Parameterizing rational points on a singular projective cubic

Let C = Z(F) be a singular rational projective cubic. It follows from Bézout’s theorem (applied over \overline{\mathbb{Q}}) that there can only be one singularity; if not, the line joining two singularities would intersect C with total multiplicity at least 4, but a cubic and a line intersect with multiplicity at most 3. Moreover, another application of Bézout’s theorem proves that the singularity must be a double point; if it were higher, once again we’d arrive at a contradiction after joining the singularity and another point of the curve with a line.

Let’s call the singularity p. We know that p‘s coordinates are algebraic over \mathbb{Q}, so we can pick a finite Galois extension k/\mathbb{Q} containing all of them. Now, we consider the action of the Galois group G=\mathrm{Gal}(k/\mathbb{Q}) over C. Since the Galois group acts on curves, it must preserve its singularities (since a singularity is a common zero of both F and its derivatives). Therefore, it must fix p, and so we conclude that p is a rational point.

Now, any non-tangent line at p with rational slope must intersect the cubic exactly one more time (once again by Bézout). Thus, we completely parameterize the cubic’s rational points in terms of its singularity.

Posted in Uncategorized | Tagged , | Leave a comment

A non-free, torsion-free abelian group

In this post we’ll prove that M:=\prod_{n} \mathbb{Z} is not free as a \mathbb{Z}-module, although it is torsion-free. Recall that, in the finitely generated case, torsion-free modules are free, by the structure theorem.

For the sake of contradiction, assume that M is free. Let N be the submodule of M consisting of sequences (x_n) such that the sequence (p^{k_n}) goes to infinity, where p is a fixed prime and k_n is the biggest natural such that p^{k_n} divides x_n. Since \mathbb{Z} is a PID, every submodule of a free module is free, and so N is free. Moreover, we have an injection f:M\to N defined by f((x_n)) = (p^n x_n). Since M is uncountably generated, it follows that N is, too.

Now we consider the \mathbb{Z}/p\mathbb{Z}-vector space N/pN. Since we supposed N was an uncountably generated free \mathbb{Z}-module, N/pN should be an uncountably generated vector space (since N\simeq \bigoplus_{I} \mathbb{Z} implies N/pN \simeq \bigoplus_{I} \mathbb{Z}/p\mathbb{Z}). On the other hand, every element in N/pN can be represented as a sequence that is eventually zero, and there are a countable number of such sequences. So N/pN is countable; in particular, it’s countably generated, which contradicts our previous observation. We then conclude that M is not free, as we wanted to prove.

Notice that this proves that a direct product of projective modules need not be projective as well, since any projective module over a PID is free.

Posted in Uncategorized | Tagged | Leave a comment

Finite-dimensional subspaces of normed spaces are complemented

Let S be a finite-dimensional subspace of a normed space E and pick a basis \{v_1,\dots,v_n\} for S. Let \varphi_i:E\to k be a bounded linear functional that extends the map v_j\mapsto 0 if j\neq i and v_i\mapsto 1. Such a functional exists by the Hahn-Banach theorem; even more, we can pick one such that its norm is bounded by 1. We now define \psi:E\to E as

\psi(v) =\sum_{i=1}^n \varphi_i(v) v_i

It’s easy to see that E is bounded; in fact \Vert \psi\Vert \leq \sum_{i=1}^n \Vert v_i\Vert, since each \varphi_i is bounded by 1. Therefore, the kernel of this map is closed. Finally, we have a direct sum decomposition E = \mathrm{ker}\psi\oplus\mathrm{img}\,\psi, since \psi is a projector. One easily sees that \mathrm{img}\,\psi=S, and so S is complemented.

Posted in Uncategorized | Tagged | Leave a comment

Cyclic modules over local algebras are indecomposable

Let A be a local algebra and let M be a cyclic A-module. Suppose M=V\oplus W, with V,W\neq 0. Let m be a generator for M; then m=v+w with v\in V, w\in W, both non-zero. Since M is cyclic, there exists some a\in A such that am=v. But v=am=av+aw, and since both V and W are A-modules, this implies av=v and aw=0. In particular, a is not invertible. Therefore, we must have that 1-a is invertible, since A is local. But now (1-a)m = m-am = v+w-v=w, and so (1-a)v=0. This is a contradiction, so we conclude that M must be indecomposable.

Posted in Uncategorized | Tagged , | Leave a comment

The representation type of C_p x C_p in the modular case

In this post we’ll prove that G=C_p\times C_p has infinite representation type in the modular case. This illustrates the fact that the representation theory of groups in the modular case is a hard problem; we have infinite representation type even for very simple groups.

The group algebra kG is isomorphic to k[x,y]/\left(x^p-1, y^p-1\right), and since \mathrm{char}\, k = p, we have that x^p-1=(x-1)^p and so kG = k[x,y]/(x^p,y^p). We’ll now exhibit an infinite family of non-isomorphic indecomposable modules for this algebra.

Take V an n-dimensional k-vector space; we’ll consider M=V\oplus V as a kG-module with x and y acting in the following way:

x(v,w)=(0,v)
y(v,w)=(0,g(v))

where g:V\to V is a k-linear morphism such that its matrix is an n\times n Jordan block of eigenvalue 1 in some basis. We obviously have x^p=y^p=0, so this endows M with a kG-module structure.

