Optimal linear sofic approximations of countable groups
Abstract.
Let be a group. The notion of linear sofic approximations of over an arbitrary field was introduced and systematically studied by Arzhantseva and Păunescu [AP17]. Inspired by one of the results of [AP17], we introduce and study the invariant that captures the quality of linear sofic approximations of over . In this work we show that when has characteristic zero and is linear sofic over , then takes values in the interval and cannot be replaced by any larger value. Further, we show that under the same conditions, when is torsion free. These results answer a question posed by Arzhantseva and Păunescu [AP17] for fields of characteristic zero. One of the new ingredients of our proofs is an effective non-concentration estimates for random walks on finitely generated abelian groups, which may be of independent interest.
1. Introduction
Let be a family of groups, each equipped with a bi-invariant bounded metric. Bi-invariance means that for every , we have the equality
A -approximation of a countable group consists of an increasing sequence of positive integers and a sequence of maps
satisfying the following two properties:
-
(1)
(Asymptotic homomorphism) For all , one has
-
(2)
(Uniform injectivity) There exists such that for all ,
We will then also say that is -approximable by . Perhaps the most prominent and well-studied classes of approximable groups are the sofic and hyperlinear groups, which correspond, respectively, to approximation by the family of symmetric groups equipped with the normalized Hamming distance and the family of unitary groups equipped with normalized Hilbert-Schmidt distance. The class of sofic groups was introduced by Gromov in connection with the so-called Gottschalk surjunctivity conjecture [Gro99], while the terminology is due to Weiss [Wei00]. Hyperlinear groups first appeared in the context of Conne’s embedding conjecture. The term hyperlinear was coined by Rădulescu [R0̆8]. Sofic groups are shown to be hyperlinear [ES05, Theorem 2], It is unknown whether every group is sofic, or even hyperlinear.
The class of linear sofic groups over an arbitrary field was introduced by Arzhantseva and Păunescu [AP17, Definition 4.1 and the paragraph following Definition 4.2] who proved fundamental results about this class of groups. This mode of approximation defining linear sofic groups uses general linear groups (over a general field fixed in the discussion) as target groups while the metric is defined using the normalized rank. In this regards, linear sofic groups provide a hybrid form of approximation
In order to define this metric, let be a field. For matrices , we define
The following definition of linear sofic groups will be more convenient for our purpose. The equivalence of two definitions is proven in [AP17, Proposition 4.4].
Definition 1.1.
Let be a field, a countable group and . We say that is -linear sofic over if for every finite set and every and every , there exists and a map satisfying the following two properties:
-
(AH)
For all , one has .
-
(D)
For all , .
Such a map is called an -map. Roughly speaking, (AH) guarantees that is almost a homomorphism, while (D) shows that distinct elements are separated out. Following [AP17] we say that is linear sofic over if it is -linear sofic for some . It is clear that if , then every -linear sofic group is -linear sofic. For a countable group , we write
Note that whenever is not -linear sofic over for any , we define to be zero.
Remark 1.2.
The notion of metric approximation can also be defined using the notion of metric ultraproducts. This alternative definition allows one to avoid limiting processes that require passing to subsequences, and thereby simplifies certain arguments, see [AP17] for examples. Since this point of view will not provide us with any special advantages, we will not use this definition.
1.1. The amplification argument
Let be as above, and assume that the diameter of with respect to is normalized to be . It is natural to ask whether a -approximable group for some is always -approximable. Elek and Szabó [ES05] proved that this is the case for sofic groups. A similar statement (with a modified proof) holds for hyperlinear groups. Note that this implies that analogously defined and can only take values in the set , and the longstanding open question asking whether all groups are sofic is equivalent to for all groups .
Let us recall that both proofs are based on a basic tool, often referred to as amplification, which uses the identity
(1.1) |
for matrices and . This identity allows one to show that there exists a function such that if one starts with a map with then the tensor power defined by satisfies . Moreover, starting from any , the sequence of iterates converges to . Hence, by iterating the tensor power operation, one can arrive at arbitrarily well sofic approximation. The case of hyperlinear groups is dealt with in a similar fashion. As it was observed in [AP17], (1.1) does not have an analog for linear sofic approximations. In [AP17], Arzhantseva and Păunescu invented a new amplification argument to prove that every linear sofic group is -linear sofic. A particularly innovative aspect of this argument is that it tracks two different quantities that when coupled together can be used to control the distance to the identity. Then using clever properties of ranks of tensor powers they prove that this amplification argument works. The question of whether the constant can be improved is left open in [AP17]. We will build upon their work to answer this question in the case of fields of characteristic zero.
1.2. Statement of results
In this paper, we will address the question of optimality of linear sofic approximations. The main results of this paper is the theorem below.
Theorem A.
Let be a countable linear sofic group over . Then
-
(1)
If is torsion-free, then is -linear sofic over .
