14
$\begingroup$

I apologize in advance if this is well-known, but I can't seem to find the answer in the literature. Let me be precise about my question. I am looking for concrete examples of locally compact Hausdorff groups $G$ such that that there exists $a, b \in G$ with $a \ne b$, but for any continuous representation $\pi : G \to \operatorname{GL}_n(k)$ with $k$ a valuated field and $n$ a positive integer, we have $\pi(a) = \pi(b)$.

I am also interested in a weaker form: the case where $k$ is the field $\mathbb{C}$ of complex numbers.

If it turns out that there are no such groups, please indicate a reference or sketch a proof.

Also a relevant statement which my gut tells me should be true, but I don't know enough Lie theory yet to give a complete proof: for every real Lie group $G$, there are enough finite dimensional, continuous complex representations of $G$ to distinguish points of $G$. I was thinking maybe this can be proved by exploiting the close relationship between representations of a Lie group and representations of its Lie algebra. It would be nice if someone could give a sketch if this can indeed be achieved or describe why this "hand-waving" actually does not work.

Edit: Sorry for the poor choice of terminology. To truly get the "weaker form" as YCor pointed out, I should replace "valuated field" by "a field $k$ with an absolute value" $|\cdot|: k \to \mathbb{R}_{\ge 0}$, such that (1) $|ab| = |a| |b|$; (2) $|a| = 0$ if and only if $a=0$; (3) $|1 + a| \le 2$ for all $a$ with $|a| \le 1$ (or equivalently by a well-known argument, the triangle inequality). To exclude the discrete topology induced by this absolute value, we exclude the trivial absolute value that $|a| = 1$ for all $a \in k^\times$.

$\endgroup$
7
  • 5
    $\begingroup$ I didn't check carefully but I suspect Higman's group will serve as a (discrete) counterexample in the $k={\bf C}$ case. en.wikipedia.org/wiki/Higman_group $\endgroup$
    – Terry Tao
    Dec 7, 2021 at 1:55
  • 8
    $\begingroup$ $SL_2\mathbb C$ is simply connected, but $SL_2\mathbb R$ has infinite fundamental group. Its double cover is a counterexample. It is connected, so it doesn't have any $p$-adic representations. Its finite dimensional real reprentations yield representations of its Lie algebra, which complexify to complex representations of $SL_2\mathbb C$, which doesn't have a double cover so can't detect the kernel of the extension. $\endgroup$ Dec 7, 2021 at 1:57
  • 1
    $\begingroup$ The object to understand, for a group $G$, is the intersection $K_G$ of all kernels of all finite-dimensional continuous representations over all valued fields (and similarly $K'_G$ defined considering finite-dimensional representations). You're asking when $K_G=\{1\}$ or $K'_G=\{1\}$. For instance for compact groups $G$, Peter-Weyl ensures $K'_G=\{1\}$. On the other hand, if $G$ is a finitely generated group then its "finite residual" (the intersection of all its finite index subgroups) is contained in $K_G\cap K'_G$. In particular if $G$ has no nontrivial finite quotient, $K_G=K'_G=G$. $\endgroup$
    – YCor
    Dec 7, 2021 at 8:28
  • $\begingroup$ As regards the difference between $K_G$ and $K'_G$: it seems to me that $K_G=G$ for every connected group $G$, since every valued field is totally disconnected (if I understand what you mean by valued field). While $K_G=\{1\}$ for many connected Lie groups. At the opposite, for $\mathrm{SL}_n(\mathbf{Q}_p)$ we have $K_G=G$ and $K'_G=\{1\}$. $\endgroup$
    – YCor
    Dec 7, 2021 at 8:32
  • 1
    $\begingroup$ You've ignored my previous comment while editing: I explained that the "weaker form" is not weaker. Unless you mean by "valued field" something other that what I imagine (which is in particular totally disconnected). $\endgroup$
    – YCor
    Dec 7, 2021 at 17:26

3 Answers 3

12
$\begingroup$

There is an example which satisfies something much stronger: there exist nontrivial groups $G$ such that any homomorphism (not even necessarily continuous) $\pi:G\to GL_n(k)$ for any field $k$ (and, in fact, any commutative integral domain) is trivial, so in particular for any $a,b\in G$ we have $\pi(a)=\pi(b)$. Since any group can be given a locally compact Hausdorff topology, namely the discrete topology, these will in particular answer your question.