Let’s now describe the endomorphism algebra of this module. If f\in \mathrm{End}_{kG}(M), then f=\varphi \oplus \psi where \varphi, \psi:V\to V\oplus V are k-linear morphisms. For f to be kG-linear, the following identity must hold:

From this, we deduce that \psi_1=0 and so \psi_2 =\varphi_2. Doing a similar analysis for y, we obtain that g\varphi_1=\varphi_1 g and so f is a kG-linear morphism iff

where \varphi_1 commutes with g. An easy calculation shows that any such \varphi_1 must be a lower triangular matrix constant in each diagonal in the Jordan basis for g. If \varphi_1 is invertible (that is, if its main diagonal in the Jordan basis for g is not identically zero), then f is (one has to check that the obvious k-linear inverse is kG-linear too, but this is easy). If \varphi_1 is not invertible, then it is a stricly lower triangular matrix in some basis, and so it is nilpotent, and therefore f is too. We have then proved that every kG-endomorphism of M is either invertible or nilpotent. This is a sufficient condition for \mathrm{End}_{kG}(M) to be local, and so M is indecomposable.

Since we can produce an indecomposable module for every natural 2n, this gives an infinite number of non-isomorphic indecomposable kG-modules.

Posted in Uncategorized | Tagged | Leave a comment

Representations of p-groups in the modular case

In this post we’ll prove that if G is a p-group and \mathrm{char}\, k =p, then the only simple kG-module is the trivial one.

Let S be a simple kG-module and pick 0\neq x\in S. We now define H as the subgroup of the additive group of S generated by the elements \{gx:g\in G\}. Since we’re working in characteristic p, a subgroup of the additive group of a vector space must be a p-group of exponent p. Notice that H must be finite, since it’s a finitely generated abelian group of finite exponent.

We have an obvious action of G over H, and we can see that orbits of this action must be of cardinality p^k (allowing k to be zero), by the orbit-stabilizer theorem. Since the orbit of zero is obviously trivial, we have that there’s at least another element y with trivial orbit. Thus \langle x\rangle is a kG-submodule of dimension 1, and so \langle x\rangle=k=S by simplicity. This proves our claim.

Notice that there are non-trivial kG-modules; for instance, take any non-zero element of G and let it act over k^p as a Jordan block of size n of eigenvalue 1, and let all the elements h not in \langle g\rangle act trivially. Then this is a non-trivial representation of G. Since a direct sum of trivial representations is itself trivial, we have proved that no non-trivial module can be expressed as a direct sum of simple modules, and so the group algebra kG is not semisimple. This is a partial converse to Maschke’s theorem (we have only proved it for p-groups, although it holds in full generality).

Posted in Uncategorized | Tagged | Leave a comment

A bijective proof of a classical identity involving the totient function

In this post we’ll give a simple bijective proof of the well-known identity \sum_{d\mid n} \varphi(d)=n.

If we write down the expressions \frac 1n,\dots,\frac nn and then put each one of them in their irreducible form, we have that each divisor d of n appears exactly \varphi(d) times as a denominator. Thus, we have n fractions, comprised of \varphi(d) fractions with denominator d for each d\mid n, and so \sum_{d\mid n} \varphi(d)=n, as we wanted.

Posted in Uncategorized | Tagged | Leave a comment

Two sets are equinumerous iff their symmetric groups are isomorphic

Let X and Y be sets; we’ll prove that |X|=|Y| iff S(X)\simeq S(Y), following Andreas Blass’ argument here.

Suppose |X|=|Y| and pick a bijection f:X\to Y. Then \varphi \mapsto f\circ\varphi\circ f^{-1} is an isomorphism between S(X) and S(Y).

Reciprocally, suppose S(X)\simeq S(Y). In the finite case, we can recover the cardinalities of |X| and |Y| easily, so we’ll restrict to the infinite case. We’ll prove that the smallest non-trivial conjugacy class of S(X) has cardinality |X|, and since the cardinality of conjugacy classes is an isomorphism invariant, this will in turn prove that |X|=|Y|. This is easy to check: the set of all transpositions is a conjugacy class and has cardinality |X|\times |X\setminus\{x\}| = |X|. Moreover, if A is a non-trivial conjugacy class, we have that |A| has cardinality at least |X|, since we can produce that many elements conjugating a non-trivial element of A with transpositions. This proves our claim.

Posted in Uncategorized | Tagged | Leave a comment

The irreducible representations of the dihedral group

In this post we’ll characterize all of the irreducible representations of D_n over an algebraically closed field k. We will suppose \mathrm{char}\, k \nmid 2n so that Maschke’s theorem holds and kD_n has finite representation type.

