x)=b(x)=\pm 1 for at least n values, and again a(x)=b(x) so that f_n(x) is a perfect square. However, when n is odd this is an impossibility by degree, so n must be even. For n < 4 we can observe f_n(4) is irreducible, and for n > 4 and n even we can observe f_n(3/2) is negative, a contradiction since f_n(x) is a perfect square.

When n=4, with a process similar to that for (1)(d), we can observe f_4(x)=(x^2-5x+5)^2 is its decomposition.~\square

Friday, May 17, 2013

Bezout Subring of Q[x] (9.3.5)

Dummit and Foote Abstract Algebra, section 9.3, exercise 5:

MathJax TeX Test Page Let R=\mathbb{Z}+x\mathbb{Q}[x] \subset \mathbb{Q}[x].

(a) Let f(x),g(x)∈\mathbb{Q}[x] with x^r the maximal power of x dividing both f(x) and g(x), and let f_r and g_r be the coefficients of x^r in f(x) and g(x) respectively. Then \mathbb{Z}f_r+\mathbb{Z}g_r=\mathbb{Z}d_r for some nonzero d_r\mathbb{Q}. Prove there is a polynomial d(x)∈\mathbb{Q}[x] such that d(x) is a gcd of f(x) and g(x), and its minimal term is d_rx^r.
(b) Prove f(x)=d(x)q_1(x) and g(x)=d(x)q_2(x) for some q_1(x),q_2(x)∈R.
(c) Prove d(x)=f(x)a(x)+g(x)b(x) for some a(x),b(x)∈R.
(d) Conclude Rf(x)+Rg(x)=Rd(x) in \mathbb{Q}[x], implying in part that R is a Bezout domain.
(e) Show that there must be ideals of R which are not finitely generated, and in fact x\mathbb{Q}[x] is an ideal of R that is not finitely generated.

Proof: (a) We can compute the gcd of f(x) and g(x) in the Euclidean domain \mathbb{Q}[x]. By the Euclidean algorithm we have d(x)=v_1(x)f(x)+v_2g(x) for some polynomials v_1(x),v_2(x)∈\mathbb{Q}, so that x^r \mid d(x). Furthermore, letting f_r be the one of its pair that is guaranteed nonzero, by d(x) \mid f(x) we must have the coefficient of x^r in d(x) is nonzero. Since this gcd is unique up to multiplication by a nonzero rational number, we can arrange for a gcd d(x) with minimal term d_rx^r.

(b) Since \mathbb{Z}f_r+\mathbb{Z}g_r=\mathbb{Z}d_r we have d_rq_1=f_r and d_rq_2=g_r for some q_1,q_2∈\mathbb{Z}. By definition of a gcd we have f(x)=d(x)q_1(x) and g(x)=d(x)q_2(x) for some q_1(x),q_2(x)∈\mathbb{Q}[x], and by observation we see the constant terms of q_1(x) and q_2(x) are q_1 and q_2 respectively, putting these polynomials in R.

(c) We can observe d(x)=f(x)a(x)+g(x)b(x) for some a(x),b(x)∈\mathbb{Q}[x], and by 9.2.11 we have a full solution set for a(x),b(x) given bya'(x)=a(x)+m(x) \dfrac{g(x)}{d(x)}b'(x)=b(x)-m(x) \dfrac{f(x)}{d(x)}as m(x) ranges over \mathbb{Q}[x].

Assume f_r and g_r are nonzero. First, observe that d_r=f_r\alpha+g_r\beta for some \alpha,\beta∈\mathbb{Z}, so that 1=\dfrac{f_r}{d_r}\alpha+\dfrac{g_r}{d_r}\beta, implying that v_f=\dfrac{f_r}{d_r} is relatively prime to v_g=\dfrac{g_r}{d_r} (which are, foremost, seen to be integers).

Letting a and b be the constant terms of a(x) and b(x) respectively, we can see by our mentioned equation that (f_rx^r)a+(g_rx^r)b=d_rx^r, so that f_ra+g_rb=d_r implying v_fa+v_gb=1. As such, we can see v_fz_1+v_gz_2=v_fa+v_gb for some z_1,z_2∈\mathbb{Z}, so that v_fz_1-v_fa=-v_gz_2+v_gb and we can set m=\dfrac{z_1-a}{v_g}=\dfrac{-z_2+b}{v_f}so that with some manipulation and replacement we havez_1=a+m\dfrac{g_r}{d_r}z_2=b-m\dfrac{f_r}{d_r}Now, by letting m(x)=m and taking the solutions a'(x) and b'(x) as above, by noting that the constant terms of \dfrac{f(x)}{d(x)} and \dfrac{g(x)}{d(x)} must be \dfrac{f_r}{d_r} and \dfrac{g_r}{d_r} respectively we see that the equations above actually describe the constant terms of a'(x) and b'(x), which implies a'(x),b'(x)∈R.

Assume one of f_r or g_r is zero, letting f_r be the nonzero. Since now d_r \mid f_r and f_r \mid d_r we have f_r=\pm d_r, so that in the equation d(x)=f(x)a(x)+g(x)b(x) we can see the constant term of a(x) must be \pm 1 as the latter summand doesn't affect the sum coefficient of x^r. Continuing the terminology of the last procession we still have the same complete set of solutions, although observing the constant terms z_1 and z_2 of a'(x) and b'(x) simplifying toz_1=a+m\dfrac{g_r}{d_r}=a=\pm 1z_2=b-m\dfrac{f_r}{d_r}=b-mSo choose m(x)=b so that a'(x),b'(x)∈R are valid solutions.

