remresetThe remreset package \newaliascntPropositionequation\aliascntresettheProposition \newaliascntTheoremequation\aliascntresettheTheorem \newaliascntMain Theoremequation\aliascntresettheMain Theorem \newaliascntCorollaryequation\aliascntresettheCorollary \AtEndEnvironmentCorollary \newaliascntLemmaequation\aliascntresettheLemma \newaliascntDefinitionequation\aliascntresettheDefinition \newaliascntNotationequation\aliascntresettheNotation \newaliascntQuestionequation\aliascntresettheQuestion \newaliascntExampleequation\aliascntresettheExample \AtEndEnvironmentExample \newaliascntExamplesequation\aliascntresettheExamples \AtEndEnvironmentExamples \newaliascntRemarkequation\aliascntresettheRemark \AtEndEnvironmentRemark \newaliascntAssumptionequation\aliascntresettheAssumption
Fractal behavior of tensor powers of the two dimensional space in prime characteristic
Abstract.
We study the number of indecomposable summands in tensor powers of the vector representation of SL2. Our main focus is on positive characteristic where this sequence of numbers and its generating function show fractal behavior akin to Mahler functions.
Key words and phrases:
Tensor products, asymptotic behavior, fractals, Mahler functions, subcategories of finite tensor categories.2020 Mathematics Subject Classification:
Primary: 11N45, 18M05; Secondary: 11B85, 26A12, 30B30.1. Introduction
Let be a field and let be a group, possibly an algebraic group or affine group scheme over . For any finite dimensional -representation over let be an integer such that
where are indecomposable -representations. This number is well-defined by the Krull–Schmidt theorem. Now let be a finite dimensional -representation and define the integer sequence
The study of the growth of is the main motivation of this paper.
The related question to study the growth of , where is the length of , is usually much easier and we discuss this along the way, e.g. in Section 2A and Section 2BV below.
We write for ‘asymptotically equal’. One could hope that
(1.1) |
In practice, is often a constant or alternates between finitely many constants, but sometimes is more complicated.
We do not know in what generality Equation 1.1 holds, but expressions of this form are very common in the theory of asymptotics of generating functions, see for example [FS09] or [Mis20] for the relation between counting problems and the analysis of generating functions.
1A. The exponential factor
Even without assuming (1.1) one can define the value , and it is proven in [COT24, Theorem 1.4] that
(1A.1) |
Thus, Equation 1A.1 gives as the exponential factor of the growth rate of .
The goal of this paper is to provide a more precise asymptotics, that is, one of the form Equation 1.1, for the sequence in the case and or, equivalently, . This is a nontrivial case where one can hope to understand the asymptotics explicitly. In this setting the subexponential factor is determined by Section 1B, as well as Section 2A and Section 2A.
We do not calculate , but it appears to be oscillating, cf. Section 1B. See also Section 8 where we determine for .
1B. The main theorem – the subexponential factor
If is finite, then is a finite group and the asymptotic behavior of can be easily obtained, see for example Section 2A below or [LTV23, Example 8] which settle the case of a finite field. We give an example which is prototypical in this situation:
Example \theExample.
For we get
(We highlight that the subexponential factor is .) Note that does not exist, but the ratio is rather governed by the periodic function .
From now on let be a two dimensional vector space over an infinite field. Let be the algebraic group . Let be the sequence as above. As a matter of fact, the sequence depends only on the characteristic of (which is only implicit in our notation), so we only need to study it for one infinite field for each . That is, we have
(1B.1) |
where is either for or otherwise.
The following example is classical, and settles the case :
Example \theExample.
Assume . Then is the sequence of middle binomial coefficients
Then Stirling’s formula implies
as one easily checks. In this case the periodic function is constant, i.e. . Since this situation is semisimple, the same formula works for the length instead of the number of indecomposable summands.
For the remainder of the paper let . Then the asymptotic behavior of the sequence turns out to be more complicated. Namely let us define
(1B.2) |
As we will see, plays the role of in Equation 1.1. To give some numerical expressions, we have for example:
It is easy to see that all are irrational. Moreover, all these numbers are transcendental, as follows from the (logarithm form of the) Gelfond–Schneider theorem which implies that , for a prime number and , is transcendental unless is a power of .
Remark \theRemark.
By very similar arguments, all that we will see below that are not manifestly rational will be irrational and even transcendental.
Notation \theNotation.
Throughout, we will use big O notation (or rather its generalization often called Bachmann–Landau notation). Most important for us from this family of notations are as above, small , big , big and . That is, given two real-valued functions , which are both defined on or on for the final point, we will write
(1B.3) |
We sometimes omit the subscript if no confusion can arise.
Our main result is:
Main Theorem \theMain.Theorem.
For any there exist such that we have
Thus, .
Moreover, for we have a quite complete picture and we prove stronger results: we show that Equation 1.1 holds and we determine , see Section 7 and Section 8. We expect that similar methods can be used to handle the case .
Remark \theRemark.
As usual in modular representation theory, the case behaves like . Indeed, we have
so we can compare Section 1B with Section 1B.
Remark \theRemark.
In contrast to characteristic zero as in Section 1B, the limit
does not exist, i.e. for there is no constant such that .
Example \theExample.
For we have and the logplot of the even values of (splitting the sequence into even and odd makes it more regular, see for example Section 5) gives
These illustrations show the graph of for and (the fluctuations in the beginning vanish quickly) and we see the tiny oscillation, which is proven to be true in Section 8.
Remark \theRemark.
For Section 1B, or rather the stronger version in Section 7B, was independently proven in [La24] using quite different methods. Moreover, and similar but different to Section 8, the paper [La24] also identifies from Equation 1.1. For [La24] even implies some results for other representations as well.
Remark \theRemark.
Let us comment on the quantum case; the analog of Section 1B for quantum . If the quantum parameter is not a root of unity, then the same discussion as in Section 1B works. The next case, where the quantum parameter is a root of unity and the underlying field is of characteristic zero can be deduced from [LPRS24]. For the final possibility, the so-called mixed case as e.g. in [STWZ23], we expect a result very similar to Section 1B.
1C. Proof outline of Section 1B
It is a classical fact that all direct summands of are tilting representations over . In the special case of the characters of indecomposable tilting representations are explicitly known; using this we derive a recursion for the sequence , see Section 3B.
Then we translate this recursion into a functional equation for the generating function of the sequence , see Section 3A. Then we analyze the behavior of the function in the vicinity of its radius of convergence and deduce Section 1B by employing suitable Tauberian theorems in Section 6.
One of the steps is the following elementary looking inequality from Section 5A:
1D. Odds and ends
Here are a few open questions that might be interesting to explore. (i) is proven for , see Section 7 and Section 8, and similar methods apply in general.
-
(i)
In Section 5A we prove that . A natural question is whether we have . Moreover, recall from Section 1B that there is no constant such that . However, one could conjecture that there exists a continuous real function of period one such that
(1D.1) For this is proven in Section 7B.
-
(ii)
Recall from Section 1B that is bounded by two constants, and it would be interesting to have good explicit values for these constants, see Section 9D for some possible values. For a more precise asymptotic formula one would need to analyze the oscillation as in Section 1B.
-
(iii)
One could try to compute other asymptotic formulas. An example that comes to mind is to verify an analog of Section 1B (with the same subexponential factor) for the three dimensional tilting -representation (in case ). When this three dimensional -representation is not tilting and getting asymptotic formulas might be very hard.