As for examples of such groups $G$, we can take any finitely generated which has no finite quotients, e.g. the Higman group mentioned in the comment by Terry Tao.

This result is apparently due to Mal'cev, but I don't have a reference at hand so here is a sketch of an argument. It is enough to show that if you have a finitely generated subgroup $\Gamma$ of a group $GL_n(R)$ for some integral domain $R$ (e.g. the image of $G$ under a representation), then $\Gamma$ is residually finite (hence trivial, if $\Gamma$ is a quotient of $G$ which has no finite quotients). The main idea of the proof is to replace $R$ by some ring which is finitely generated over $\mathbb Z$, meaning it is a quotient of a polynomial algebra over $\mathbb Z$. Then $R$ is a Jacobson ring, so in particular the intersection of all maximal ideals of $R$ is trivial, and moreover for all maximal ideals $m$ of $R$ we have $R/m$ finite. Now if $\gamma\in\Gamma$ is any nontrivial element, then there is a maximal ideal $m$ not containing all coefficients of $\gamma$, so $\gamma$ has nontrivial image in the finite group $GL_n(R/m)$.

$\endgroup$
4
  • $\begingroup$ I marked this as accepted answer though I couldn't completely check all the details. It would be more helpful if you could provide a reference when you have the time. $\endgroup$ Dec 10, 2021 at 13:36
  • 1
    $\begingroup$ @RickSternbach You can find the details for instance in section 10 of these notes $\endgroup$
    – Wojowu
    Dec 10, 2021 at 13:46
  • 2
    $\begingroup$ In fact, there are groups which have no non-trivial homomorphism into $\text{G}(R)$ for any commutative unital ring $R$. For this, take an infinite simple group with property (T), see appendix A in arxiv.org/abs/1608.06265. There exists finitely presented such groups by the work of Caprace-Remy, see arxiv.org/abs/math/0607664. $\endgroup$
    – Uri Bader
    Dec 12, 2021 at 8:57
  • $\begingroup$ @UriBader Thank you for the references! I believed this more general statement to be the case too, but I couldn't find it in the literature. $\endgroup$
    – Wojowu
    Dec 12, 2021 at 10:08
9
$\begingroup$

If you take an infinite simple group $G$ (say Thompson’s $T$) and put the discrete topology on it this will have your property when k is a finite field (since the image in $GL_n(k)$ will necessarily be trivial.

Say further that $G$ is finitely presented (as indeed for example Thompson's $T$ is). Then you can generalize this to $\mathbb{C}$ as follows by following the techniques in this paper. Basically, a non-trivial homomorphism into $GL_n(\mathbb{C})$ can be turned into a homomorphism into $GL_n(\mathbb{F}_p)$ for some prime $p$. See theorem 3.4. There they use it to turn a $\mathbb{C}$ representation with non-commutative image into a $\mathbb{F}_p$ representation with non-commutative image, but you can do the same thing for "non-trivial" instead of "non-commutative" $\mathbb{C}$ representation. I will spell it out below.

The key result it relies on is a very cool and useful theorem that says "solutions to systems of equations over $\mathbb{C}$ imply over some finite field." Specifically, if $f_1(x_1, \dots, x_N), \dots f_m(x_1, \dots, x_N)$ is a system of polynomials with integer coefficients that has a solution over $\mathbb{C}$, then it also has a solution over $\mathbb{F}_p$ for some prime $p$. Edit: This was using a theorem cited in the attached paper. I replaced this with an easy to prove theorem that gives you a solution over a finite field (not necessarily prime order), proven below.

Next we want to construct a system of polynomials such that a solution to that system in a field $k$ will correspond to a $k$-representation.