-
(2)
Unconditionally, is -linear sofic over . Moreover, the constant cannot be improved.
Note that the assertion in Theorem A is in stark contrast with the case of sofic and hyperlinear groups. An interesting observation in [AP17] (see the paragraph before Proposition 5.12) is that the amplification argument does not see the interaction between group elements and will equally work for a subset of a group. Theorem A, however, shows that the optimal constant does indeed depend on the group structure and can even change by passing to a subgroup of finite index.
Remark 1.3.
Although we stated Theorem A over , one can easily see that it implies the same statement over all fields of characteristic zero; see Remark 5.1. However, an important part of the argument that is based on Lemma 4.2 does not work when has positive characteristic. See Remark 5.2 for more details. Henceforth, we write instead of .
We briefly outline the proof of Theorem A, which is following the main strategy of [AP17]. Given a finite subset and , we start with an -map (in the sense of Definition 1.1) provided by [AP17]. Using a sequence of functorial operations (see 2.1 for definitions) we replace with an -map which has the additional property that for every , at least of eigenvalues of matrix are . We will then show that the rank of tensor powers of are controlled by the return probability of a certain random walk on a finitely generated subgroup of . We will establish required effective non-concentration estimates in Theorem 3.9. This part of proof uses a variety of tools ranging from Fourier analysis to additive combinatorics. Let us note that the effectiveness of these bounds is a key element of the proof: as the asymptotic homomorphism condition (AH) deteriorates after every iteration of tensor power, we need to know in advance the number of required iterations so that we can start with an appropriate . The counter-intuitive move of adding ones as eigenvalues is needed for this purpose. When is close to unipotent, this argument completely breaks down. In this case, again using the method of [AP17] we will instead show the normalized number of Jordan block of tensor powers tends to zero with an effective bound for speed. This is carried out in Theorem 4.1 by translating the problem to estimating integrals of certain trigonometric sums. In summary, our proof can be viewed as a version of amplification argument where we use additional functors in the process.
Another result of this paper involves determining for finite groups.
Theorem B.
Let be a finite group.
-
(1)
There exists a finite-dimensional linear representation of such that
In particular, is a rational number.
-
(2)
iff has a fixed-point free complex representation.
-
(3)
Let denote the cyclic group of order . For prime and we have
In particular, as .
One of the main ingredients in the proof of Theorem B is the notion of stability. Broadly described, stability of a group in a mode of metric approximation demands that almost representations be small deformations of exact representations. Finite groups are easily seen to enjoy this property with respect to linear sofic approximations, see Proposition 6.2. Once this is established, the problem reduces to representation theory of finite groups. Proof of Part (a) of Theorem B is based on the simplex method in linear programming. Part (3) implements this for groups , . We finally remark that all finite groups to which (2) of Theorem A applies have been classified by Joseph Wolf, [Wol67]. They include groups such as . The next result establishes the value of for certain classes of groups over fields of positive characteristic.
Theorem C.
Let be a field of characteristic , and let be a finite group such that is the smallest prime dividing . Then
One of the problems posed in [AP17] is whether the notions of linear sofic approximation over and are equivalent. Theorem B and Theorem C together show that in general the values of and need not coincide for a finite group . This may be viewed as a quantitative reason for the difficulty of the problem of equivalence. We note that quantitative approaches to other metric approximations have also been considered before, see [AC20].
This paper is organized as follows: In Section 2, we will collect basic facts related to the rank metric, and basic theory of random walks on abelian groups and explain how they relate to our question. In Section 3, we prove various non-concentration estimates for random walks on abelian groups. Section 4 contains the proof of Theorem 4.1, involving matrices in Jordan canonical form. These ingredients are put together in Section 5 to prove Theorem A, except for the optimality claim. The optimality, as well as the proof of Theorems B and C, are discussed in Section 6.
Acknowledgement The authors would like to thank Goulnara Arzhantseva for helpful comments and suggestions on an earlier version of this paper. We also thank the anonymous referee for a careful reading of the paper and numerous remarks that significantly improved the exposition of this paper. Special thanks are due to Iosif Pinelis for providing the reference to Theorem 3.2.
2. Preliminaries and notation
In this section, we will set some notation and gather a number of basic facts needed in the rest of the paper. We will denote the group of invertible matrices over the field by . This space can be turned into a metric space by defining for :
We will often suppress the subscript and simply write . Every has eigenvalues , each counted with multiplicity. We write
For , let us denote by the set of all -th roots of unity in . It will later be convenient to consider the following quantity:
The number of Jordan blocks of (in its Jordan canonical form) will be denoted by . The number of Jordan blocks corresponding with on the diagonal will be denoted by . Note that
The following lemma plays a key role in the arguments used in [AP17]:
Lemma 2.1 ([AP17]).