(d) This is evident from parts (a) and (c) as greatest common divisors in R both exist and are expressible as R-linear combinations of elements from R, so that (~f(x),g(x)~)=(~d(x)~). This implies every finitely generated ideal of R is principle, i.e. R is a Bezout domain.

(e) Since R has been shown in the previous exercise to not be a UFD, it cannot be a PID, so that R must contain infinitely generated nonprinciple ideals. Now, assume x\mathbb{Q}[x]=(a_1,...,a_n) is finitely generated; then x\mathbb{Q} \subseteq x\mathbb{Q}[x]=a_1R+...+a_nR. By observing the first degree terms, we have \mathbb{Q}=a_1\mathbb{Z}+...+a_n\mathbb{Z}, which translates to saying that \mathbb{Q} as an additive group is finitely generated by the a_i, a contradiction by 2.4.14(d).~\square

Thursday, May 16, 2013

Polynomial Factorization Practice (9.3.4)

Dummit and Foote Abstract Algebra, section 9.3, exercise 4:

MathJax TeX Test Page Let R=\mathbb{Z}+x\mathbb{Q}[x] \subset \mathbb{Q}[x].

(a) Prove R is an integral domain and R^\times = \pm 1.
(b) Show that the irreducibles of R are the irreducibles of \mathbb{Z} together with the irreducible polynomials of \mathbb{Q}[x] with constant term \pm 1. Prove these irreducibles are prime.
(c) Show x cannot be written as the product of irreducibles in R (in particular, x is not irreducible) and conclude that R is not a UFD.
(d) Show x is not prime in R and describe R/(x).

Proof: (a) Since this set is clearly closed under subtraction and multiplication, it is a subring, and by its containment in \mathbb{Q}[x] it is an integral domain. Note that the degree norm N has the property N(\alpha \beta)=N(\alpha)+N(\beta) as in a typical polynomial ring, so that the only possible units are within \mathbb{Z}, and as such the only units are \pm 1.

(b) Should p=\alpha \beta we have 0=N(p)=N(\alpha)+N(\beta), so that \alpha,\beta∈\mathbb{Z} and thus by the primality of p we have one of the two is equal to \pm 1, i.e. is a unit. Letting p(x)=q_nx^n + ... + q_1x \pm 1 be an irreducible element of \mathbb{Q}[x] with constant term \pm 1, assuming p(x)=\alpha \beta for nonunit \alpha,\beta∈R, by necessity \alpha and \beta have constant term \pm 1 so that N(\alpha),N(\beta) > 0. But now \alpha and \beta are nonunits in \mathbb{Q}[x], violating the irreducibility of p(x). Finally, assuming a(x) is an irreducible element of R by necessity we have its constant term is \pm 1 else for a prime divisor p of the term we have a(x)=p \dfrac{a(x)}{p} is factorization when a(x)≠\pm p. Assuming a(x)=b(x)c(x) for some nonunit b(x),c(x)∈\mathbb{Q}[x], we have the constant term of b(x) is \dfrac{m}{n} and necessarily the constant term of c(x) is \dfrac{n}{m} so that b(x)=\dfrac{m}{n}b'(x) and c(x)=\dfrac{n}{m}c'(x) for b'(x),c'(x)∈R and now a(x)=b'(x)c'(x) is a valid factorization of a(x) in R by 0 < N(b(x))=N(b'(x)) and 0 < N(c(x))=N(c'(x)) so that b'(x),c'(x)∉R^\times.

For prime p the ideal (p) \subseteq R is the kernel of the surjective homomorphism φ : R → \mathbb{Z}/p\mathbb{Z} by q_nx^n + ... + q_1x + z \mapsto z + p\mathbb{Z}. For irreducible q∈\mathbb{Q}[x] with constant term \pm 1, we have q is prime and thus R/(q) is contained in the integral domain \mathbb{Q}/(q) and is therefore an integral domain itself, proving q is prime.

(c) Note that any vx∈R for v∈\mathbb{Q} is not irreducible, since vx=(2)(\dfrac{v}{2}x) is a valid factorization. Now, if x=\pi_1...\pi_n for irreducible \pi_i, then 1=N(\pi_1)+...+N(\pi_n) so that exactly one \pi_k is of the form qx+z and for i≠k we have \pi_i∈\mathbb{Z}. By our remark above z≠0, but now the product \pi_1...\pi_n has a nonzero constant term, a contradiction.

(d) If x were prime in R, it would be irreducible, a contradiction by the above. We now claim that R/(x)=\{\overline{a+bx}~|~a∈\mathbb{Z},b∈\mathbb{Q} \land 0≤b<1\}. Since clearly \supseteq, we must show the reverse: For arbitrary a(x)∈R, we have a(x)=q_nx^n + ... + q_1x+z=(q_nx^{n-1}+...q_2x)x+(w+q_1')x+z=(q_nx^{n-1}+...q_2x+w)x+q_1'x+z for some w∈\mathbb{Z} and 0≤q_1'<1, so that \overline{a(x)}∈\{\overline{a+bx}~|~a∈\mathbb{Z},b∈\mathbb{Q} \land 0≤b<1\}. From here, the isomorphism from R/(x) onto \mathbb{Z}+(\mathbb{Q}/\mathbb{Z})x is evident, where \mathbb{Q}/\mathbb{Z} is the ring extension of \mathbb{Q}/\mathbb{Z} as additive groups by standard multiplication.~\square

Tuesday, May 14, 2013

Polynomial Ideals and Primality (9.1.13)

Dummit and Foote Abstract Algebra, section 9.1, exercise 13:

MathJax TeX Test Page Letting F be a field, prove F[x,y]/(y^2-x) \not ≅ F[x,y]/(y^2-x^2).