This we do as follows. Since our group $G$ is finitely presented (say with generators $a_1, \dots, a_s$ and relations $w_1, \dots, w_r$), for any $n$, we can write down matrices $A_1, \dots, A_s, B_1, \dots, B_s$ with variable entries (like $a_{i, j}$ and $b_{i, j}$) and consider the products of these matrices corresponding to the relations $w_i$, where we replace $a_i$ with $A_i$ and $a_i^{-1}$ with $B_i$. By equating the resulting polynomial-entried matrix with the identity matrix, we get $n^2$ integer-coefficient polynomial equations.

We then do this for all matrices simultaneously, plus the additional relations $A_i B_i - I_n = 0$, to get a big system of integer-coefficient polynomials for which a solution in $k$ is exactly the data of a $k$ representation of $G$.

Now, let's say we have a non-trivial $\pi \colon G \to GL_n(\mathbb{C})$. If it has a non-trivial image, there is some $x \in G$ with $\pi(x) \neq 1$. Write $x$ as a word $w$ in the alphabet $a_1^\pm, \dots a_s^\pm$. Denote by $X$ the product of $A_i, B_i$s corresponding to the word $w$ (and thus to the element $x \in G$). We want to augment the above system of equations to say that the image of $w$ in $GL_n(\mathbb{C})$ is not the identity matrix.

This is just a bit of hacking -- we want to enforce that for some $i, j$ the $i,j$ entry of $I_n$ differs from $X$. I.e., some $\delta_{i, j} - X_{i,j}$ is non-zero. For each $i, j$ introduce new variables $z_{i, j}$ and $r_{i, j}$ and add equations

  • $z_{i, j} (\delta_{i, j} - X_{i, j}) - (1 - r_{i, j})$
  • $z_{i, j} r_{i, j}$

You can check that these force $r_{i, j}$ to be 0 if $\delta_{i, j} - X_{i, j}$ is non-zero, and 1 if $\delta_{i, j} - X_{i, j}$ is 0.

Now we want to say that some $r_{i, j}$ is 0. This we can do by adding $\prod_{i, j} r_{i, j}$ to our system of equations.

At this point, a solution to the system over $k$ is exactly a $k$ representation that sends $x$ to a non-trivial element of $GL_n(k)$. We assumed such a representation exists for $\mathbb{C}$. Thus, the lemma proved below yields a representation for some finite field. But this is a contradiction, as we showed earlier that there are no representations over finite fields. Thus, there cannot be representations over $\mathbb{C}$ either.

Edit: After some reflection I realized we can do without the results of the paper in this context by proving an easier version of a similar theorem.

Lemma. Fix any prime $p$. If $f_1(x_1, \dots, x_N), \dots f_m(x_1, \dots, x_N) \in \mathbb{Z}[x_1, \dots, x_N]$ have a common zero over $\mathbb{C}$, then they have a common zero over some finite field of characteristic $p$.

Proof.

Consider the ideal $I_p = (f_1, \dots, f_m) \subseteq \mathbb{F}_p[x_1, \dots, x_N]$.

  1. First suppose this ideal is trivial (i.e., equal to $\mathbb{F}_p[x_1, \dots, x_N]$). Then there exist some $g_1, \dots, g_m \in \mathbb{F}_p[x_1, \dots, x_N]$ with $$ g_1 f_1 + \dots + g_m f_m = 1 \mod p $$ Now let $G_i \in \mathbb{Z}[x_1, \dots, x_N]$ be any polynomial which when reduced mod $p$ is $g_i$. Then $$ G_1 f_1 + \dots + G_m f_m = 1 + a p $$ for some $a \in \mathbb{Z}$. Then we have $$ \frac{G_1}{1 + ap} f_1 + \dots + \frac{G_m}{1+ ap} f_m = 1 $$ which means the ideal $(f_1, \dots, f_m) \subseteq \mathbb{C}[x_1, \dots, x_N]$ is trivial. But that means there can be no solution to the $f_i$ over $\mathbb{C}$, which is a contradiction to our assumption.