For every , we have
The next lemma will be essential for tracking the multiplicity of eigenvalue in tensor powers of a matrix :
Lemma 2.2 ([AP17], Lemma 5.1).
Suppose that is the set of eigenvalues of a complex matrix , each counted with multiplicity. Then, for the set of eigenvalues of counted with multiplicity is given by
Proof.
There exists such that is upper-triangular. Hence, without loss of generality, we can assume that is upper-triangular with diagonal entries . It is easy to see that will be upper-triangular in an appropriate ordering of the bases, with diagonal entries given by the list above. Special case is dealt with in Lemma 5.1 of [AP17]. ∎
2.1. Three functorial operations
Let us now consider three functorial operations that can be applied to a family of matrices in . These will be used to replace an -map by an -map, which has some better properties. An alternative point of view is that each operation can be viewed as post-composition of the initial map by a representation of .
-
(1)
(Tensors) Consider the representations
where denotes the number of tensors. We will denote by .
-
(2)
(Direct sums) For , let
where the number of summands is . Instead of , we write .
-
(3)
(Adding Identity) For , consider the representation
where is the identity matrix of size . We write for .
Lemma 2.3.
Let and be matrices. Then for we have
-
(1)
-
(2)
-
(3)
Proof.
Parts (1) and (2) are straightforward computations. For (3) note that
The claim follows by a simple computation. ∎
Proposition 2.4.
Let be a linear sofic group. Then for every finite , and there exists an -map such that for all .
Proof.
Lemma 2.5.
Let and be matrices. Then .
Proof.
For , we have The claim follows from [AP17, Proposition 5] and the triangle inequality. The general case follows by a simple inductive argument. ∎
2.2. Preliminaries from probability theory
It will be convenient to reinterpret Lemma 2.2 in a probabilistic language. Let be a countable abelian group with the neutral element denoted by . By a probability measure on , we mean a map such that . We will then write . For , we define . We say that is finitely supported if there exists a finite set such that . The smallest set with this property is called the support of . Probability measures considered in this paper are finitely supported.
The convolution of probability measures measures and on is the probability measure defined by
One can see that the convolution is commutative and associative. The -th convolution power of will be denoted by . Given a probability measure on the group , the -random walk on (or the random walk governed by ) is the random process defined as follows. Let be a sequence of independent random variables, where the law of is . Define the process by and . It is easy to see that the law of is . We will use the notation to denote the probability of an event . Similarly, denotes the expected value of a random variable .
Given with eigenvalues , define the probability measure on
Proposition 2.6.
Let be a invertible complex matrix. Then
-
(1)
is a finitely supported probability measure on the multiplicative group .
-
(2)
-
(3)
For all integer we have
Proof.
Parts (1) and (2) follow from the definition of . Part (3) is an immediate corollary of Lemma 2.2. ∎
3. Effective non-concentration bounds on abelian groups
This section is devoted to proving effective non-concentration bounds for random walks on abelian groups. We will use the additive notation in most of this section. However, we will eventually apply these results to subgroups of the multiplicative group of non-zero complex numbers. Let be a finitely supported probability measure on a finitely generated abelian group . Our goal is to prove effective upper bounds for the return probability .
Lemma 3.1.
Let be a finitely supported probability measure on such that for some . Then
where is an absolute constant.
Note that the key aspect of Lemma 3.1 is that the decay rate is controlled only by without no further assumption on the distribution of . This fact will be crucial in our application. The proof of Lemma 3.1 is based on a non-concentration estimate in classical probability theory. Before stating the theorem, we need a few definitions. The concentration function of a random variable is defined by
We will use a theorem of Rogozin [Rog61], which generalizes a special case due to Kolmogorov[Kol58]. Our statement of the theorem is taken from [Ess66, Theorem 1], where a new proof using Fourier analysis is given. A variation of this proof can also be found in [Pet75, Chapter].
Theorem 3.2 (Rogozin).
Suppose are independent random variables and . Then for every non-negative we have we have
where is an absolute constant.
Proof of Lemma 3.1.
Let denote an embedding of into as an abelian group. Let denote the push-forward of which is a finitely supported probability measure defined by for every . Let be independent identically distributed random variables with distribution . Then by choosing all equal to we obtain
Letting we obtain where . Since , we have for all . Since , we have . The claim follows by noticing that . ∎
Remark 3.3.
It might be tempting to expect a similar upper bound for under the weaker assumption that . This is, however, not true. To see this, consider the probability measure with and . Although , we have for all .
We will now consider a variant of Lemma 3.1 for finite cyclic groups . In this case the uniform measure is the stationary measure and hence (under irreducibility assumptions) will converge to as . Note, however, that the random walk and its frequency of visits to zero can be bounded by how much the measure is supported on small subgroups. The next lemma provides an effective upper upper bound for only depending on the mass given to elements of small order.