Proof: We show something a bit stronger, that F[x,y]/(y^n-x) \not ≅ F[x,y]/(x^2-y^2) for any n, with F only an integral domain.

First we show y-x,y+x∉(y^2-x^2)=(y-x)(y+x) so that there are zero divisors in F[x,y]/(y^2-x^2). Assuming the contrary we have for f some polynomial y-x=(y-x)(y+x)f⇒(y+x)f=1, an impossibility. Similarly for y+x∉(y^2-x^2).

Now we show F[x,y]/(y^n-x)=F[y]/(y^n-x)≅F[y] is an integral domain, which will complete the proof. To see the first equality, notice that \supseteq is clear and any polynomial of F[x,y]/(y^n-x) can be represented by the image of a polynomial devoid of variables in x since \overline{y}^n=\overline{x}. Letting φ be the natural surjective homomorphism of F[y] onto F[y]/(y^n-x), we can observe that if \text{ker}~φ≠0 then 0≠(y^n-x)f∈F[y] for some f∈F[x,y], an impossibility by (y^n-x)f=y^nf-xf and observing the result of the xf term by choosing the result of the multiplication monomial with maximal degree of x in f, which cannot be affected by any terms in xf (else the f monomials' y and x degrees are the same and can be rewritten as one) nor in y^nf.~\square

Monday, May 13, 2013

Prime Ideal Squared (9.1.12)

Dummit and Foote Abstract Algebra, section 9.1, exercise 12:

MathJax TeX Test Page Let the overbar denote passage from \mathbb{Q}[x,y,z] into \mathbb{Q}/(xy-z^2). Prove that \overline{P}=(\overline{x},\overline{z}) is a prime ideal. Show that \overline{xy}∈\overline{P}^2 but that no power of \overline{y} lies in \overline{P}^2, showing that \overline{P} is a prime ideal whose square is not a primary ideal.

Proof: It is easy to check that for a surjective homomorphism φ, we have (φ(a_1),φ(a_2),...)=φ(a_1,a_2,...), so that (\overline{x},\overline{z})=(x,z)/(xy-z^2). Evidently (xy-z^2) \subseteq (x,z) so that \overline{\mathbb{Q}[x,y,z]}/\overline{(x,z)}≅\mathbb{Q}[x,y,z]/(x,z)≅\mathbb{Q}[y] is an integral domain, so that \overline{P} is a prime ideal. Furthermore, \overline{xy-z^2}=\overline{0} so that \overline{xy}=\overline{z}^2∈\overline{P}^2 as claimed. Now, if \overline{y}^n∈\overline{P}^2 \subseteq \overline{P} for some n∈\mathbb{Z}^+ then we have the natural image of \overline{y}^n in the previously mentioned quotient isomorphic to \mathbb{Q}[y] is zero, an impossibility by the commutativity of natural projections.~\square

Sunday, May 12, 2013

Characterization of PIDs (8.3.11)

Dummit and Foote Abstract Algebra, section 8.3, exercise 11:

MathJax TeX Test Page ProveR~\text{a Principle Ideal Domain}~⇔R~\text{a UFD}~\land R~\text{a Bezout Domain}Proof: The direction is evident, so it suffices to demonstrate the converse. Observe a typical ideal I \subseteq R, and choose arbitrary x∈I with x=u\pi_1^{\alpha_1}...\pi_n^{\alpha_n} being its unique decomposition up to units. For any e∈\mathbb{Z}^+ \cup \{0\} and i with 1≤i≤n let A_{i,e}=\{r∈I ~ | ~ \pi_i^e \mid r\}. For each i let \beta_i be the largest integer such that A_{i,\beta_i}=I, which is guaranteed to exist for arbitrary i as the exponent \alpha_i of \pi_i is finite so that as a bound x∉A_{i,\alpha_i+1}. For each i, choose a_i such that a_i ∈ I \setminus A_{i,\beta_i+1}, so that a_i=\pi_i^{\beta_i}n_i with \pi_i \not \mid n_i. Letting A=\bigcup_{1≤i≤n} a_i and taking successive greatest common divisors of the a_i and then taking a greatest common divisor of this and x, we have a final common divisor d=\pi_1^{\beta_1}...\pi_n^{\beta_n} (since these are the only irreducibles dividing x, and for each i there is an element a_i with its minimal \pi_i exponent \beta_i, and all the exponents of the a_j and x of \pi_i is at least \beta_i as a_j,x∈I=A_{i,\beta_i}), writable as an R-linear combination of elements from A \cup \{x\} \subseteq I by 7.2.7. We claim (d)=I: Evidently (d) \subseteq I, and for the reverse containment, for any r∈I we have \pi_i^{\beta_i} \mid r for applicable i since r∈I=A_{i,\beta_i}, implying I \subseteq (d) and now I=(d) is a principle ideal.~\square

Saturday, May 11, 2013

Quotients of the Gaussian Integers (8.3.6-7)

Dummit and Foote Abstract Algebra, section 8.3, exercises 6-7:

MathJax TeX Test Page 6. (a) Prove \mathbb{Z}[i]/(1+i) is a field of order 2.
(b) Let q∈\mathbb{Z} be a prime with q ≡ 3 \mod{4}. Prove \mathbb{Z}[i]/(q) is a field of order q^2.
(c) Let p∈\mathbb{Z} be a prime with p ≡ 1 \mod{4} and let p=\pi \overline{\pi} be its factorization in \mathbb{Z}[i]. Prove \mathbb{Z}[i]/(p)≅\mathbb{Z}[i]/(\pi) \times \mathbb{Z}[i]/(\overline{\pi}) as rings, \mathbb{Z}[i]/(p) is of order p^2 and therefore \mathbb{Z}[i]/(\pi) and \mathbb{Z}[i]/(\overline{\pi}) are fields of order p.
7. Let \pi∈\mathbb{Z}[i] be irreducible.
(a) For n \geq 0 an integer, show that (\pi^{n+1}) \subseteq (\pi^n) and that multiplication by \pi^n induces an isomorphism \mathbb{Z}[i]/(\pi)≅(\pi^n)/(\pi^{n+1}) as additive abelian groups.
(b) Prove |\mathbb{Z}[i]/(\pi^n)|=|\mathbb{Z}[i]/(\pi)|^n.
(c) Prove for \alpha≠0 that |\mathbb{Z}[i]/(\alpha)|=N(\alpha).

Proof: (6) (a) By the division algorithm, every element in \mathbb{Z}[i]/(1+i) can be represented by an element of norm < 2. We have -i=-(1+i)+1,i=(1+i)-1, and -1=-(1+i)(1-i)+1, so that the order is ≤2. Since 1+i is not a unit, we have (1+i) \subset \mathbb{Z}[i], so that the order must be 2. Since \mathbb{Z}[i]/(1+i) is now a finite integral domain by the primality of 1+i, we have it is a field.

(b) Observe arbitrary a+bi∈\mathbb{Z}[i]. Let a=a'+v_aq and b=b'+v_bq with 0≤ a',b' < q. We have a+bi=(a'+v_aq)+(b'+v_bq)i=a'+b'i+q(v_a+v_bi), so that in this sense a+bi can be reduced modulo q. This allows a maximum order of q^2. Now, assume reduced \overline{a+bi} = \overline{c+di} in \mathbb{Z}[i]/(q), so that a+bi=c+di+qr for some r∈\mathbb{Z}[i]. We therefore have (a-c)+(b-d)i=qr, i.e. q divides (a-c) and (b-d). Since a,b,c,d < q, we have a=c and b=d, i.e. a+bi=c+di. Therefore there are precisely q^2 elements. This order counting holds when q is an arbitrary integer.

Once again let \overline{a+bi} be a typical (nonzero) element. Since q is irreducible thus prime thus \mathbb{Z}[i]/(q) is an integral domain, we have \overline{(a+bi)(a-bi)}≠\overline{0}. Letting v^{-1} denote the inverse of \overline{(a+bi)(a-bi)} modulo q, we have \overline{(a+bi)(v^{-1}(a-bi))}=\overline{1}, so that \mathbb{Z}[i]/(q) is a field.

Alternatively, by a previous exercise \mathbb{Z}[i]/I is finite for any nonzero ideal I, so these finite integral domains are clearly fields.

(c) Since (\pi) is prime and thus maximal by a property of PIDs, and since \pi is not associate to \overline{\pi} i.e. \pi r ≠ \overline{\pi} for any r∈\mathbb{Z}[i], we have (\pi,\overline{\pi})=\mathbb{Z}[i] so that the Chinese Remainder Theorem holds by (\pi)(\overline{\pi})=(\pi\overline{\pi})=(p). Parallel to the proof above, \mathbb{Z}[i]/(p) is of order p^2 and since (\pi),(\overline{\pi})≠\mathbb{Z}[i] we have they are both finite integral domains (fields) of order p.

(7) (a) The containment should be evident, so we must demonstrate the group isomorphism by φ(x+(\pi))=x\pi^n+(\pi^{n+1}). Well-definedness:x+(\pi)=y+(\pi)⇒x=y+\pi r⇒\pi r = x-y ⇒ \pi^{n+1} r = x\pi^n-y\pi^n⇒x\pi^n+(\pi^{n+1})=y\pi^n+(\pi^{n+1})⇒φ(x+(\pi))=φ(y+(\pi))Homomorphism:φ(x+(\pi)+y+(\pi))=φ((x+y)+(\pi))=x\pi^n+y\pi^n+(\pi^{n+1})=φ(x+(\pi))+φ(y+(\pi)) Surjectivity: x∈(\pi^n)/(\pi^{n+1})⇒x=\pi^n r + (\pi^{n+1})⇒φ(r+(\pi))=x Injectivity: φ(x+(\pi))=(\pi^{n+1})⇒x\pi^n∈(\pi^{n+1})⇒\pi \mid x⇒x+(\pi)=0+(\pi) (b) Assume \pi is not associate to 1+i. Let N(\pi)=p. We claim \mathbb{Z}[i]/(p^n)≅\mathbb{Z}[i]/(\pi^n) \times \mathbb{Z}[i]/(\overline{\pi^n}). The greatest common divisor of \pi^n and \overline{\pi^n} according to proposition 13 is 1, and since \mathbb{Z}[i] is a Euclidean Domain it can written as a linear combination of \pi^n and \overline{\pi^n}, implying (\pi^n,\overline{\pi^n})=\mathbb{Z}[i]. Since (\pi^n)(\overline{\pi^n})=(p^n), the Chinese Remainder Theorem takes care of the claim.