Acknowledgments. We thank Michael Larsen for sharing a draft of [La24], which uses very interesting methods that are quite different from the ones in this work. The methods are exciting and important, so we all agreed that it makes sense to write two separate papers. We also thank Henning Haahr Andersen, David He, Abel Lacabanne and Pedro Vaz for useful discussions and comments, and are very grateful to Kenichi Shimizu for pointing out a gap in the proof of Appendix A in the first version.
We express our appreciation to ChatGPT for their assistance with proofreading. In addition, D.T. extends heartfelt thanks to the tree constant for inspiration.
P.E.’s work was partially supported by the NSF grant DMS-2001318, D.T. was supported by the ARC Future Fellowship FT230100489.
2. Fractal behavior of growth problems
Before coming to the main parts of this paper, let us briefly indicate the main difficulty (and maybe the most exciting part) of growth problems in the above sense: a certain type of fractal behavior of these sequences and of their generating functions. For the sake of this paper, and partially justified by the discussion in this section, we use fractal behavior to mean that the exponent of the subexponential factor is transcendental (note that e.g. from Equation 1B.2 is transcendental), or at least irrational. For instance, could be the fractal dimension of some fractal.
2A. No fractal behavior
For of characteristic zero, let be a connected reductive algebraic group over . [Bia93, Theorem 2.2] and [CEO24, Theorem 2.5] give the asymptotic for the numbers for an arbitrary finite dimensional -representation. The example of and its vector representation is given in Section 1B where .
In general, [Bia93, Theorem 2.2] and [CEO24, Theorem 2.5] prove (since this case is semisimple the same holds true for the length instead of the number of indecomposable summands):
Proposition \theProposition.
In the above setting, the asymptotic takes the form
Thus, the subexponential factor always has some half integer exponent.∎
It is not a coincidence that since:
“The nature of the generating function’s singularities determines the associated subexponential factor.” |
The above strategy is well-known in symbolic dynamics, and under certain assumptions on the singularities of the generating function one always gets half integer powers for the subexponential growth term. For example, this works if the generating function is algebraic.
In fact, there are well developed algorithms to compute the asymptotics if the generating function is sufficiently nice (e.g. meromorphic), see for example [Mis20, Section 7.7]. The algorithm presented therein always produces an exponent and is enough to treat the case of a connected reductive algebraic group in characteristic zero.
Here is another example where one always gets a half integer coefficient, namely zero, regardless of the characteristic. As we will argue below, this means that these cases do not show fractal behavior with respect to the growth problems we consider.
Let be a finite group and let be a representation of over , with of arbitrary characteristic. We assume for simplicity that is a faithful -representation; if is not faithful we replace by a quotient and continue as below. Let be the central cyclic subgroup of containing all scalar matrices. Let . The category of finite dimensional -representations splits into a direct sum
where acts on objects of in the same way as in .
Let be the set of isomorphism classes of simple objects of ; for any let be its projective cover. For any we define
Let be the regular representation of and let be its decomposition into summands . A result of Bryant–Kovács [BK72, Theorem 2] says that for sufficiently large with we have that is a direct summand of . Similarly to , let , where denotes the length of representations. Assuming , it follows easily that for we have
In particular, each of the sequences and has at most limit points. In particular, [Mis20, Section 7.7] applies and the subexponential factor has half integer exponent. A similar analysis works without the assumption .
This immediately proves that , in this case, is periodic with period , the subexponential factor is trivial and the exponential factor is given by the dimension. Precisely:
Proposition \theProposition.
For a finite group and a -representation we always have
for some and all .∎
The scalars in Section 2A are easy to compute. For see for example [CEO24, Proposition 2.1] and [LTV23, (2A.1)] for characteristic zero, and [LTV24, Section 5] for arbitrary characteristic.
Example \theExample.
For instance if , and is the symmetric group or and is a faithful representation of we have and .
Remark \theRemark.
Let be a finite tensor category, see e.g. [EGNO15, Chapter 6] for details. Let be a tensor generator of . We have a decomposition of over the universal grading group of . Write for indecomposables with of degree . Let be the subgroup of generated by . Then is a cyclic group , and statements similar to Section 2A hold. Details are omitted, but the key statement hereby is proven in Appendix A in the appendix.
To summarize, in the two above settings the subexponential exponent is a half-integer, in particular not transcendental, and the function is constant up to some period .
2B. Fractal behavior
For our problem, where is of prime characteristic, is difficult also because the generating function that we compute in Section 3A does not have nice enough properties to run the classical strategies. For example, we will see that we have to face a dense set of singularities, see e.g. Section 4B. And in fact, the exponent we get is not a half integer but rather the transcendental number from Equation 1B.2.
In a bit more details, see Section 3C for the precise statement, we get a functional equation for the generating function of the numbers that takes the form
(2B.1) |
Functions of this type are called 2-Mahler functions of degree .
Remark \theRemark.
The name originates in Mahler’s approach to transcendence and algebraic independence results for the values at algebraic points in the study of power series satisfying functional equations of a certain type. Mahler’s original functional equation is of the form , which is an example of what we call a 2-Mahler function of degree 2.
Let denote rational functions. The Mahler functions above have been generalized under the umbrella of -Mahler functions of degree satisfying
(2B.2) |
but we will not need this generalization. We refer to [Nis96] for a nice discussion of Mahler functions and a list of historical references. Mahler functions occur in combinatorics as generating functions of partitions and related structures.
A crucial fact is that such a Mahler function often grows with exponent for a transcendental when approaching its relevant singularity. We will use this in Section 4.
Moreover, connections from Mahler functions that grow with transcendental to fractals are well-studied. To the best of our knowledge, general theorems relating them are not known, and connections are only example-based. In the rest of this section, we give several examples that are easier to deal with than our main result.
The common source for these examples is the well-known principle that projecting -adic objects onto the real world leads to fractals. At the same time, modular representations of algebraic groups are known to exhibit -adic patterns, stemming from Steinberg-type tensor product theorems. Therefore, one can expect that combinatorial invariants of modular representations of algebraic groups (such as dimensions, weight multiplicities, etc.) tend to exhibit fractal behavior. In fact, fractal behavior in the study of reductive groups in prime characteristic has been observed several times, and is part of folk knowledge.
Most relevant for this paper are fractal patterns within the study of tilting representations. For example, for an algebraically closed field of prime characteristic , let . The multiplicities of Weyl in tilting -representations have fractal patterns of step size and so do their characters, by Donkin’s tensor product formula [Don93, Proposition 2.1] (this is a bit easier to see in the reformulation given in [STWZ23, Section 3]). The same is reflected in the Temperley–Lieb combinatorics, see e.g. [BLS19, Theorem 2.8] or [Spe23, Figure 3], which is exploited in [KST24].
For higher rank there are also several known instances of fractal behavior of tilting characters, see for example the picture of cells in [And04, Figure 1] or billiards of tilting characters as in [LW18, Section 4].
2BI. Modular representations and the Cantor set
One of the simplest examples of a fractal is the (bounded) Cantor set , which can be realized as the set of real numbers in which admit a ternary expansion without digit with the probably familiar picture
This example is related to representation theory of for being algebraically closed of characteristic (the standard choice is ) as follows.
The finite dimensional simple -representations are indexed by , their highest weight. Expressing the number -adically, say with , one gets
where the exponent (i) denotes Frobenius twists which acts on characters by . This is a special case of Steinberg’s tensor product theorem. Since the character of for is , this gives a description of the weights of for .