Lemma 3.4 (Non-concentration for random walks on cyclic groups).
Let be a probability measure on . Further, suppose that for some and positive integer the following hold:
-
(1)
.
-
(2)
For every subgroup with , we have .
Then for all we have
where is an absolute constant.
The following proof is an adaptation of the proof of Littlewood-Offord estimate [NW21, Theorem 6.3]. This theorem is proven under different conditions, where instead of condition (2), a stronger assumption on the measure of all subgroups is imposed.
The proof uses the Fourier transform of measure and some of Let be a probability measure on . Write for . The following basic property of is used several times in the sequel. For each , we have . We will define the (Fourier) transform by
Lemma 3.5.
Let be a probability measure on . Then we have
-
(1)
.
-
(2)
For all , we have .
-
(3)
Proof.
Part (1) is clear. For (2), suppose that and are two probability measures on . Then we have
(3.1) |
Now, (2) follows by induction. For (3) note that
(3.2) |
where the last equality follows from the fact that for every , we have unless , in which case it is equal to . ∎
We can now start with the proof of Lemma 3.4.
Proof of Lemma 3.4.
Using parts (2) and (3) of Lemma 3.5 we have
Applying riangle inequality we deduce
Writing and using the inequality that holds for all we obtain
Set and define . Note that since , hence for all .
By separating the sum into level sets we obtain the following inequality:
(3.3) |
For an integer , and subsets we write
We use the shorthand for , where is the number of summands. The following lemma is proven in [Map10, Proposition 3.5] for all finite fields. A simple verification shows that it applies verbatim to all finite cyclic groups. For reader’s convenience we will sketch the modification needed in the proof.
Lemma 3.6.
For any and integer , we have
Proof.
It suffices to show that for all we have
This, in turn can be re-written as
This follows from the trigonometric inequality proven in [Map10, Proposition 3.1]. ∎
We will also need a lemma from additive combinatorics. Define the set of symmetries of a set by . Note that is a subgroup of and if then . The lemma follows by a simple inductive argument from [TV12, Theorem 5.5].
Lemma 3.7 (Kneser’s bound).
Let be an abelian group and are finite subsets. Then we have
In particular, for any finite we have
where , with the sum containing copies of .
We will now claim that if is a subgroup of with then there exists with . To prove this note that
Expanding the definition of in , and summing for we obtain
(3.4) |
Here, denotes the dual of , consisting of such that for all . Note that is a subgroup of . If then and hence . If then . Since , we have in this case as well. This implies that . This shows that implying that there exists such that .
Let us now suppose . Let be the largest positive integer with . We claim that . In fact, if this is not the case, using Lemma 3.6 we have
it follows that contains a subgroup with more than elements. We will then have
which contradicts that claim proved above. It follows from Lemma 3.7 that for all , we have
By the choice of we have
implying that . Putting all these together we obtain
Inserting the latter bound in the range and the trivial bound for into (3.3), we arrive at
(3.5) |
This ends the proof of Lemma 3.4.
∎
Lemma 3.8.
Let be a probability measure on and such that . Then for some constant and all we have
Proof.
This is proven is similar to the proof of Lemma 3.4. As before we write
We now claim that for all we have
To show this, for denote by the largest positive integer with . We claim that . Suppose that this is not the case. Recall that , where . Since we have . Since , we have . This implies that
it follows that contains a subgroup with more than elements, hence has to be equal to , from which it follows that . On the other hand, using the Plancherel formula we have
where the second inequality follows from . This implies that . Hence there exists such that . A similar argument as above finishes the proof. ∎
We will now combine Lemmas 3.1, 3.4, 3.8 to prove a non-concentration theorem for random walks on abelian groups with cyclic torsion component. Let where and are free and torsion components of . For , denote the projection from onto by , and for probability measure on , we write for the push-forward of under .
Theorem 3.9 (Non-concentration for random walks on abelian groups).
Let be a finitely generated abelian group with a cyclic torsion subgroup and a finitely supported probability measure on . Let and be a positive integer. Assume that
-
(1)
.
-
(2)
, where is the subset of consisting of elements of order at most .
Then for all we have
where is an absolute constant. Moreover, only under assumption (1) above, we have
Proof.
Clearly we have
Moreover, the inequality is satisfied for both . We consider two cases: if , then it follows from Lemma 3.1 that
Hence, suppose that . We claim that in this case we must have . If this is not the case, then which contradicts assumption (2). We will now verify condition (2) of Lemma 3.4 by finding an upper bound on the measure (under ) of the set of elements of order at most in .
(3.6) |
In particular, if is a subgroup with , then the order of every element of is at most , and hence . We can now apply Lemma 3.4 and Lemma 3.8 (with instead of ) to obtain the result. ∎
Corollary 3.10.
