On the rate of convergence of continued fraction statistics of random rationals
Abstract.
We show that the statistics of the continued fraction expansion of a randomly chosen rational in the unit interval, with a fixed large denominator , approaches the Gauss-Kuzmin statistics with polynomial rate in . This improves on previous results giving the convergence without rate. As an application of this effective rate of convergence, we show that the statistics of a randomly chosen rational in the unit interval, with a fixed large denominator and prime numerator, also approaches the Gauss-Kuzmin statistics. Our results are obtained as applications of improved non-escape of mass and equidistribution statements for the geodesic flow on the space .
1. Introduction
In this article we wish to revisit some of the results proved in [DS18] and strengthen them. The main novel observation in the current article is that one of the two assumptions appearing in the main theorems of [DS18] follows from the other and hence can be dropped. We then add to this by presenting a new arithmetic application of the stronger result.
1.1. Background and the main application
It is well known since ancient times that a real number admits a continued fraction expansion (abbreviated hereafter as c.f.e); namely, there are positive integers such that
In fact, to be more precise, irrational ’s correspond bijectively to infinite sequences , where the above equation is interpreted as a limit, and rational ’s correspond to finite sequences of positive integers, where here the correspondence is not - but almost so: each rational can be written in exactly two ways,
where and . As usual, we abbreviate and write when is irrational and when is rational. As a convention we will always use the shorter expansion for rationals, which does not end with . The length of the c.f.e of a rational number is defined by
(1.1) |
Our discussion will revolve around the continued fraction expansions of rationals with large denominators, but to motivate the phenomena discussed it is actually more natural to start with irrationals.
It is well known (see for example the introduction of [AS18]) that given any finite string of positive integers , for Lebesgue almost any irrational , the asymptotic frequency of appearance of in the c.f.e of exists and is independent of . That is, the limit
(1.2) |
exists and is independent of , as the notation suggests (in fact equals where is a certain interval in , see Remark 3.2). For example, for a randomly chosen , the asymptotic density of appearance of ’s in its c.f.e is with probability .
A similar density could be defined for rationals: if then we can define
In [DS18] the authors investigated the densities of rationals in reduced form for large denominators , as well as the length of the c.f.e of such rationals. In particular, the following result is proved.
Theorem 1.1 ([DS18]).
For any , let be a random rational with denominator and numerator co-prime to chosen uniformly (namely, one picks randomly and uniformly from the possible numerators co-prime to ). Then for any , any string , and any , we have
-
(1)
-
(2)
In this paper, we are able to improve this result, and show that not only the probability in Theorem 1.1 converges to zero, but it does so at a polynomial rate.
Theorem 1.2.
For any , let be a random rational with denominator and numerator co-prime to chosen uniformly. Then, for any , any string , and any , there exist and such that for all large enough:
-
(1)
-
(2)
Our methods do not provide any estimate of the values and in Theorem 1.2, only their existence. It would be interesting to find out how fast do these probabilities must decay, and in particular the dependence of these powers on .
Theorem 1.2 is in fact equivalent to the following generalization of Theorem 1.1. We show that the probability for the c.f.e of to ‘behave badly’ (in the sense of Theorem 1.1) approaches , even if we choose uniformly at random from a set of smaller size with , instead of choosing from . Here and throughout this paper, we identify between and the set .
Theorem 1.3.
For any , let so that , and let be a random rational with denominator and numerator chosen uniformly (we denote this probability measure over by ). Then, for any , any string , and any , we have
-
(1)
-
(2)
As an application, we deduce the following prime numerator version of Theorem 1.1. It follows from the prime number theorem in combination with Theorem 1.3.
Theorem 1.4.
For any , let be a random rational with denominator and prime numerator co-prime to chosen uniformly. Then for any , any string , and any , we have
-
(1)
-
(2)
Related questions regarding the distribution of the coefficients in continued fractions of random rationals, and particularly the expected length of the expansion, were previously discussed in the literature. In [Hei69], Heilbronn computed the asymptotics of the average length over , also giving an error term. This error term was improved by Porter [Por75] and then further improved by Ustinov [Ust08].