This also shows that we may consider, often infinite dimensional, representations where is a -adic integer, which is made explicit in Haboush’s generalized Steinberg tensor product theorem for distribution algebras [Hab80, Theorem 4.9]. Concretely, given , we may define to be the infinite tensor product
which, by definition, is the span of tensor products of vectors in such that almost all of them are (fixed) highest weight vectors.
The group does not act in this space, but admits an action of the distribution algebra , for which this representation is generated by the highest weight vector . This gives a -grading with even degrees, placing vectors of weight in degree . The Hilbert series for this grading is thus
This is a holomorphic function for , and we will mostly consider it on the interval , i.e. as .
Consider now the special case , . Hence, for all and we get
This function satisfies the functional equation
This is a 2-Mahler function of degree , and the relevant singularity is at .
The Taylor coefficients of form the Cantor set sequence defined by
This sequence is well-studied, see e.g. [CE21, Section 1] and [OEI23, A292686], and is its generating function.
The corresponding fractal is constructed as follows. Let denote the set of degrees of , so that is the set of nonnegative integers with ternary representations where all digits of take values only. For , define . We then have . Let . Then the closure of in is the unbounded Cantor set , which is invariant under multiplication by . The set is the intersection , and (nested union). Note that the set comes with a natural measure , the Hausdorff measure of : the weak limit of the counting measures of rescaled by dividing by .
Now, by the standard theory of Mahler functions, see e.g. [BC17, Theorem 1], we get
for some with . Moreover, is found by substituting into the Mahler equation, i.e. from the condition as . This yields , hence
The number is transcendental, and the Hausdorff (and Minkowski) dimension of the Cantor sets and .
Rewriting the above using , see Equation 1B.3 for the notation, we get
Then Tauberian theory (as for example in Section 6A) implies
This describes the asymptotics of Cesáro sums of the Cantor set sequence, which counts the dimension of the subspace spanned by vectors in of degree .
One may further ask how the sequence behaves when or, related, how the function (equivalently, ) behaves when . This behavior can be analyzed as follows. Let for and for . We have
for the function
This is a periodic function in the sense that (note that the product is absolutely convergent). This implies that the function has no limit as and asymptotically has oscillatory behavior: it approaches the periodic function .
Plotting this illustrates the overall growth rate and the oscillation:
(2B.3) |
To analyze this a bit further, it is easy to see that the analytic function in the region , which satisfies the equation , is nothing but the Laplace transform of the Hausdorff measure of the set :
This implies that
Similarly, the sequence for large behaves as the devil’s staircase function:
which is periodic under the map . This is visualized in Equation 2B.3. Hence, this sequence does not have a limit as and exhibits oscillatory behavior, approaching the periodic function . Note that this function is continuous but not differentiable, nor absolutely continuous, since it is the integral of a singular measure (the Hausdorff measure of ). It is, however, Hölder continuous with exponent . We can also find the range of oscillation of . Indeed, it is easy to see that (attained for ) while (approached for ).
Remark \theRemark.
The function is holomorphic in the strip and periodic under , so we may consider its Fourier coefficients . They are related to the Fourier coefficients of the measure by the formula
Here and throughout denotes the gamma function evaluated at (not to be confused with the group ). Since has unit volume on the period, we have , . Hence:
The numbers are tiny for large , see the above loglogplot.
2BII. Generalization to general primes and highest weights
A similar analysis applies for any prime and rational highest weights (excluding positive integers, for which is finite dimensional). Namely, in this case has an infinite periodic expansion, which can be computed in the standard way: write as where and , and pick the minimal such that divides . Then the repeating string of the expansion of is the base expression of the number (a string of length , with zeros at the beginning if needed). If , where are its base digits, then
The corresponding fractal is the set of real numbers which have a base expansion with th digit in , which has Hausdorff (and Minkowski) dimension . Note that for , (so ), while for we get an actual fractal, i.e. is transcendental. For example, for we have and as explained above.
The deeper analysis of the oscillating behavior of the character of and the sequence of Cesáro sums of its coefficients (i.e., the sequence of dimensions of weight spaces ) also extends mutatis mutandis to general and . In fact, the power behavior of the sequence is observed for sufficiently generic . Namely, if and there exists a limit
then we have
This includes rational , and also if is chosen randomly (all digits are independent and uniformly distributed) then
For example, for , let . Consider the base Cantor set sequence:
Thus, we have:
All of these values are transcendental.
Then the analog of Equation 2B.3 for and is:
2BIII. Generalization to higher rank
The above analysis can also be extended to simply connected simple algebraic groups of arbitrary rank . In this case, a weight is an -tuple , (the coefficients of with respect to fundamental weights), and a fractal attached to the simple highest weight module over the distribution algebra can be built in a Euclidean space , where is the set of positive roots of , for which we choose an ordering. For example, suppose , where is an integral weight with . Then the fractal attached to depends on a choice of a PBW basis of .
Namely, pick a collection such that the set of vectors , (product in the chosen order) forms a basis of (where is the highest weight vector). Then we define as the set of functions which have a base expansion such that for all , the th digit of belongs to . The Hausdorff (and Minkowski) dimension of equals
for any choice of . (However, it must be mentioned that is notoriously hard to compute in general.)
Also, comes equipped with a natural measure , obtained by suitable rescaling of the counting measures as before (a multiple of the Hausdorff measure). Finally, we have a proper map given by , and the measure does not depend on , nor on the ordering of positive roots (it expresses the large-scale asymptotics of the character of ). This measure completely determines the large-scale behavior of .
2BIV. Modular representations and Sierpinski gaskets
Fix a prime . To give another example, recall Sierpinski’s gasket or triangle, which is Pascal’s triangle modulo . For we have (we only illustrate cutoffs):
Sierpinski’s triangle is combinatorially given by the limit of the above pictures, keeping the black boxes and disregarding the white boxes.
This example is related to representation theory of as follows.
For each let us plot the weights of for on the line on the coordinate plane: the weight corresponds to the point . Let us rescale the obtained set by the factor and denote the resulting set by . Thus is the set of pairs , , where are -digit numbers in base such that there are no carries in computing (i.e., for all , ). This shows that .
Let and be the closure of . Hence, is the compact set of pairs such that , in base , and . This set is the Sierpinski triangle from above.
Remark \theRemark.
The set also has a direct representation-theoretic interpretation. Namely it corresponds to the set of weights in the simple representations of the (infinite type) affine group scheme , obtained by ‘perfecting’ , see [CW24, §7.1]. It would be interesting to have a similar direct interpretation for from Section 2BI. This is less obvious since distribution algebras of perfected group schemes are trivial, see [CW24, Lemma 3.1.2].
It is easy to see that . From this it is not hard to deduce the well-known fact that the Hausdorff dimension of is the number given by
This number is transcendental, and can also be seen at the level of the generating function given by
Steinberg’s tensor product theorem yields that is a 2-Mahler equation of degree :
We have , so we have for some
This implies that for some , we have, for large enough :
This is illustrated in Equation 2B.4 below. A more detailed asymptotics of can be obtained as follows. Note that
Thus, we see that
where (note that the product below is absolutely convergent)
is a periodic function in the sense that . This implies that the sequence also exhibits oscillatory behavior:
(2B.4) |
Let be the direct image of the (suitably normalized) Hausdorff measure on under the map . Then the analytic function on is the Laplace transform of :
Similarly, the sequence for large behaves as follows:
Thus, this sequence approaches the periodic function at infinity. As before, this function is continuous but not differentiable, nor absolutely continuous.