Now, we have \mathbb{Z}/(p^n) is of order p^{2n}, therefore |\mathbb{Z}[i]/(\pi^n)| \cdot |\mathbb{Z}[i]/(\overline{\pi^n})| = p^{2n} and in particular these former two are orders of p power. We are now in a position to prove our claim by induction; assuming the claim doesn't hold for some particular n, we have either |\mathbb{Z}[i]/(\pi^n)| < p^n or |\mathbb{Z}[i]/(\overline{\pi^n})| < p^n. Without loss of generality we can assume the former, so that |\mathbb{Z}[i]/(\pi^n)|=|\mathbb{Z}[i]/(\pi^m)| for some m < n by induction. Write \mathbb{Z}[i]/(\pi^m)≅(\mathbb{Z}[i]/(\pi^n))/((\pi^m)/(\pi^n)) so that necessarily (\pi^m) \subseteq (\pi^n), a contradiction by \pi^m \not ∈ (\pi^n).

So assume \pi=1+i, which by the associativity of 1+i to 1-i implies this is the last case we have to consider. We see that this is the case when n=2m is even, as \mathbb{Z}[i]/(\pi^{2m})=\mathbb{Z}/((\pi^2)^m)=\mathbb{Z}/(2^m), which has 2^{2m}=2^n=|\mathbb{Z}[i]/(\pi)|^n elements as claimed.

So assume n=2m+1. We first show that |\mathbb{Z}[i]/(\pi^{2m+1})| is of 2 power; by way of contradiction, assume there is some prime p dividing the order of this quotient. Since the quotient may be considered a group under addition, where exponentiation of addition is multiplication, by Cauchy's Theorem we have some \overline{a+bi}∈\mathbb{Z}[i]/(\pi^{2m+1}) such that \overline{p} is the smallest positive integer for which \overline{p(a+bi)}=\overline{0}. In other words, \pi^{2m+1} doesn't divide a+bi so that there are less than 2m+1 factors of \pi in the decomposition of a+bi, yet since \pi \not \mid p for any p≠2 we must also have \pi^{2m+1} \not \mid p(a+bi), a contradiction by \overline{p(a+bi)}=\overline{0}.

We have (\pi^{2m+2}) \subset (\pi^{2m+1}) \subset (\pi^{2m}), so that |(\mathbb{Z}[i]/(\pi^{2m+2}))/((\pi^{2m})/(\pi^{2m+2}))| <|(\mathbb{Z}[i]/(\pi^{2m+2}))/((\pi^{2m+1})/(\pi^{2m+2}))| <|(\mathbb{Z}[i]/(\pi^{2m+2}))/((\pi^{2m+2})/(\pi^{2m+2}))|implying |\mathbb{Z}[i]/(\pi^{2m})| = 2^{2m} < |\mathbb{Z}[i]/(\pi^{2m+1})| < |\mathbb{Z}[i]/(\pi^{2m+2})| = 2^{2m+2}Since |\mathbb{Z}[i]/(\pi^{2m+1})| is of 2 power, we must have |\mathbb{Z}[i]/(\pi^{2m+1})|=2^{2m+1}=|\mathbb{Z}[i]/(\pi)|^{2m+1}.

(c) Let \alpha = \pi_1^{a_1}...\pi_n^{a_n} be its decomposition, so that |\mathbb{Z}[i]/(\alpha)|=|\mathbb{Z}[i]/((\pi_1^{a_1})...(\pi_n^{a_n}))|, the latter of which we claim is equivalent to |\mathbb{Z}[i]/(\pi_1^{a_1})| ... |\mathbb{Z}[i]/(\pi_n^{a_n})| by the Chinese Remainder Theorem; taking any (\pi_x^{a_x}) and (\pi_y^{a_y}) with x≠y we can see (\pi_x^{a_x},\pi_y^{a_y})=\mathbb{Z}[i] since the greatest common divisor of the pair is 1. The original claim now follows by|\mathbb{Z}[i]/(\alpha)|=|\mathbb{Z}[i]/(\pi_1^{a_1})...(\pi_n^{a_n})|=|\mathbb{Z}[i]/(\pi_1^{a_1})| ... |\mathbb{Z}[i]/(\pi_n^{a_n})| =|\mathbb{Z}[i]/(\pi_1)|^{a_1} ... |\mathbb{Z}[i]/(\pi_n)|^{a_n}=N(\pi_1)^{a_1} ... N(\pi_n)^{a_n} = N(\alpha)~\square

Wednesday, May 8, 2013

Rings of Fractions of Principle Ideal Domains (8.2.8)

Dummit and Foote Abstract Algebra, section 8.2, exercise 8:

MathJax TeX Test Page Let R be a principle ideal domain, and let D be a multiplicatively closed subset of R. Prove that D^{-1}R is a principle ideal domain.

Proof: Let I be a typical ideal of D^{-1}R, and J be the subset of elements of R appearing as numerators of elements of I. We claim J is an ideal: For any a,c∈J, we have some \dfrac{a}{b},\dfrac{c}{d}∈I by definition, so \dfrac{a}{b} \cdot \dfrac{b^2}{b} = \dfrac{ae}{e}∈I for any e∈D and similarly \dfrac{ce}{e}∈I. Now (\dfrac{ae}{e}-\dfrac{ce}{e}) \cdot \dfrac{e}{e^2}=\dfrac{(a-c)e}{e} \cdot \dfrac{e}{e^2}=\dfrac{a-c}{e}∈I so that a-c∈J and J is closed under subtraction. For any r∈R, we have \dfrac{re}{e} \cdot \dfrac{a}{b}=\dfrac{ra}{b}∈I so that ra∈J and J is an ideal. Due to the hypothesis, J=(\alpha) for some \alpha∈R.