Let be a matrix in and . Assume that
-
(1)
.
-
(2)
Then for every there exists such that for all we have
Moreover, only under assumption (1), for every there exists such that, for all we have
Proof.
Denote by the subgroup of generated by the eigenvalues of . Denote the subgroup of torsion elements in by and let be a subgroup of such that . First suppose that conditions (1) and (2) hold. These guarantee that satisfies (1) and (2) of Theorem 3.9. It follows that for all we have
Given , we can find such that for all we have
This inequality together with the bound establishes the claim. The second part of the statement (involving only assumption (1)) follows by a similar argument. ∎
4. Dynamics of the number of Jordan blocks under tensor powers
In this section, we will study the effect of amplification on matrices with close to . If is such a matrix, for to be away from zero, a definite proportion of its Jordan blocks have to be of size at least . We will show that under this condition, the proportion of the number of Jordan blocks to the size of matrix tends to zero. Moreover, we will give an effective bound for the number of iterations required to reduce this ratio below any given positive .
We will denote by a Jordan block of size with eigenvalue . Let be a complex matrix. We denote by the number of Jordan blocks in the Jordan decomposition of divided by , hence , with iff is diagonalizable. Define a sequence of matrices by setting
(4.1) |
The main result of this section is the following theorem:
Theorem 4.1.
For every and every there exists such that if and then , where is defined by (4.1).
The proof of this theorem will involve estimating certain trigonometric integrals. Before reformulating the problem, we will recall some elementary facts.
Remark 4.3.
It follows from this lemma that the number of Jordan blocks of a given size in tensor powers of only depends on the size of Jordan blocks in and not on the corresponding eigenvalues. We will use this simple fact later.
In order to study the asymptotic behavior of the number of Jordan blocks in tensor powers of , it will be convenient to set up some algebraic framework. Let us denote by the set of all matrices in the Jordan normal form with eigenvalue on the diagonal, where varies over the set of all natural numbers. Note that if then . We will denote by the set of all formal linear combinations , where are integers and for all but finitely many values of . We equip with a binary operation defined by
and extended linearly to . Finally, for , consider the Laurent polynomial
and denote by the set of all (finite) integer linear combinations of . Note that since have different degrees, they do form a basis for the -module .
Now we will construct natural maps between these objects that will allow us to encode the number and size of Jordan blocks in tensor powers of a matrix using polynomials . First, define the map as follows. Let be a matrix with blocks of size for . Then define
Note that is a bijection. We now define by
Lemma 4.4.
The following hold:
-
(1)
For all , we have
-
(2)
For all we have
Proof.
For (1), note that when consists of one block of size and is a block of size , this is a restatement of Lemma 4.2. In this case, we have and and is precisely the expression defined by . It follows immediately from the definition of tensor product of matrices that the equality extends to all linear combinations with non-negative integer coefficients. This establishes (1).
In order to prove (2), we will show that For , the following identity holds:
Without loss of generality, assume that . We will proceed by calculating the coefficient of for on both sides. Note that appears on either the right-hand side or the left-hand side if and only if for some . Since the coefficients of and are equal, it suffices to consider the coefficient of for non-negative values of , which correspond to . Fix in this range. It follows from the definition that appears in if and only if , or, equivalently if . Since we have , the term appears exactly times on the right-hand side. We now show that the coefficient of on the left-hand side is the same. Note that the coefficient of the term with in is equal to the number of pairs such that appears in and appears in where . These conditions can be reformulated as bellow:
(4.2) |
Let us distinguish two cases: First suppose that . In this case we choose and any such choice determines uniquely. Hence the number of solutions to (4.2) is . When , then is allowed to vary in the range , and since , will automatically satisfy the inequality . In conclusion, the coefficient of on the left-hand side equals as well. This proves the claim. It thus follows that holds for and . Since is extended to by linearity, the general case follows immediately. ∎
Corollary 4.5.
Let be a invertible complex matrix. Let be the multiplicity of Jordan blocks of size in the Jordan decomposition of .
Proof.
Using Remark 4.3, we can assume that all eigenvalues are equal to , or . It is clear that . Part (1) of Lemma 4.4 implies that
Since , applying to the previous equation and using part (2) of Lemma 4.4 give
The claim follows by induction on . ∎
It will be more convenient to work with a trigonometric generating function. This is carried out in the next lemma.
Corollary 4.6.
Lemma 4.7.
Let and . For any real number such that we have
In particular, holds for all .
Proof.
We will proceed by induction on . For the required inequality is obvious. Assume that it also holds for . One can write:
Thus the required inequality holds for as well and we are done. ∎
Corollary 4.8.
For all we have .
For a measurable subset we denote by is the Lebesgue measure of . We will also denote the number of Jordan blocks of size in by .
Lemma 4.9.
Suppose that is in the Jordan normal form. For , set
Then
Proof.