2BV. Lengths of tensor powers
Recall that denotes the length of a representation. Finally let us discuss the fractal behavior of the integer sequence where is the vector representation of . We will not explicitly consider any fractals in this example, however the origin of what we, based on the previous examples, call ‘fractal behavior’ is still found in the same principles, such as Steinberg’s tensor product theorem.
. For simplicity consider the case first. In this case one can show that the generating function
is holomorphic for and satisfies the functional equation
In particular, this shows that for all . This is not a Mahler equation, but it turns into one under a simple change of variable. Namely, setting and , we have
i.e., is a 2-Mahler function of degree .
One gets
Using this, we can compute as previously the asymptotics of as . Set
which is transcendental. We then have, again by the theory of Mahler functions,
for some with . Hence, we get
So from the Tauberian theory, cf. Section 6A, it follows that
Or, equivalently,
As previously, the behavior of as can be analyzed as follows. We have
where (for as in Section 2BI above)
This is a periodic function in the sense that , and the function asymptotically approaches the periodic function as . Writing , we obtain that the function approaches as , i.e.,
Or, equivalently,
But
so we get
Thus, setting , we obtain
Hence, the function
is periodic with period and analytic for , i.e., the strip of holomorphy is twice as wide as in previous examples. This happens because the role of the prime in this example is played by the number (rather than ), i.e., as , the function behaves as a Mahler function of degree four, rather than a Mahler function of degree two. As we will see, this will lead to much greater regularity of the coefficient sequence .
We would now like to understand the asymptotics of the sequence in more detail. If was known to behave sufficiently regularly, we could read off its asymptotics from the asymptotics of the Cesáro sums by Abel resummation:
The required regularity is guaranteed by the following lemma, showing that the sequence is decreasing.
Lemma \theLemma.
We have .
Proof.
Let , where we agree that . Then the functional equation for implies that
So the coefficients of satisfy the recursion
This shows that for all , which implies the statement. ∎
From this it follows (with some work) that
where is a periodic function in the sense that (period doubling with respect to the previous examples). Thus, behaves roughly like with (as expected, growing faster than in characteristic zero, where it is , see Section 1B). Here is the plot (for ; note that ):
To see the oscillation, we zoom into :
As before, the Fourier coefficients of the 1-periodic function are related to the Fourier coefficients of the function by the formula
So since is analytic in the strip , we have for all , hence
Thus, the function is analytic in the strip .
This is a significant difference from the previous examples, where the analogous function was not even absolutely continuous (nor differentiable). This happens due to presence of the change of variable which has a quadratic branch point at and thus causes doubling of periods and widths of strips of holomorphy.
General . The analysis is essentially the same as for , and we will omit a discussion. Let us simply point out that the generating function satisfies
which gives
Hence, we have again a 2-Mahler function of degree . From this we get
as the exponent of the subexponential factor.
2BVI. Conclusion and goal
The primary objective of this paper is to study a similar but more complicated problem of precisely estimating the number of indecomposable summands of . Unlike the study of the length, which grows faster than the number of indecomposable summands due to the non semisimple nature of the category of representations of , this problem introduces subtleties. For example, since the 2-Mahler equation of degree it leads to is inhomogeneous, the generating function is not a single product but rather a sum of products. Still, the asymptotic behavior of the sequences and functions we consider ends up being very similar to the above examples. There will be an especial similarity with the example in this subsection (e.g. length of ). Namely, we also observe doubling of periods and strip widths, resulting in a very high degree of regularity of the sequence of interest.
We will however abstain from exploring specific fractals within this context.
3. Generating function
We fix a prime . Our first goal is to explicitly describe the generating function for the sequence as in Equation 1B.1. All objects will depend on , but to keep notation light we usually omit from notation.
3A. The function
Let and be formal variables. By expansion, we have a ring inclusion
Moreover, the above restricts to inclusions
The image of the latter inclusion consists precisely of those rational functions in invariant under . We will use this several times tacitly below.
We are now ready to study the generating function for the sequence . It is a bit more convenient to shift the generating function for as follows. Let
We will also regard as a holomorphic function with domain which vanishes at infinity.
It will be convenient for formal manipulations to focus on . We therefore set
Of course, by construction, we can also interpret as a holomorphic function (the singularity of this function that will be important later is now at ) on
(3A.1) |
Indeed, direct manipulation shows that is the set of complex numbers with
yielding the intersection and joint exterior of two disks, as displayed in Equation 3A.1.
Our main result of this section is:
Theorem \theTheorem.
We have the following explicit formula for :
Remark \theRemark.
As pointed out to us by Henning Haahr Andersen, a finer version (with a quite different formulation) of Section 3A has appeared in [Erd95].
The remainder of this section is devoted to the proof of Section 3A.
Example \theExample.
Example \theExample.
We can also consider the generating function for a field of characteristic zero for the numbers , see Section 1B. Then we find
This is well-known and can be derived from [OEI23, A001405].
3B. Recursion relations
Recall the category of tilting representations for , see e.g. [Don93] and [Rin91] (additional details can be found in [AST18]), or, using the identification with the Temperley–Lieb calculus, [And19] or [TW21]. All the terminology and facts about tilting representations that we use below can be found or derived from these references. We denote the category by .
We identify the weight lattice of with and the dominant weights with . Let be the split Grothendieck ring of . Then has a -basis , where is the indecomposable tilting representation of highest weight . We then have ring isomorphisms
where is the subring of Laurent polynomials symmetric under , and associates the formal character to a representation. In particular, the composed isomorphism identifies and . We will freely use both isomorphisms and switch notation accordingly.
Let be the group homomorphism
Hence, we have .
By Donkin’s tensor product formula, for the operation maps indecomposable tilting representations to indecomposable tilting representations; concretely:
Here is the Frobenius twist of . In particular, we have
where is the endomorphism of determined by . Consequently, we have
(3B.1) |
for all .
We will interpret (3B.1) as recursion relations to compute . To this end, we need to know the following few characters: it is well-known and easy to compute that
and, for ,
(3B.2) |
We can reformulate this in terms of the Chebyshev polynomials of the first kind, (not to be confused with the notation for tilting modules) and of the second kind, . Indeed, they can be defined via
(3B.3) |
We then find:
Example \theExample.
For , the two equations become
The first relation allows one to compute in terms of lower cases and the second relation allows one to compute . Concretely, one obtains
This gives a very efficient way of computing the numbers .
Let be the group endomorphism of given by
(3B.4) |
where is a primitive complex th root of unity. In particular, we have
Lemma \theLemma.
The function satisfies
Proof.
For convenience, we extend the definition of to
-linearly. Then we can write
Using (3B.1) and (3B.2), we can then conclude that
(3B.5) |
where and for . We now complete the proof of the first displayed equation, corresponding to , the other cases being similar.
Writing
we find that
Consequently, we find
Comparing this expression with (3B.5) indeed proves the first equation of the lemma. ∎
3C. Functional equation
The following proposition shows that is a 2-Mahler function of degree .
Proposition \theProposition.
We have
where is explicitly given by
For , is also determined as the rational function in such that is the Padé approximant of order of .
Example \theExample.
Consider the case . The two equations in Section 3B are
Taking a linear combination of the two equations neutralising the terms in yields
This is equivalent to the functional equation in Section 3C specialized at . The proof of the case given below is a refinement of this argument.