First, let’s recall the following corollary of the Artin-Wedderburn theorem:

Corollary: Let G be a finite group such that the group algebra kG is semisimple. Let \{S_1,\dots,S_r\} be a complete system of representatives for the isomorphism classes of simple kG-modules and let D_i=\mathrm{End}_{kG}(S_i) and n_i = \mathrm{dim}_{D_i} \mathrm{hom}_{kG}(S_i,kG), so that there’s an isomorphism of algebras kG\simeq M_{n_1}(D_1^\mathrm{op})\times\dots\times M_{n_r}(D_r^\mathrm{op}). Then:

(i) \sum_{i=1}^r n_i^2 \mathrm{dim}_k D_i = \vert G\vert (the so-called sum-of-squares formula).
(ii) If \mathrm{dim}_k S_i=1, then n_i=\mathrm{dim}_k D_i=1.
(iii) If k is algebraically closed and n_i=1 then \mathrm{dim}_k S_i=1.
(iv) If k is algebraically closed, the number of simple modules S_i of k-dimension 1 is exactly the same as \vert G/G'\vert.
(v) r\leq\vert \mathrm{cl}(G)\vert (where \mathrm{cl}(G) stands for the set of conjugacy classes of G) and the equality holds if k is algebraically closed.

Let’s begin with our characterization. Let V be an irreducible n-dimensional representation of D_n, and recall that \langle r,t\vert r^n=t^2=1, tr=r^{-1}t\rangle is a presentation for D_n. The equation r^n=1 implies that the matrix representing r (which we’ll simply call r for the remainder of this post) is diagonalizable, since its minimal polynomial divides x^n-1, which has simple roots because of our hypothesis on the characteristic of the field. Moreover, every eigenvalue of r is an n-th root of unity.

Suppose that r=\mathrm{id}; since t has an eigenvector v, then \langle v\rangle is a non-trivial subrepresentation of V and so must be V by irreducibility. In this case, t is a 1\times 1 matrix such that t^2 = 1, so t=\pm 1. Both possibilities can occur, and in fact correspond to the trivial and sign representations, respectively.

Now, if r\neq\mathrm{id}, take v an eigenvector of r with eigenvalue \lambda \neq 1. Then \lambda t(v) = tr(v) = r^{-1}t(v); and so rt(v) = \lambda^{-1}t(v). One now can easily check that \langle v, t(v)\rangle is a subrepresentation, and we see that V has dimension at most 2. Therefore, in this case, the representation is given by

Recall that \lambda was an n-th root of unity. If n is even, then -1 is an n-th root of unity, and so picking \lambda=-1 makes r a scalar matrix. Therefore any eigenvalue of t gives rise to a 1-dimensional subrepresentation. This makes sense, since D_n/D_n' is \mathbb{Z}/2\mathbb{Z} or \mathbb{Z}/2\mathbb{Z}\oplus \mathbb{Z}/2\mathbb{Z}, and so we’ll have a different number of irreducible representations of dimension 1 depending on the parity of n.

In any other case, that is, \lambda\neq \pm 1, this representation is clearly irreducible since r and t share no eigenvalues.

It’s easy to see that if V_\lambda denotes the irreducible representation we just described, then V_\lambda\simeq V_{\lambda^{-1}}. Let’s prove that if \lambda\neq \mu and \lambda\neq\mu^{-1}, then V_\lambda\not\simeq V_\mu. By (i) \sum_{i=1}^r n_i^2 \mathrm{dim}_k D_i = \vert G\vert = 2n; and since D_i\simeq k (a finite dimensional k-algebra which is a domain must be a field; and therefore D_i is a finite extension of k, which is algebraically closed) we have that \sum_{i=1}^r n_i^2 = 1^2 + 1^2 + \sum 2^2 = 2n in the odd case and \sum_{i=1}^r n_i^2 = 1^2 + 1^2 + 1^2 + 1^2 + \sum 2^2 = 2n in the even case. An easy counting argument then shows that our V_\lambda representations are pairwise non-isomorphic, as we wanted.

This completely classifies the representations of D_n, as every representation is a direct sum of the irreducible representations we just classified.

Moreover, we can give explicit submodules of kD_n isomorphic to each simple module. Given any character \chi, the submodule \langle \sum_{g\in D_n}\chi(g^{-1})g\rangle is isomorphic to the simple module of dimension 1 associated with the character \chi. This construction works in general for any representation induced by a character. Now, for the 2-dimensional case, we have that the submodule \langle t\sum_{i=1}^n (\lambda r)^i, \sum_{i=1}^n (\lambda r)^i\rangle is isomorphic to the simple submodule V_\lambda.

Posted in Uncategorized | Tagged | Leave a comment

Dense subsets of a metric space sometimes determine its cardinality

Suppose X is a metric space and D is a dense subset of X such that \vert D\vert> \aleph_0. Then, for every x we can produce a sequence d_n^x of elements in D such that d_n^x\rightarrow x. This in turn defines a mapping f:X\to D^\mathbb{N}, which is injective (notice that we are heavily using the axiom of choice here). Since \vert D\vert > \aleph_0, we have that \vert D^\mathbb{N}\vert = \vert D\vert and so D actually determines the cardinality of X, which obviously is not the case when D is countable.

One thing to keep in mind is that the crucial step in this proof is the characterization of convergence via sequences, which holds for metric spaces, but not in more general contexts.

Posted in Uncategorized | Tagged | 2 Comments