We now claim I=(\dfrac{\alpha e}{e}). Since \alpha∈J, as before we have \dfrac{\alpha e}{e}∈I so that (\dfrac{\alpha e}{e}) \subseteq I. Now, observe arbitrary \dfrac{x}{y}∈I, and write x=z\alpha. We have \dfrac{\alpha e}{e} \cdot \dfrac{z}{b} = \dfrac{a}{b}, so that I \subseteq (\dfrac{\alpha e}{e}) and now I=(\dfrac{\alpha e}{e}) is a principle ideal, and D^{-1}R is a principle ideal domain.~\square

Sunday, May 5, 2013

(p-1)th Roots of the p-Adic Integers (7.6.11e)

Dummit and Foote Abstract Algebra, section 7.6, exercise 11e:

MathJax TeX Test Page Show that if a_1 \not ≡ 0 \mod{p}, then there is an element a=(a_i) in the inverse limit \mathbb{Z}_p satisfying a_j^{p-1} ≡ 1 \mod{p^j} and \mu_{j1}(a_j)=a_1 for all j. Deduce that \mathbb{Z}_p contains p-1 distinct (p-1)th roots of 1.

Proof: We claim that for any a_1 \not ≡ 0 \mod{p}, the coordinates of a are uniquely determined and satisfy the given criteria. We proceed by induction on the nth coordinate: The case holds true for n=1, so observe the n+1th coordinate. a_{n+1}^{p-1} ≡ 1 \mod{p^{n+1}} if and only if its associated polynomial representation (b_0+b_1p+...+b_np^n)^{p-1} ≡ 1 \mod{p^{n+1}} holds true, so this is equivalent to solving for b_n since a_n=b_0+b_1p+...+b_{n-1}p^{n-1} has been uniquely determined.(b_0+b_1p+...+b_np^n)^{p-1} ≡ 1 \mod{p^{n+1}}⇔((b_0+b_1p+...+b_{n-1}p^{n-1})+(b_np^n))^{p-1} ≡ 1 \mod{p^{n+1}}⇔\sum_{k=0}^{p-1} {p-1\choose k}(b_0+b_1p+...+b_{n-1}p^{n-1})^k(b_np^{n})^{p-1-k}≡1 \mod{p^{n+1}}⇔(b_0+b_1p+...+b_{n-1}p^{n-1})^{p-1}+(p-1)(b_0+b_1p+...+b_{n-1}p^{n-1})^{p-2}(b_np^n) ≡ 1 \mod{p^{n+1}}Since (b_0+b_1p+...+b_{n-1}p^{n-1})^{p-1}≡ 1 \mod{p^n} has been uniquely determined, we have (b_0+b_1p+...+b_{n-1}p^{n-1})^{p-1}=1+vp^n for some integer v.1+vp^n+(p-1)(b_0+b_1p+...+b_{n-1}p^{n-1})^{p-2}(b_np^n) ≡ 1 \mod{p^{n+1}}⇔1+vp^n-b_0^{p-2}b_np^n ≡ 1 \mod{p^{n+1}}⇔1+(v-a_1^{p-2}b_n)p^n ≡ 1 \mod{p^{n+1}}⇔v-a_1^{p-2}b_n ≡ 0 \mod{p}⇔b_n≡a_1^{2-p}v \mod{p}~\square

Direct Limits (7.6.8)

Dummit and Foote Abstract Algebra, section 7.6, exercise 8-9:

MathJax TeX Test Page 8. Let I be an indexing set with partial order , with A_i an additive abelian group for all i∈I. Furthermore, I is a directed set: For any i,j∈I, there exists k∈I such that i,j≤k. Suppose that for every pair of indices i,j with i≤j there exists a map \rho_{ij} : A_i → A_j such that \rho_{jk} \circ \rho_{ij} for all i≤j≤k and \rho_{ii} = 1. Letting B be the disjoint union of all the A_i, define a relation \sim on B bya \sim b ⇔ ∃k∈I (i,j≤k \land \rho_{ik}(a)=\rho_{jk}(b))for a∈A_i and b∈A_j.
(a) Show that \sim is an equivalence relation on B. The set of equivalence classes is called the direct limit of the direct system \{A_i\} and is denoted \varinjlim A_i. For the rest of the exercise let A = \varinjlim A_i.
(b) Let \overline{x} denote the class of x and define \rho_i : A_i → A by \rho_i(a)=\overline{a}. Show that if \rho_{ij} is injective for all i,j then \rho_i is injective for all i (so that each A_i can be viewed as a subset of A).
(c) Assume \rho_{ij} are all group homomorphisms. For a∈A_i,b∈A_j show that the operation\overline{a}+\overline{b}=\overline{\rho_{ik}(a)+\rho_{jk}(b)} is well defined, for any k such that i,j≤k. Show that this makes A into an abelian group, and that \rho_i in part (b) are group homomorphisms.
(d) Show that if A_i are commutative rings with 1 and \rho_{ij} are ring homomorphisms that send 1 to 1, then A may likewise be given the structure of a commutative ring with 1 such that \rho_i are all ring homomorphisms.
(e) Under the hypotheses in (c), prove that the direct limit has the following universal property: if C is any abelian group such that for each i∈I there is a homomorphism φ_i : A_i → C with φ_i = φ_j \circ \rho_{ij} whenever i≤j, then there is a unique homomorphism φ : A → C such that φ \circ \rho_i = φ_i for all i.