Proof of Section 3C.
We can obtain the functional equation by taking an appropriate linear combination of the equalities in Section 3B. Indeed, we use the interpretation of in (3B.4), and we will sum with appropriate -dependent coefficients to make all terms containing for cancel. We can start by adding up the equalities in Section 3B for with coefficients so that the resulting coefficients of yield a value independent of .
So we must have
(3C.1) |
for all . Consider
By (3C.1) and its complex conjugate, , for all . It follows that, up to scaling,
It follows that, setting , we can choose:
Summing over the equations as intended we thus get
So multiplying the first equation, , by , and adding it to the above equation, we obtain
This can be written as
subsequently as
and finally as the expression in the theorem.
The values for are equal to the corresponding numbers in characteristic zero. Consequently, we have
If is the Chebyshev polynomial of the first kind, then we set
In other words:
The functional equation then implies
Since is of degree (as a polynomial in ), we have (as functions in )
Rewriting the functional equation, we thus find
It thus follows that is the Padé approximant of order of
if and is a rational function in of degree .
It remains to verify the final statement. For consider therefore the expansion
The left factor on the last line is the inverse of a polynomial in of degree , while the right factor is a polynomial in of degree . In particular, is a rational function in of degree , concluding the proof. ∎
3D. Conclusion
Proof of Section 3A.
We can now employ the functional equation from Section 3C to derive the closed expression for in Section 3A. Set
So
(3D.1) |
Thus setting , Section 3C can be rewritten as
So we get
or equivalently
The conclusion of Section 3A then follows from substituting (iterations of) Equation 3D.1. ∎
4. Asymptotics of the generating function
Recall that is a fixed prime. In order to understand the sequence from Equation 1B.1 better, we focus on the generating function as approaches the radius of convergence , see Section 3, as usual in the theory of asymptotics of generating functions.
4A. Asymptotics
Concretely, we will focus on the behavior for of , viewed as a smooth function
This is well-defined, as lies in from Equation 3A.1. This can be seen directly from the formula , or the plot Equation 3A.1. We will now prove one of our main results:
Theorem \theTheorem.
We have
Here is real analytic, and bounded away from and .
Proof.
Recall from Section 3C that we have
A calculation gives
This is a -Mahler function of degree , cf. Equation 2B.1, meaning it is of the form
Let be a variable. After clearing denominators so that the are polynomials, the so-called characteristic polynomial (as recalled e.g. in [BC17]) of a Mahler functional equation as in Equation 2B.2 is
This in our example is
which has a root at . At the relevant singularity we get
which is the eigenvalue of the Mahler function . Now the classical theory of Mahler functions, see e.g. [BC17, Theorem 1], implies that the log with base the degree of the Mahler equation of this eigenvalue is minus the exponent of in the asymptotic expansion. The remaining parts of the theorem follow also from [BC17, Theorem 1]. ∎
In the reminder of this section we make Section 4A more explicit.
4B. The oscillating factor
We define the following function on , rescaling :
Lemma \theLemma.
As a function on we have:
Proof.
Directly from Section 3A. ∎
We will now show that there exists a function with , defined pointwise, satisfying . To this end, consider the product in given by
Define the power series
Proposition \theProposition.
For , we have
Moreover, is a real analytic and oscillatory term, , and is bounded away from and .
Remark \theRemark.
The function in Section 4B is a rescaling of from Section 4A, so describes the oscillation of .
Proof of Section 4B.
It is easy to see that is well-defined for : Firstly, the factors in the expression converge to rapidly, and this implies that itself converges to some number in , and we can assume that is equal to . Second, the left term in the sum goes to and takes values in for negative , so the negative part of the sum converges. Finally, for positive the summands go to rapidly and the sum also converges.
As in the previous paragraph, converges to for . We then obtain
The claim follows. ∎
Proposition \theProposition.
The series converges absolutely and uniformly on compact sets in the region but has a dense set of singularities on the imaginary axis.
Proof.
This follows as for proven in Section 4C below. Details are omitted. ∎
4C. Example for
For , the expression for simplifies to
In particular, we have
Lemma \theLemma.
The product converges absolutely and uniformly on compact sets (not containing zeros and poles) to a meromorphic function of .
Proof.
This holds since exponentially fast as . ∎
The poles of factors in are solutions of the equation , i.e.,
On the other hand, zeros occur when nontrivial cube roots of , so
and they have multiplicities (the number of factors in ).
Proposition \theProposition.
The series converges absolutely and uniformly on compact sets in the region but has a dense set of singularities on the imaginary axis.
Proof.
This holds by the above discussion. ∎
5. Monotonicity
In this section we address the monotonicity of our main sequence.
5A. Neighboring values of
We will now prove:
Theorem \theTheorem.
We have , or equivalently , for all .
Proof.
The proof will occupy this section, and is split into a few lemmas. For the first lemma up next, let for be a sequence of Laurent polynomials such that, for some , we have
(5A.1) |
Note that , where we use as in Section 3A. Now we will prove that actually .
Lemma \theLemma.
If is a positive power series in , then the same is true for for all .
Proof.
We have
So the statement follows by induction on . ∎
For any and an integer , let us define
Similarly, define
Lemma \theLemma.
-
(a)
The function has positive Taylor coefficients in .
-
(b)
The function has positive Taylor coefficients in .
Proof.
(a). If is an integer, let , and if is an honest half integer, let . Then satisfy (5A.1) with or . Moreover, in the integer case and in the non-integer case.
Thus, the result follows from Section 5A, as we can write
(b). Let . Then satisfy (5A.1), and . Thus, as in part (a), the result follows from Section 5A. ∎
We let
Moreover, we consider the renormalization
Note that , so that and are not positive. But we prove:
Lemma \theLemma.
The Taylor coefficients in of the function satisfy for .
Example \theExample.
Proof of Section 5A.
Recall the 2-Mahler coefficients from Section 3C, and also the functional equation for given in the proof of Section 3C, namely:
Here is the th Chebyshev polynomial of the first kind with , as recalled in Equation 3B.3. Set also .
Generalzing the calculation in Section 5B below, we get the following expression (note that which gives a rescaling of the above functional equation):
We let , with the last terms being either or , depending on the parity. A calculation gives
To see this observe that this is equivalent to . This in turn follows by setting and we get
We thus get
We get:
We rewrite this as
where
Recall that the polynomial has real roots. Furthermore, or depends on . Both observations together imply that we have that is a positive series in . The same applies to . So it suffices to check that the series and are positive. The next two lemmas imply these facts.
Lemma \theLemma.
The function has positive Taylor coefficients in . The same holds for .
Proof.
For we have
which is manifestly positive.
That is positive is then immediate from . ∎
Lemma \theLemma.
The function has positive Taylor coefficients in .
Proof.
A calculation shows that
The , as indicated above, are positive for . Indeed, , , , are positive by Section 5A and is positive by Section 5A and Section 5A. ∎
Section 5A implies that is positive, while is positive by Section 5A. This finishes the proof of Section 5A. ∎
Taking all together yields Section 5A. ∎
5B. Example for
For the calculation in Section 5A is rather straightforward. In this case Section 3C implies the functional equation
This gives a recursion of the coefficients , namely:
This recursion has positive coefficients, hence, Section 5A follows.
6. The main theorem
We now prove Section 1B after an auxiliary lemma for which we use Equation 1B.3.