Lemma 4.10.
Suppose is not unipotent, that is, . Then we have .
Proof.
First, notice that generates a Jordan block of size 1 if and only if the equation holds for some . Equivalently for . This equation has a solution if and only if . Therefore, we have
∎
Lemma 4.11.
For , define . Then for all we have
Proof.
As are fixed during this proof we write instead of . First note that
Using we can write:
∎
Corollary 4.12.
-
(1)
For all odd integers we have
-
(2)
For all even integers we have
Proof.
A direct computation shows that and . Repeated application of Lemma 4.11 extends this to all integers . In a similar fashion, checking that together with implies the claim for all odd values of . The last inequality is easy to check for . For we have:
and for we can write:
∎
Corollary 4.13.
Let and be defined as above. Then we have
Proof.
Proof of Theorem 4.1.
Let be such that , where . Let . We will find a function such that for all we have . By repeating this process we see that for all and all we have . Let be defined by so that . It is clear that
will satisfy the required property.
If we have we set , and we are done. Note that Proposition 5.8. of [AP17], we have for any matrix . Since , it follows that if we once the inequality holds for then the same inequality will continue to hold for all . Set let us assume that . Then we have
It follows from the choice of that . This implies that
Set so that the second term is bounded by . Now, let be the smallest positive integer such that
It is clear that for this value of we have
The claim will now follow from Corollary 4.13. ∎
5. Proof of Theorem A
In this section, we will prove parts (1) and (2) of Theorem A. The proof crucially depends on the inequality
from Lemma 2.1. Let be a torsion-free group and a finite subset of with . We will show that for every and an -map into exists. Let and set . By Lemma 2.4, there exists an -map such that
(5.1) |
for all . Here, is a small quantity whose values will be determined later. We will find such that for , the tensor power is an -map. Fix some . We will consider two different cases.
Case (1): Suppose that . Then it follows from Lemma 2.5 and Theorem 4.1 that for we have . This implies that
On the other hand, using Lemma 2.5, we have
as soon as . From here, we have by the triangle inequality that
Note that if is any matrix and then . In other words, we have . This implies that
Case (2): Suppose that . The proportion of blocks corresponding to eigenvalue is , and the proportion of other blocks is at most This implies that
Since we obtain
Also note that
It thus follows that
as long as . If denotes the size of the matrix , then we know that has at least Jordan blocks. If denotes the number of blocks corresponding to the eigenvalue , then noting that each such block contributes a one-dimensional subspace to , we deduce that . Note that by the assumption , that is, has at least Jordan blocks. This means that the total number of eigenvalues (with multiplicity) coming from blocks of size at least is at most . Hence, the multiplicity of as an eigenvalue of is at most .
Note that if are eigenvalues of then are eigenvalues of . If for some then . This implies that
(5.2) |
for . Applying Lemma 3.10 it follows that for , we have
It follows that . In conclusion, for , we have
for all . Note, however, that
Hence we need to choose with for the desired inequality to hold. The proof of (2) is similar and somewhat simpler. This time, we work directly with (and not ) and apply part (2) of Corollary 3.10.
Remark 5.1.
One can deduce from Theorem A the more general version in which the field of complex numbers is replaced by an arbitrary field of characteristic zero . In order to see this, suppose that is -linear sofic over . This implies that for every finite set and every and every , there exists and a map that satisfies properties and of Definition 1.1. Since is finite, one can replace by a finitely generated subfield of (depending on ). However, every such field is isomorphic to a subfield of . This implies that the arguments given above show that one can amplify . Now, note that all the amplifications are constructed via the functorial operations describe in 2.1. This implies that the image remains in for some , from which the claim follows.
6. Stability and the proof of Theorems B and C
In this section, we will prove Theorem B. Along the way, we will also address the more general question of determining when is an arbitrary finite group. As a byproduct, we will show that the bound cannot be improved for finite groups. This will be carried out through computation of , from which it will follow that as
The special case of will then prove the claim. One ingredient of the proof is the notion of stability for linear sofic representations. Studying stability for different modes of metric approximation has been an active area of research in the last decade. Stability of finite groups for sofic approximation was proved by Glebsky and Rivera [GR09]. Arzhantseva and Păunescu [AP15] showed that abelian groups are stable for sofic approximation. This result was generalized by Becker, Lubotzky, and Thom [BLT19] who established a criterion in terms of invariant random subgroups for (sofic) stability in the class of amenable groups. For other related results, see for instance [AP15, DCGLT20, BLT19, BL20] and references therein. Some progress towards proving the stability of in linear sofic approximation has been made in [EG21]. Our first theorem establishes the stability of finite groups in the normalized rank metric. Before stating and proving this result, we will need a simple fact from linear algebra.
Lemma 6.1.
Suppose that are subspaces of with , for . Then, we have .