Proof: (a) Reflexivity: \rho_{ii}(a)=a=\rho_{ii}(a)⇒a \sim a. Symmetry: a \sim b⇒ \rho_{ik}(a)=\rho_{jk}(b) ⇒ \rho_{jk}(b)=\rho_{ik}(a) ⇒ b \sim a for some k. Transitivity: Let c∈A_l. a \sim b \land b \sim c ⇒ \rho_{ik}(a)=\rho_{jk}(b) \land \rho_{jm}(b)=\rho_{lm}(c) for some k and m with i,j≤k and j,l≤m. Let k,m≤n. We have \rho_{in}(a)=\rho_{kn} \circ \rho_{ik}(a)=\rho_{kn} \circ \rho_{jk}(b)=\rho_{jn}(b)=\rho_{mn} \circ \rho_{jm}(b)=\rho_{mn} \circ \rho_{lm}(c)=\rho_{ln}(c) so that a \sim c.
(b) Assume a,b∈A_i and \rho_{i}(a)=\rho_{i}(b), so that \overline{a}=\overline{b} ⇒ a \sim b. Then \rho_{ij}(a)=\rho_{ij}(b) for some j, so that a=b.
(c) Start with well-definedness. Let \overline{a'}=\overline{a} by \rho_{i'n}(a')=\rho_{in}(a) for some i,i'≤n, and denote a'∈A_{i'}, and likewise for b' with m in place of n. It suffices to show \rho_{i'k'}(a')+\rho_{j'k'}(b') \sim \rho_{ik}(a) + \rho_{jk}(b) for arbitrary applicable k'. Choose k^o such that n,k,k',m≤k^o (by choosing a value greater than or equal to n and k, and another greater than or equal to m and k', and letting k^o be a value greater than or equal to both). We have\rho_{k'k^o}(\rho_{i'k'}(a')+\rho_{j'k'}(b'))=\rho_{k'k^o} \circ \rho_{i'k'}(a')+\rho_{k'k^o} \circ \rho_{j'k'}(b')=\rho_{i'k^o}(a')+\rho_{j'k^o}(b')=\rho_{nk^o} \circ \rho_{i'n}(a')+\rho_{mk^o} \circ \rho_{j'm}(b')=\rho_{nk^o} \circ \rho_{in}(a) + \rho_{mk^o} \circ \rho_{jm}(b)=\rho_{ik^o}(a)+\rho_{jk^o}(b)=\rho_{kk^o} \circ \rho_{ik}(a)+\rho_{kk^o} \circ \rho_{jk}(b)=\rho_{kk^o}(\rho_{ik}(a)+\rho_{jk}(b))Due to the homomorpicity of \rho_{ij}, we have \overline{e}=0 in A for any e∈A_i a zero and -\overline{a}=\overline{-a}. Associativity, closure, and abelianness are assured by their demonstration in A_i. Furthermore, for a,b∈A_i, we have \rho_{i}(a)+\rho_{i}(b)=\overline{a}+\overline{b}=\overline{\rho_{ii}(a)+\rho_{ii}(b)}=\rho_{i}(a+b).
(d) This is simply part (c) with the operation of addition replaced by multiplication, together with the evident fact that \overline{e}=1 in A for any e∈A_i an identity.
(e) Any such homomorphism φ is uniquely defined by φ(\overline{a})=φ \circ \rho_{i}(a)=φ_i(a), so now well definedness and homomorphicity must be demonstrated: φ(\overline{a'})=φ_{i'}(a)=φ_n \circ \rho_{i'n}(a')=φ_n \circ \rho_{in}(a)=φ_i(a)=φ(\overline{a}), and φ(\overline{a}+\overline{b})=φ(\overline{\rho_{ik}(a)+\rho_{jk}(b)})=φ_k(\rho_{ik}(a)+\rho_{jk}(b))=φ_i(a)+φ_j(b)=φ(\overline{a})+φ(\overline{b})where addition is interchangeable with multiplication.~\square

Saturday, May 4, 2013

Collapsing Units Between Modular Rings (7.6.7)

Dummit and Foote Abstract Algebra, section 7.6, exercise 7:

MathJax TeX Test Page Let n \mid m, and prove that the natural projection map (\mathbb{Z}/m\mathbb{Z})^{\times} → (\mathbb{Z}/n\mathbb{Z})^{\times} is surjective.

Proof: Firstly, if u is a unit of \mathbb{Z}/m\mathbb{Z}, then (u,m)=1, so that (u,n)=1, so that u is a unit of \mathbb{Z}/n\mathbb{Z}, proving the mapping is as claimed. To show surjectivity, let p_i be a prime not dividing n but dividing m for all applicable i. For a typical unit x∈(\mathbb{Z}/n\mathbb{Z})^{\times}, solve the simultaneous system of equationsy ≡ x \mod{n}y ≡ 1 \mod{\prod p_i}which is always possible given that n and \prod p_i are coprime. For any prime q dividing m, we have q divides either n or \prod p_i. In the former case, since y is coprime to n, we have y coprime to q. In the latter case, since y is coprime to \prod p_i, we have y coprime to q. Since y is coprime to all the prime divisors of m, it is coprime to m itself, and is thus a unit of (\mathbb{Z}/m\mathbb{Z})^{\times} which naturally projects to x in (\mathbb{Z}/n\mathbb{Z})^{\times}.~\square

Friday, May 3, 2013

Classification of Finite Boolean Rings with Identity (7.6.1-2)

Dummit and Foote Abstract Algebra, section 7.6, exercises 1-2:

MathJax TeX Test Page 1. An element e∈R is idempotent if e^2=e. Assume e is idempotent and within the center of R. Prove Re and R(1-e) are two-sided ideals of R and R≅Re \times R(1-e). Show e and 1-e are identities for Re and R(1-e) respectively.
2. Let R be a finite Boolean ring with identity. Prove R≅\mathbb{Z}/2\mathbb{Z} \times ... \times \mathbb{Z}/2\mathbb{Z}.