6A. A Tauberian lemma
The following type of result, often called Tauberian theory, is standard and just reformulated to suit our needs, see for instance [Tit58, §7.53] or [BGT89, Theorem 2.10.2] for related results.
Lemma \theLemma.
-
(a)
Consider a sequence with , for which the series has radius of convergence . If, for some , we have,
for some , then:
-
(b)
Assume that, additionally to (a), for some and we have
for all . Then:
Proof.
Claim (a) We set .
We first prove the upper bound on . For every we have
for some . The function attains its minimum at (which is larger than for sufficiently large), allowing us to reformulate the above inequality as
The latter then implies that for some , we have .
Deriving the inequality in the other direction is more subtle. However, as explained in [Dus20], this (in fact, both inequalities) is a special case of the de Haan–Stadtmüller theorem, see [BGT89, Theorem 2.10.2].
Claim (b) By the conclusion of part (a), we know that for some
We will use this freely.
We again start with the upper bound. For all we find from monotonicity that
From this we can derive for all for .
For the lower bound we consider the case first (in which case we can take and the second inequality in (b) is trivial). For any we have
Since , we can choose for which . For , we can rewrite the above inequality as
By construction the factor in front of is positive and we are done.
In case we can similarly show that
(6A.1) |
for some . On the other hand, we have
In particular, we can take so that
which together with (6A.1) concludes the proof. ∎
Remark \theRemark.
The conclusion in Section 6A.(b) does not follow without the additional assumption in case . Indeed, it suffices to take the sequence
Then and , but .
We in fact already had an example of this type: in Section 2BI the Cantor set sequence satisfies
for . However, .
6B. Conclusion
We are ready to prove our main result:
Proof of Section 1B.
Set for . Then Section 5A implies that . We also know that , since each indecomposable summand in is responsible for at least one in . Hence, , so the sequence satisfies all conditions in parts (a) and (b) of Section 6A with .
For the purpose of this proof we will write to mean for an arbitrary . Recall that
Section 4A implies that
Recall further that
Solving for in the region then yields
Since the last factor is bounded on away from , we find
Section 6A.(b) then implies
and we are done. ∎
7. Additional results in characteristic two
Throughout this section let . This restriction is mostly for convenience: with some work the statements and proofs below generalize to all primes.
7A. Neighboring values of in characteristic two
We will now strengthen Section 5A, where we can focus on difference two for the indexes since for . The auxiliary sequence that we use is , for .
Proposition \theProposition.
We have and .
Since , Section 7A shows that .
Proof of Section 7A.
We start with an analysis of .
Lemma \theLemma.
We have for all .
Proof.
By Section 3C, we have . So, if , then we get
Hence, we get for that:
This gives a positive recursion for , which implies the statement. ∎
The numbers are nonnegative by Section 5A. Moreover, Section 7A implies that which gives
This yields and . Putting these together gives , and iterating this procedure then gives
We also have, say for the even values,
for some , where the final inequality is Section 1B. Thus,
for some , which, by rewriting, proves the statement. ∎
7B. The main theorem revisited
The following generalizes Section 1B:
Proposition \theProposition.
Equation 1D.1 is true.
Proof.
Let . Note that this ignores the odd values since runs only over the even values of , but since for all this does not play a role below and just simplifies the notation.
We observe that , so that Section 1B implies that we can sandwich both, and , at the same time:
However, this does not imply the result yet, we need to know a bit more about the sequence . Firstly, it follows from Section 7A that there is a constant such that
(7B.1) |
Also and, by Section 3B, for we have
Thus, when running over even values, we get
(7B.2) |
Note also that, for , we have
So using this, combined with (7B.1) and (7B.2), we get
Hence, for all there exists a limit
and is continuous on with , and bounded away from and . ∎
8. Fourier coefficients
The functions that we have met above are oscillating, and in this section we analyze their Fourier coefficients.
8A. An Abelian lemma
Tauberian theory as in Section 6A, roughly speaking, says that given a certain behavior of a generating function, we get a certain behavior of the associated sequence. There is an “inverse” to Tauberian theory often called Abelian theory.
Below we will need the following well-known result from Abelian theory:
Lemma \theLemma.
Assume that we have two functions and which converge for and diverge for . Then
Proof.
This is for example explained at the beginning of [Tit58, Section 7.5]. ∎
8B. Some generalities on asymptotics of Fourier coefficients
Let and . Suppose we have a sequence such that we have asymptotically
for some continuous -periodic function with . The associated generating function is the series
Lemma \theLemma.
The series absolutely converges for with singularity at .
Proof.
This follows since . ∎
As before, let denote the gamma function. Recall that the Fourier coefficient formula of a -periodic function (so that the integral expression up next makes sense) are given by , for . That is, for .
We get the following asymptotic of :
Proposition \theProposition.
Retain the assumptions above, and denote the Fourier coefficients of by . We have
where is the -periodic function given by
Proof.
For the argument below we note that is the relevant singularity of , and we need to analyze the growth rate of for .
Recalling that , we use Section 8A and get
We rewrite this as
So setting and using the periodicity of , namely , we get
where the inner sum is over (not necessarily reduced) fractions with denominator . Note that the main contribution to the asymptotics comes from large , as . Hence, we may replace the inner sum by an integral times (reciprocal length of step), and the range of summation for the outer sum by . After doing this we get:
Making the substitution allows us to rewrite the above as
Substituting back to , the above yields
Replacing the dummy variable by yields
We can thus define the 1-periodic function
That this periodic function corresponds to the one in the proposition follows easily from the definition of the Gamma function. ∎
Theorem \theTheorem.
Proof.
By Section 8B and Fourier inversion. ∎
8C. Analysis of the Fourier coefficients of the generating function for
Assume that Equation 1D.1 holds. (For this assumption is true by Section 7.) We now apply the results in Section 8B to our example of in characteristic . That is, we look at for the in Section 8B.
Consider the generating function of :
The relevant singularity is at . Below we will also use the generating function in defined mutatis mutandis as in Section 3A.
Moreover, associated to we can define similarly as from Section 4B but defined now in , and we get the same type of statements for as for .
Recall from Equation 1B.2. Setting (which we expect to be correct based on numerical data) and , we get:
Proposition \theProposition.
Proof.
Recall from Section 4B that . We can write this as . Then Section 8B as well as Section 8B apply. ∎
9. Numerical data
All of the below can be found on the GitHub page associated to this project [CEOT24], where the reader can find code that can be run online. For convenience, we list a few numerical examples here as well.
Remark \theRemark.
All plots below are logplots (also called semi-log plots). This means that the plots have one axis (the y-axis for us) on a logarithmic scale, the other on a linear scale.
9A. The main sequence
Below we will always use with meaning that the characteristic is large compared to the considered range, which leads to the same data as in characteristic zero.
We first give tables for for :
1 | 1 | 1 | 3 | 3 | 9 | 9 | 29 | 29 | 99 | 99 | 351 | 351 | 1273 | 1273 | 4679 | |
1 | 1 | 2 | 2 | 5 | 6 | 15 | 21 | 50 | 77 | 176 | 286 | 637 | 1066 | 2340 | 3978 | |
1 | 1 | 2 | 3 | 6 | 9 | 19 | 28 | 62 | 91 | 208 | 308 | 716 | 1079 | 2522 | 3886 | |
1 | 1 | 2 | 3 | 6 | 10 | 20 | 34 | 69 | 117 | 242 | 407 | 858 | 1431 | 3069 | 5085 | |
1 | 1 | 2 | 3 | 6 | 10 | 20 | 35 | 70 | 126 | 252 | 462 | 924 | 1716 | 3432 | 6435 |
.