Proof.
We will proceed by induction on . For there is nothing to prove. Assume that the claim is shown for . Then, using the induction hypothesis, one can write:
∎
Proposition 6.2.
Let be a finite group, and is such that for all we have
Then there exists a representation such that for every we have
Proof.
For consider the subspace defined by
and set . We claim that is a -invariant subspace of . Assume that is an arbitrary element of . It suffices to prove that for any , is an element of as well. Since , we can write:
This implies that , proving the claim. In summary, is a -invariant subspace with the property that the restriction of to is a representation of . Let be a subspace complement of . For , define to be the linear transformation that acts on via and on by identity. It is clear that defined in this way is a -representations. Finally, for every we have:
This finishes the proof. ∎
Remark 6.3.
It is noteworthy that our proof establishes stability with a linear estimate. However, the constant depends on . It would be interesting to see if this dependency can be relaxed for certain families of groups.
We will start by providing a simple description of irreducible representations of the group . We start by setting some notation. Recall that denotes the character . Also, for , we write . For each , define by . It is a well known [Luo09] fact that , as constitute all irreducible representations of .
Proposition 6.4.
For we have
Proof.
Let
denote the direct sum of all other than the trivial representation . One can regard as a diagonal matrix of size with diagonal entries for non-zero . We claim that for every , the set of has cardinality . In fact, the condition corresponds to the equation for which has exactly non-zero solutions. Hence , from which it follows that
To prove the reverse inequality, we first suppose is a -dimensional representation of . Decompose into a direct sum of irreducible representations and denote the multiplicity of by :
Set
and write . This implies that for every non-zero , we have
(6.1) |
Without loss of generality, we can assume that , since by removing the trivial representation the value of can only increase. Note also that for every , there are exactly elements with . This implies that
This together with (6.1) implies that .
Now, to finish the proof assume that Choose such that . Let be such that
holds for all . Use Lemma 6.2 to find a representation such that . Since for all non-zero , it follows that for all
This is a contradiction. ∎
Corollary 6.5.
The best constant for the class of all groups is .
6.1. The value of for finite groups
Let be an arbitrary finite group. Note that it follows from Proposition 6.2 that
where ranges over all finite-dimensional representations of . The first observation is that the supremum can be upgraded to a maximum.
Proposition 6.6.
Let be a finite group. Then there exists a finite-dimensional representation of such that
In particular, is a rational number.
Proof.
Denote by the set of all irreducible representations of up to isomorphism. Pick to be a set of representatives for all conjugacy classes of . We will assume that is the trivial representation and is the identity element of . Consider the matrix where
Note that does not depend on the choice of the representative . Set , and write
For each representation which does not contain the trivial representation, we can decompose as , and define to be the column vector
of normalized multiplicities of irreducible representations of . It is easy to see that holds iff for all
These conditions can be more succinctly expressed as
where we write for two vectors if every entry of is non-negative. By assumption, for every , there exists a representation such that holds. In other words, the set
is non-empty. By compactness of , it follows that there exists a point such that . Note that these constitute a system of inequalities involving and with rational coefficients. Now using the fact the the set of solutions to this system is a rational polytope (or equivalently using the proof of Farkas’ lemma [Mat07]) we deduce that both and are rational. This proves the claim.
∎
Remark 6.7.
There are noncyclic finite groups for which . For instance, let be a finite subgroup of . Such groups are cyclic of odd order and double covers of finite subgroups of , which include the alternating groups of letters, and the symmetric group on letters. We claim that the natural representation of in has the property that for the eigenvalues of are not . This is clear since if one eigenvalue is , then the other has to be as well, contradicting the faithfulness of . These groups include, for instance, .
Proposition 6.8.
Let be a finite group. Then iff has a fixed-point free complex representation.
These groups have been classified by Joseph A. Wolf. The classification is rather complicated. We refer the reader to [Wol67] for proofs and to [Nak74, Theorem (1.7)] for a concise statement and the table listing these groups. The difficulty of classifying finite groups with suggests that the problem of determining in terms of may be a challenging one.
Remark 6.9.
Given a field , one can also study the notion of linear sofic approximation over . It is not known whether linear sofic groups over and other fields coincide. However, one can show that and do not need to coincide. This will be seen in the next subsection.
6.2. Optimal linear sofic approximation over fields of positive characteristic
In this subsection, we will prove Theorem C
Proof of Theorem C.
Let be a field of characteristic . It is easy to see that the proof of Proposition 6.2 works over without any changes. Hence, for any finite group we have
where runs over all representations
Suppose is a linear representation. Let be an element of order in . Write where is a matrix over . From , and has characteristic , it follows that . Let denote the Jordan canonical form of , consisting of blocks. It follows from that each block in is a nilpotent block of size at most , implying . Since we have
This shows that . To prove the reverse inequality, let denote the vector space consisting of all functions . Clearly . Consider the left regular representation of on defined by
Let , and consider the subspace
Any is invariant from the left by the subgroup generated by . It follows that
Hence
proving the claim. ∎
References
- [AC20] Goulnara Arzhantseva and Pierre-Alain Cherix, Quantifying metric approximations of discrete groups, 2020.