  2. So, it must be the case that $I_p$ is not trivial. Then, by the weak nullstellensatz, the $f_i$ have a common solution $\alpha_1, \dots, \alpha_m$ in $\overline{\mathbb{F}_p}$, the algebraic closure of $\mathbb{F}_p$. Since each of the $\alpha_i$ is contained is some finite extension of $\mathbb{F}_p$, we can take a finite extension $k$ of $\mathbb{F}_p$ that contains all the $\alpha_i$. So we are done: $k$ is our finite field of characteristic $p$ which has a solution $(\alpha_1, \dots, \alpha_m)$ to our system of equations.

$\endgroup$
6
  • $\begingroup$ "infinite simple" should be "finitely generated infinite simple". Otherwise $\mathrm{SL}_3(\mathbf{Q})$ is a counterexample to the first assertion. $\endgroup$
    – YCor
    Dec 7, 2021 at 8:29
  • $\begingroup$ By the first assertion do you mean “an infinite simple group G has no non-trivial representations in finite fields”? This is true because the image is a finite quotient of G of which there is only the trivial group. I further imposed the group should be finitely presented in the second paragraph $\endgroup$ Dec 7, 2021 at 19:06
  • 2
    $\begingroup$ Sorry, indeed I didn't see you assumed to be finite and thought you were implicitly using Malcev's theorem that f.g. subgroups of $GL_n$ over any field, are residually finite. So my point is that a finitely generated group with no nontrivial finite quotient has no nontrivial (finite-dim) linear representation over any field. By the way, Malcev's result is classical and finite generation is enough (no finite presentability is needed). $\endgroup$
    – YCor
    Dec 8, 2021 at 8:28
  • 1
    $\begingroup$ (...) Malcev's result is proved by showing that every f.g. domain is residually a finite field. Once this is known, it follows that every f.g. group having a non-trivial $n$-dimensional representation over some field, also has a non-trivial $n$-dimensional representation over some finite field. $\endgroup$
    – YCor
    Dec 8, 2021 at 8:31
  • 1
    $\begingroup$ Just pointing out that you mean Thompson's group $T$ or $V$, not $F$ ($F$ is not simple). $\endgroup$ Dec 12, 2021 at 0:55
1
$\begingroup$

Towards the weaker form of the question, i.e., when $k = \mathbb{C}$, one can also give the following example.

First, recall the following fact. Any finite dimensional complex representation of a totally disconnected locally compact group, which is continuous with respect to the complex topology factors through a discrete quotient, see the Uri Bader's answer to this question. In particular it will be smooth (i.e., the stabilizer of any vector will be open).

Example. Now, let $p$ be a prime. The $p$-adic group $G = {\rm GL}_2(\mathbb{Q}_p)$ is locally profinite. By the above fact, any finite-dimensional continuous complex representation of $G$ will be smooth. Now, it is a well-known fact that only finite-dimensional smooth representations of $G$ are one-dimensional. (E.g., one checks that the kernel of any such representation contains all upper and lower triangular matrices, hence contains ${\rm SL}_2(\mathbb{Q}_p)$.) Then, for any $a,b \in {\rm SL}_2(\mathbb{Q}_p)$ one has $\pi(a) = \pi(b)$ for all $\pi$ as above.

Remark. The fact above breaks down without the finite-dimensionality assumption: see, for example, this answer.

$\endgroup$
1
  • 1
    $\begingroup$ I don't think it's "weaker", it's a variant of the question. That continuous complex representations, vs continuous representation over valued fields, separate points, are independent conditions, as I explained in a comment (where I explicitly mentioned that $\mathrm{SL}_2(\mathbf{Q}_p)$ has no nontrivial continuous complex representation). $\endgroup$
    – YCor
    Dec 7, 2021 at 12:26

Your Answer

By clicking “Post Your Answer”, you agree to our terms of service and acknowledge you have read our privacy policy.

Not the answer you're looking for? Browse other questions tagged or ask your own question.