Other works considered the length question on different sets. In [Bi05], Bykovski considered for any the average length for the c.f.e of , over all not necessarily coprime to . Others considered a larger probability space, comprised of all pairs where is coprime to and for some . On this space, Dixon [Dix70] showed a large deviation result, namely bounded the probability that deviates from the expected value by a certain amount. This result was later improved by Hensely [Hen94]. Furthermore, in [BV05] Baladi and Vallée consider a central limit theorem on this space. In comparison with these works, we consider too a large deviation question, but on a significantly smaller probability space where the denominator is considered fixed. Furthermore, we also consider the distribution of words in the c.f.e expansions rather than only its length.
1.2. The diagonal flow and the main result
We start with the setting in which our mathematical discussion will take place, namely the space of unimodular lattices in Euclidean plane. We will use similar notation to [DS18]. Let , , , and let denote the identity coset. For a point we let denote the probability measure supported on the singleton . Consider the diagonal flow
The unstable horospherical subgroup of with respect to is given by
For a given set denote
(1.3) |
that is, is the probability measure given by the continuous average in the diagonal direction on the segment of the discrete counting probability measure on the points . For an illustration of the geodesic segments on which we average, see Figure 1 and the discussion before it. Recall that we say that a sequence of probability measures on does not exhibit escape of mass if any weak* limit of it is a probability measure. The following is one of the main results of [DS18].
Theorem 1.5.
[DS18, Theorem 1.7] Let be subsets such that
-
(1)
,
-
(2)
the sequence of measures does not exhibit escape of mass.
Then , where is the Haar probability measure on .
The current paper concerns itself with proving that the second item in Theorem 1.5 is superfluous since it actually follows from the first item. Our main result is the following bound on the possible escape of mass for divergent orbits.
Theorem 1.6.
Let . Suppose that are given so that
Then any weak∗ limit of the sequence satisfies
Corollary 1.7.
Let be subsets satisfying , then .
Remark 1.8.
To illustrate the trajectories , we show in Figure 1 their projection to the upper half plane , which is identified with the quotient , where . The trajectories are shown in the standard fundamental domain, given by points with and . As discussed in [DS18], at times , these trajectories start and end at height in the upper half plane, and for the trajectories diverge directly into the cusp. Hence the -portion of the geodesic trajectory is the interesting one to study.
As can be observed, there is a variance between the trajectories for different values of , with some comprising of just one long excursion to the cusp, and others disproportionally bounded. This is why averaging on a relatively large set is required to obtain equidistribution, and otherwise one can construct counter examples as we have in §1.4 . Furthermore, there are visible symmetries between the trajectories for different values of , which were proved in [DS18] and are also used in this current paper (see Lemma 2.5).

1.3. Prime numerator Zaremba
In this part of the introduction we wish to present a conjecture we refer to as the prime numerator Zaremba conjecture. The discussion in this subsection has a different flavour to it as we do not present any theorems but rather discuss a conjecture and present supporting data. Nevertheless, we find the conjecture interesting and think it fits in this paper as we were inspired by Theorem 1.4 when we formulated it. We hope it will inspire other people to look into it as well.
The Zaremba conjecture asserts that for any denominator , one can find such that the c.f.e of satisfies for all . Lets call such a , a -Zaremba numerator for the denominator . This conjecture is in direct contrast to Theorem 1.1 which implies in particular, that the set of -Zaremba numerators in has size bounded by as and in particular its proportion in approaches zero. Intrigued by this, we used the computer to check if one should expect the existence of -Zaremba prime numerators and it seems that the data suggests so. We therefore suggest:
Conjecture 1.9.
For any , there exists a prime residue which is -Zaremba for . Moreover, for any which is not one of there exists a 4-Zaremba prime numerator.