For completeness, we give here the for (in order top to bottom, excluding where the sequence is [OEI23, A001405]) in a copy-able fashion:
.
.
.
.
Here is one picture to compare their growth, including the case :
Note that the sequences get a bit more regular if one runs over the, say, even values. This is in particular true in characteristic where . Below we will strategically sometimes only illustrates the even values.
Next, the logplots of compared with , and of compared with are:
9B. The length sequence
Let us now also give here the for (again in order top for to bottom for , excluding where the sequence is [OEI23, A001405]) so that it is copy-able:
.
.
.
.
As before for , here is one picture to compare their growth, including the case :
Note the following comparison of the plot to the plot from above: for , the grows a faster for and slower for , and vice versa for the .
9C. The generating function
Continuing with Section 3A, recall from Equation 1B.2 that . We compare , with the sum cut-off at , with :
The growth rates towards the singularity (left side of the picture) are almost the same and this is crucial in Section 6.
Recall from Section 4B that
We now give formulas to compute the Fourier coefficients of , and thus of , (fairly) efficiently. It turns out that is moderate but are tiny for , which causes the behavior of as in the next example, plotted on :
This looks like the sine curve because is much smaller than as we will see momentarily. To wrap-up, for one can show that
|
Let denote the Hurwitz zeta function. Consider the following reparametrization:
Recalling that , we get
The Hurwitz zeta function is quite easy to compute, so this gives a fairly efficient way to compute the Fourier coefficients .
A bit of work shows that, for general , we have:
|
Moreover, using the classical formulas for the Hurwitz zeta function one can show that is nonzero. Furthermore, let
One can also show that , for fixed , is obtained at and . Indeed, the function behaves as follows when varying and , respectively, while keeping fixed:
Here we have divided by to highlight the bound which is quite crude:
9D. Numerical data for odds and ends
Recall that Section 1B shows that, for some , we have
One can probably take and , as illustrated here for :
Moreover, Section 5A shows that we have
One could conjecture that , as illustrated here for where we know this is true by Section 7A:
Finally, with is an indecomposable tilting -representation unless . For we get for the numbers:
One could aim for a statement of the form
where . Indeed, we get the following logplots:
As one can see, we expect some nontrivial scalar to make the curves fit better. Hence, these plots might indicate an analog of Section 1B for instead of .
Appendix A Monoidal subcategories of finite tensor categories
We use the terminology of tensor categories from [EGNO15]. Recall that an object of a tensor category is a generator if any object of is isomorphic to a subquotient of a direct sum of tensor products of and its (iterated) duals.
For Section 2A, we need to show that in case is finite (meaning that has enough projective objects and finitely many simple objects), all projective objects appear as direct summands of the tensor powers of . This follows from Appendix A.(d) below and the fact that projective objects in a tensor category are also injective; see [EGNO15, Proposition 6.1.3]. Note that the theorem is a generalization of the known case of fusion categories in [DGNO10, Corollary F.7]. It can also be interpreted as a generalization of the Burnside–Brauer–Steinberg theorem for Hopf algebras; see, e.g. [PQ95], or a generalization of the main result of [BK72].
Theorem \theTheorem.
The following properties hold for any tensor category with finitely many isomorphism classes of simple objects:
-
(a)
Any full additive subcategory that is closed under taking subquotients and internal tensor products is a tensor subcategory (i.e. rigid).
-
(b)
For any , the left and right dual objects and appear as subquotients of direct sums of for .
-
(c)
For any simple object , the left and right dual objects and appear as subquotients of for some .
-
(d)
If is a generator, then every object in is a subquotient of direct sums of tensor powers of .
Proof.
Firstly, we can observe that part (a) implies part (b) and that (b) implies (d).
Now we prove that (c) actually implies (a). For this, let be a subcategory as in (a), and we assume that property (c) is true for . To show that is rigid, we show that every object is rigid. We prove this by induction on the length of . The base follows from our assumption (c).
To justify the induction step, let and include into a short exact sequence
where are nonzero. Note that and are rigid in by induction, and is defined by an element . Now, in general, given an element for objects with duals and , we have an element which defines an element obtained by composing with evaluations and coevaluations. The , the extension of and defined by is the left dual of , and thus belongs to . The right dual can be found in a similar way. Thus, is rigid.
By the above, we have (c)(a)(b)(d), and it now suffices to prove property (c). This will be done below after some preparation.
The following lemma can be seen as being part of the Frobenius–Perron theorem.
Lemma \theLemma.
Let be a nonzero square matrix with nonnegative entries and be row and column vectors with strictly positive entries such that , for some . Then and is completely reducible (conjugate by a permutation matrix to a direct sum of irreducible matrices).
Proof.
We have , so . It remains to show that if is reducible, then it is decomposable. If is reducible, then after conjugating by a permutation matrix, we have . So, if , then we have
Multiplying the second equation by , the first equation by , and subtracting, we get , which means , so is decomposable, as claimed. ∎
Now, let be a transitive unital -ring of finite rank with basis . For with , let be the matrix of left multiplication by on in the basis .
Lemma \theLemma.
-
(i)
If , then is completely reducible.
-
(ii)
Suppose that and that the -decomposition of contains . Then there exists such that the -decomposition of contains .
Proof.
(i). Let be the Frobenius–Perron dimension. Since is a character, , so the column vector is a right eigenvector of . Also, by [EGNO15, Proposition 3.3.6(2)], there exists a left (row) eigenvector of with positive entries. Thus, by Appendix A, is completely reducible.
(ii). Consider the directed graph with vertex set and edge present if and only if the -decomposition of contains . Let be the connected component of that contains . Since the -decomposition of contains , . Thus, by (i), there is an oriented path from to , i.e., there exists such that the -decomposition of contains , as desired. ∎
To complete the proof of Appendix A, it now suffices to observe that Appendix A.(c) follows from applying Appendix A(ii) to the Grothendieck ring , with the basis of simple objects, and or . ∎
Remark \theRemark.
The assumption that has finitely many simple objects in Appendix A cannot be dropped (even for categories with enough projective objects). For instance, the statement is not true if is the representation category of the multiplicative group and is a faithful one dimensional representation.
References
- [And04] H.H. Andersen. Cells in affine Weyl groups and tilting modules. In Representation theory of algebraic groups and quantum groups, volume 40 of Adv. Stud. Pure Math., pages 1–16. Math. Soc. Japan, Tokyo, 2004. doi:10.2969/aspm/04010001.
- [And19] H.H. Andersen. Simple modules for Temperley–Lieb algebras and related algebras. J. Algebra, 520:276–308, 2019. URL: https://arxiv.org/abs/1709.00248, doi:10.1016/j.jalgebra.2018.10.035.
- [AST18] H.H. Andersen, C. Stroppel, and D. Tubbenhauer. Cellular structures using -tilting modules. Pacific J. Math., 292(1):21–59, 2018. URL: https://arxiv.org/abs/1503.00224, doi:10.2140/pjm.2018.292.21.
- [BC17] J.P. Bell and M. Coons. Transcendence tests for Mahler functions. Proc. Amer. Math. Soc., 145(3):1061–1070, 2017. URL: https://arxiv.org/abs/1511.07530, doi:10.1090/proc/13297.