- [AP15] Goulnara Arzhantseva and Liviu Paunescu, Almost commuting permutations are near commuting permutations, J. Funct. Anal. 269 (2015), no. 3, 745–757. MR 3350728
- [AP17] Goulnara Arzhantseva and Liviu Păunescu, Linear sofic groups and algebras, Trans. Amer. Math. Soc. 369 (2017), no. 4, 2285–2310. MR 3592512
- [BL20] Oren Becker and Alexander Lubotzky, Group stability and Property (T), J. Funct. Anal. 278 (2020), no. 1, 108298, 20. MR 4027744
- [BLT19] Oren Becker, Alexander Lubotzky, and Andreas Thom, Stability and invariant random subgroups, Duke Math. J. 168 (2019), no. 12, 2207–2234. MR 3999445
- [DCGLT20] Marcus De Chiffre, Lev Glebsky, Alexander Lubotzky, and Andreas Thom, Stability, cohomology vanishing, and nonapproximable groups, Forum Math. Sigma 8 (2020), Paper No. e18, 37. MR 4080477
- [EG21] Gábor Elek and Łukasz Grabowski, Almost commuting matrices with respect to the rank metric, Groups Geom. Dyn. 15 (2021), no. 3, 1059–1083. MR 4322021
- [ES05] Gábor Elek and Endre Szabó, Hyperlinearity, essentially free actions and -invariants. The sofic property, Math. Ann. 332 (2005), no. 2, 421–441. MR 2178069
- [Ess66] Carl-Gustav Esseen, On the Kolmogorov-Rogozin inequality for the concentration function, Z. Wahrscheinlichkeitstheorie und Verw. Gebiete 5 (1966), 210–216. MR 205297
- [GR09] Lev Glebsky and Luis Manuel Rivera, Almost solutions of equations in permutations, Taiwanese J. Math. 13 (2009), no. 2A, 493–500. MR 2500002
- [Gro99] Mikhael Gromov, Endomorphisms of symbolic algebraic varieties, J. Eur. Math. Soc. (JEMS) 1 (1999), no. 2, 109–197. MR 1694588
- [II09] Kei-ichiro Iima and Ryo Iwamatsu, On the Jordan decomposition of tensored matrices of Jordan canonical forms, Math. J. Okayama Univ. 51 (2009), 133–148. MR 2482411
- [Kol58] Andery Kolmogorov, Sur les propriétés des fonctions de concentrations de M. P. Lévy, Ann. Inst. H. Poincaré 16 (1958), 27–34. MR 101545
- [Luo09] Bao Luong, Fourier analysis on finite abelian groups, Applied and Numerical Harmonic Analysis, Birkhäuser Boston, Ltd., Boston, MA, 2009. MR 2537697
- [Map10] Kenneth Maples, Singularity of random matrices over finite fields, 2010.
- [Mat07] Jiri Matousek, Understanding and using linear programming, Springer, Berlin, 2007.
- [Nak74] Minoru Nakaoka, Finite groups which act freely on spheres, Osaka Math. J. 11 (1974), 323–330. MR 356108
- [NW21] Hoi H Nguyen and Melanie Matchett Wood, Random integral matrices: universality of surjectivity and the cokernel, Invent. Math. 208 (2021).
- [Pet75] Valentin Vladimirovich Petrov, Sums of independent random variables, Ergebnisse der Mathematik und ihrer Grenzgebiete, Band 82, Springer-Verlag, New York-Heidelberg, 1975, Translated from the Russian by A. A. Brown. MR 0388499
- [Rog61] Boris Alekseevich Rogozin, An estimate of the concentration functions, Teor. Verojatnost. i Primenen 6 (1961), 103–105. MR 0131893
- [R0̆8] Florin Rădulescu, The von Neumann algebra of the non-residually finite Baumslag group embeds into , Hot topics in operator theory, Theta Ser. Adv. Math., vol. 9, Theta, Bucharest, 2008, pp. 173–185. MR 2436761
- [TV12] Terence Tao and Van Vu, Additive combinatorics, Cambridge University Press, 2012.
- [Wei00] Benjamin Weiss, Sofic groups and dynamical systems, vol. 62, 2000, Ergodic theory and harmonic analysis (Mumbai, 1999), pp. 350–359. MR 1803462
- [Wol67] Joseph Albert Wolf, Spaces of constant curvature, McGraw-Hill Book Co., New York-London-Sydney, 1967. MR 0217740