The data, as can be seen in Figure 2, suggests that not only there are Zaremba numerators for any , but in general the number of them increases as the denominator grows. A related statement in this regard is to study the asymptotic number of Zaremba numerators for all , and show it grows fast enough as a function of in order for the Zaremba conjecture to hold. Indeed, without the prime restriction on the numerators, this was found by Hensley to be true [Hen89]. It would be interesting to know if it still holds with the prime numerator restriction.
One may speculate from the data in Figure 2 that the graphs converge to certain limits and , respectively, as . If the limit exists, it easily follows from Hensley’s work that the relation must be satisfied, where is the set of all irrationals in for which all digits in their c.f.e are bounded by . It is therefore interesting to wonder if is also related to the Hausdorff dimension of some set. If so, it seems reasonable to guess from our data that this dimension should be strictly smaller than . Similar question can be asked for bounds other than .
1.4. Counter examples
In this section we give examples for the necessity of the assumption in Theorem 1.5 and Corollary 1.7. In §1.4.1 we show that without this assumption escape of mass can occur, and in §1.4.2 that even if there is no escape of mass, equidistribution can still fail.
1.4.1. Escape of mass
Fix any . We consider
and show that the measures exhibit escape of mass. The size of these sets satisfies
which is clear for prime , and for general follows by using some inclusion-exclusion argument (as shown for example in Lemma [DS18, Lemma 2.12]).
For escape of mass, recall Mahler’s compactness theorem (c.f. §2) which says that an element is near the cusp if the corresponding lattice has a nontrivial short vector. More precisely, for any compact subset , there is some so that any satisfies .
Note that any and satisfy
i.e. the -trajectory of starts with a long excursion to cusp, outside of , for approximately a fraction of the total time. It follows that for any compact
and therefore for any weak* limit of , showing escape of mass and in particular preventing equidistribution.
1.4.2. Equidistribution
We show next that even if there is no escape of mass, equidistribution still may fail. For a denominator , consider the -Zaremba numerators as in §1.3:
Let
By a result of Hensely [Hen90], we have
for some constant , where is the set of all irrationals in for which all digits in their c.f.e are bounded by . As
it follows from Hensley’s result that there exists an increasing sequence so that
It is well known [Jar29] that
Therefore, by choosing large enough, the collection can be assumed to be of logarithmic size as close to as we would like.
However, due to the small digits in the c.f.e of for there is some compact set which depends only on , on which is supported for all (c.f. the proof of Theorem 1.4 in [DS18, pg. 173]). It follows that any weak* limit of is a compactly supported probability measure, in particular not equal to .
1.5. Structure of the paper
Acknowledgments
We are grateful to Elon Lindenstrauss for suggesting to us to look into improving Theorem 1.5. This work has received funding from the European Research Council (ERC) under the European Union’s Horizon 2020 Research and Innovation Program, Grant agreement no. 754475. R.M. acknowledges the support by ERC 2020 grant HomDyn (grant no. 833423).
2. Non-escape of mass for divergent orbits
In this section we prove Theorem 1.6. The key input required for the proof is the relation between entropy and escape of mass, which was studied in various previous works. The case which we consider in this work, was studied in [ELMV12] by Einsiedler, Lindenstrauss, Michel and Venkatesh. Since then, generalizations were obtained for other rank situations [EKP15, Mor22], as well as for the high rank case [EK12, KKLM17, Mor23]. We give here a brief overview of the ingredients we use and the idea of the proof.
For and , let , where is the ball in of radius around the identity. Define a forward Bowen -ball to be a set of the form for . For , we consider the set . It is straightforward to show [DS18, Lemma 2.10] that the points of are well separated, in the sense that two distinct points cannot belong to the same forward Bowen -ball, where , assuming is small enough. Therefore, instead of studying the size of a given subset of , we study the size of its covering by balls.