Proof: (1) Re and R(1-e) are already left ideals. For any r_1e∈Re and r∈R, we have r_1er=(r_1r)e∈Re so that Re is a two-sided ideal. As well, we have (1-e)^2=1-2e+e^2=1-2e+e=1-e and r(1-e)=r-re=r-er=(1-e)r so that (1-e) is idempotent and central and similarly R(1-e) is a two-sided ideal.

Define a mapping φ : R → Re \times R(1-e) by r \mapsto (re,r(1-e)). This is clearly a homomorphism of rings, and we can prove surjectivity by observing φ(e)=(e^2,e(1-e))=(e,0) and φ(1-e)=((1-e)e,(1-e)^2)=(0,1-e) and now for any (r_1e,r_2(1-e)∈Re \times R(1-e) we have φ(r_1e+r_2(1-e))=φ(r_1)(e,0)+φ(r_2)(0,1-e)=(r_1e,r_2(1-e)). Finally, this mapping is injective since φ(r)=(0,0)⇒re=0 \land r(1-e)=0 ⇒r-re=r-0=0 ⇒ r=0 As we've already seen, e and (1-e) are central and idempotent, so that their identity qualities are deduced straightforwardly.~\square

(2) Lemma 1: Let φ_i : A_i → B_i be an isomorphism of rings for all applicable i. We haveA_1 \times A_2 \times ... ≅ B_1 \times B_2 \times ...Proof: Define a function that maps the ith coordinate of A_1 \times A_2 \times ... to the ith coordinate of B_1 \times B_2 \times ... as its image under φ_i. Due to the nature of the isomorphisms φ_i, we have this map is an isomorphism.~\square

Definition 1: Let R be a ring with r∈R. Define the rank of R from r as the order of the smallest subset A \subseteq R such that r∈A and A generates R.

We will proceed by induction on the rank of R from 1, where R is a finite Boolean ring with identity. When the rank from 1 is 1, since R is of characteristic 2 we have R ≅ \mathbb{Z}/2\mathbb{Z}. Assume that for some integer n we can deduce that a finite Boolean ring with rank from 1 of m≤n is necessarily isomorphic to a direct product of copies of \mathbb{Z}/2\mathbb{Z}, and now observe R with rank from 1 of n+1. Let A=\{1,a,r_3,...,r_{n+1}\} be its fulfilling generating subset, with 1≠a. Since a is necessarily central and idempotent, we have R ≅ Ra \times R(1-a) by the first exercise.

We claim that \{1 \cdot a,a \cdot a,r_3 \cdot a,...,r_{n+1} \cdot a\}=\{a,r_3a,...,r_{n+1}a\} is a generating subset of Ra containing the identity and of order ≤n so that Ra has a rank from a (the identity of Ra) of ≤n and so is isomorphic to some number of direct products of \mathbb{Z}/2\mathbb{Z} by induction. This is immediate because any series of operations taking elements from this subset is divisible by a, and for any ra∈Ra we have ra=Na for some series of operations N from A and evidently Na is replicable by a series of operations from the subset. Similarly the case holds for R(1-a) where a potential generating subset containing its identity is \{1 \cdot (1-a),a \cdot (1-a),r_3 \cdot (1-a),...,r_{n+1} \cdot (1-a)\}=\{1-a,0,r_3-r_3a,...,r_{n+1}-r_{n+1}a\}. We may remove the superfluous zero from this generating subset. Therefore by induction and the lemmaR≅Ra \times R(1-a) ≅(\mathbb{Z}/2\mathbb{Z} \times ... \times \mathbb{Z}/2\mathbb{Z}) \times (\mathbb{Z}/2\mathbb{Z} \times ... \times \mathbb{Z}/2\mathbb{Z}) ≅\mathbb{Z}/2\mathbb{Z} \times ... \times \mathbb{Z}/2\mathbb{Z}~\square In addition, this provides a unique construction of the finite Boolean ring R with identity of rank n for any n∈\mathbb{Z}^+.

Thursday, May 2, 2013

Zorn's Lemma on the Real Numbers (7.5.6)

Dummit and Foote Abstract Algebra, section 7.5, exercise 6:

MathJax TeX Test Page Prove that \mathbb{R} contains a subring A with identity and maximal under inclusion such that \dfrac{1}{2} \not ∈ A.

Proof: Let \mathcal{F} be the set of all subrings of \mathbb{R} not containing \dfrac{1}{2}, and let \mathcal{C} be a typical chain of subrings R_0 \subseteq R_1 \subseteq .... AdmitR=\bigcup_{n∈N}R_nand prove it is an upper bound of \mathcal{C}. For any a,b∈R, we have a∈R_x and b∈R_y for some x,y implying a-b∈R and ab∈R so that R is a subring. Furthermore, by definition of the union, 1∈R and \dfrac{1}{2} \not ∈ R. By Zorn's Lemma, \mathcal{F} has a maximal element.~\square