- [Bia93] P. Biane. Estimation asymptotique des multiplicités dans les puissances tensorielles d’un -module. C. R. Acad. Sci. Paris Sér. I Math., 316(8):849–852, 1993.
- [BGT89] N.H. Bingham, C.M. Goldie, and J.L. Teugels. Regular variation, volume 27 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, 1989.
- [BK72] R.M. Bryant and L.G. Kovács. Tensor products of representations of finite groups. Bull. London Math. Soc., 4:133–135, 1972. doi:10.1112/blms/4.2.133.
- [BLS19] G. Burrull, N. Libedinsky, and P. Sentinelli. p-Jones–Wenzl idempotents. Adv. Math., 352(20):246–264, 2019. URL: https://arxiv.org/abs/1902.00305, doi:10.1016/j.aim.2019.06.005.
- [CE21] M. Coons and J. Evans. A sequential view of self-similar measures; or, what the ghosts of Mahler and Cantor can teach us about dimension. J. Integer Seq., 24(2):Art. 21.2.5, 10, 2021. URL: https://arxiv.org/abs/2011.10722.
- [CEO24] K. Coulembier, P. Etingof, and V. Ostrik. Asymptotic properties of tensor powers in symmetric tensor categories. Pure Appl. Math. Q. 20 (2024), no. 3, 1141–1179. URL: https://arxiv.org/abs/2301.09804, doi:10.4310/pamq.2024.v20.n3.a4.
- [CEOT24] K. Coulembier, P. Etingof, V. Ostrik, and D. Tubbenhauer. Code and more for the paper Fractal behavior of tensor powers of the two dimensional space in prime characteristic. 2024. https://github.com/dtubbenhauer/sl2-charp.
- [COT24] K. Coulembier, V. Ostrik, and D. Tubbenhauer. Growth rates of the number of indecomposable summands in tensor powers. Algebr. Represent. Theory 27 (2024), no. 2, 1033–1062. URL: https://arxiv.org/abs/2301.00885, doi:10.1007/s10468-023-10245-7.
- [CW24] K. Coulembier and G. Williamson. Perfecting group schemes. J. Inst. Math. Jussieu 23 (2024), no. 6, 2549–2591. URL: https://arxiv.org/abs/2206.05867, doi:10.1017/S1474748024000033.
- [Del90] P. Deligne. Catégories tannakiennes. In The Grothendieck Festschrift, Vol. II, volume 87 of Progr. Math., pages 111–195. Birkhäuser Boston, Boston, MA, 1990.
- [Don93] S. Donkin. On tilting modules for algebraic groups. Math. Z., 212(1):39–60, 1993. doi:10.1007/BF02571640.
- [DGNO10] V. Drinfeld and S. Gelaki and D. Nikshych and O. Victor. On tilting modules for algebraic groups. Selecta Math. (N.S.), 16 (2010), no. 1, 1–119. URL: https://arxiv.org/abs/0906.0620, doi:10.1007/s00029-010-0017-z.
- [Dus20] M. Dus. Weak version of Karamata’s Tauberian theorem. MathOverflow, 2020. Version: 2020-Mar-03; author’s MathOverflow link https://mathoverflow.net/users/111917/m-dus. URL: https://mathoverflow.net/q/353943.
- [EGNO15] P. Etingof, S. Gelaki, D. Nikshych, and V. Ostrik. Tensor categories, volume 205 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2015. doi:10.1090/surv/205.
- [Erd95] K. Erdmann. Tensor products and dimensions of simple modules for symmetric groups. Manuscripta Math. 88 (1995), no. 3, 357–386. doi:10.1007/BF02567828.
- [FS09] P. Flajolet and R. Sedgewick. Analytic combinatorics. Cambridge University Press, Cambridge, 2009. doi:10.1017/CBO9780511801655.
- [Hab80] W.J. Haboush. Central differential operators on split semisimple groups over fields of positive characteristic. Séminaire d’Algèbre Paul Dubreil et Marie-Paule Malliavin, 32ème année (Paris, 1979), pp. 35–85, Lecture Notes in Math., 795, Springer, Berlin, 1980.
- [KST24] M. Khovanov, M. Sitaraman, and D. Tubbenhauer. Monoidal categories, representation gap and cryptography. Trans. Amer. Math. Soc. Ser. B, 11:329–395, 2024. URL: https://arxiv.org/abs/2201.01805, doi:10.1090/btran/151.
- [LTV23] A. Lacabanne, D. Tubbenhauer, and P. Vaz. Asymptotics in finite monoidal categories. Proc. Amer. Math. Soc. Ser. B, 10:398–412, 2023. URL: https://arxiv.org/abs/2307.03044, doi:10.1090/bproc/198.
- [LTV24] A. Lacabanne, D. Tubbenhauer, and P. Vaz. Asymptotics in infinite monoidal categories. 2024. URL: https://arxiv.org/abs/2404.09513.
- [LPRS24] A. Lachowska, O. Postnova, N. Reshetikhin, and D. Solovyev. Tensor powers of vector representation of at even roots of unity. 2024. URL: https://arxiv.org/abs/2404.03933.
- [La24] M. Larsen. Bounds for SL2-indecomposables in tensor powers of the natural representation in characteristic 2. 2024. URL: https://arxiv.org/abs/2405.16015.
- [LW18] G. Lusztig and G. Williamson. Billiards and tilting characters for . SIGMA Symmetry Integrability Geom. Methods Appl., 14:Paper No. 015, 22, 2018. URL: https://arxiv.org/abs/1703.05898, doi:10.3842/SIGMA.2018.015.
- [Mis20] M. Mishna. Analytic combinatorics: a multidimensional approach. Discrete Mathematics and its Applications (Boca Raton). CRC Press, Boca Raton, FL, [2020] ©2020.
- [Nis96] K. Nishioka. Mahler functions and transcendence, volume 1631 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 1996. doi:10.1007/BFb0093672.
- [OEI23] OEIS Foundation Inc. The On-Line Encyclopedia of Integer Sequences, 2023. Published electronically at http://oeis.org.
- [PQ95] D.S. Passman and D. Quinn. Burnside’s theorem for Hopf algebras. Proc. Amer. Math. Soc., 123(2):327–333, 1995. doi:10.2307/2160884.
- [Rin91] C.M. Ringel. The category of modules with good filtrations over a quasi-hereditary algebra has almost split sequences. Math. Z., 208(2):209–223, 1991. doi:10.1007/BF02571521.
- [Spe23] R.A. Spencer. The modular Temperley–Lieb algebra. Rocky Mountain J. Math., 53(1):177–208, 2023. URL: https://arxiv.org/abs/2011.01328, doi:10.1216/rmj.2023.53.177.
- [Tit58] E.C. Titchmarsh. The theory of functions. Oxford University Press, Oxford, 1958. Reprint of the second (1939) edition.
- [TW21] D. Tubbenhauer and P. Wedrich. Quivers for tilting modules. Represent. Theory, 25:440–480, 2021. URL: https://arxiv.org/abs/1907.11560, doi:10.1090/ert/569.
- [STWZ23] L. Sutton, D. Tubbenhauer, P. Wedrich, and J. Zhu. SL2 tilting modules in the mixed case. Selecta Math. (N.S.), 29(3):39, 2023. URL: https://arxiv.org/abs/2105.07724, doi:10.1007/s00029-023-00835-0.