Note that a-priori, an amount of forward Bowen -balls is required to cover a given compact subset of , as each such ball has a small diameter of in the unstable direction. However, the key idea that was discovered in previous works is that if we only want to cover the points which spend a large proportion of their time high up in the cusp, we can cut down the number of balls. Quantitatively, if we only consider points which spend more than time in the cusp, the required number of balls is . We will use this idea in the proof of Theorem 1.6, where we bound the escape of mass rate, by showing that most points of do not spend a lot of time in the cusp during the interval .
In our proof, we will use the version of the aforementioned covering result as given in [KKLM17]. For that, we first define the discrete average in the diagonal direction with steps of size of a measure to be
and the continuous average on the segment
Next, consider the distance on induced by the absolute value on , and let be the ball in of radius centered at the identity. Given a compact subset , , , , and , define the set of points which spend at least a fraction of their discrete trajectory outside by
and similarly for the continuous version
Lastly, let be defined by , for the Euclidean norm, where we consider as the space of unimodular lattices in .
Then we have the following covering counting theorem based on [KKLM17, Theorem 1.5].
Theorem 2.1.
There exist such that the following holds. For any there exists a compact set of such that for any , , and , the set can be covered with
many balls in of radius .
Remark 2.2.
Proof of Theorem 2.1.
Observe that
By [KKLM17, Theorem 1.5], the right hand side set can be covered with balls in of radius . Since , the theorem follows by conjugating with . ∎
It is convenient to work with the the following standard notations for the cusp structure of . For any , denote
While Theorem 2.1 considers discrete averages, we would like to apply it to continuous averages. For that we have the following lemma.
Lemma 2.3.
-
(1)
Let , , , and with . Then
-
(2)
Assume in addition that , for as in Theorem 2.1. Then there are which depend on , so that for any and , the set can be covered by
many balls in of radius .
Proof.
In order to prove item (1), we will show that
as the desired result follows immediately from this claim and the definitions.
In order to prove the claim above, observe that for any ,
It follows that if , then . Let
so that . If , then and the claim is trivial. Otherwise, letting we get that
where for any . It follows that
Dividing by proves the claim, and thus completes the proof of item 1.
To prove item 2, let be as in Theorem 2.1. Let large enough so that . Then, for any , we have by definition and item 1
(2.1) |
By Theorem 2.1, the RHS of (2.1), and hence the LHS as well, can be covered by
many balls in of radius
Assuming is large enough compared to , the size of this covering is also bounded by
as desired. ∎
The previous lemma handled the covering of all points in the entire open ball (with certain restrictions on the time spent in the cusp). To prove the results of this paper we need to consider instead discrete averages over rational points, as follows. Given a set , we define
with implicit in the notation (note that this definition agrees with (1.3)). Using that previous covering result, we show that the measures don’t exhibit escape of mass. Extending later to time will be an easier task.
Proposition 2.4.
Let and such that
Then for all there is a compact set so that
Proof.
Let us start, for simplicity of notation, with a general finite subset and time , and at the end specialize to and .
We begin with a general upper bound for the ‘escaping mass’
by separating into ‘good’ and ‘bad’ points according to how much time the corresponding diagonal orbit spends in the cusp. For any , let
Then it follows that:
(2.2) | ||||
Hence, we want to give a lower bound to , which will be provided by the assumption of this proposition, and an upper bound to , which is given by Lemma 2.3 as follows. Given (which will be fixed later) we fix as in Lemma 2.3. Note that by definition,
Therefore, we can use Lemma 2.3 to cover by
many balls in of radius , for large enough.
Let us now apply for and , where . Note that each ball in our covering of is of radius , hence can contain at most points from . Letting , by the proposition’s assumption, for all large enough we have that . Therefore, if is large enough, we have for
Let us now prove the proposition. Since it is trivial for , we may assume that . We can then set , and note that if is small enough. Given such , fix large enough so that in addition. Then
All together, we obtain that
We now prove Theorem 1.6. To do so, we need to move to averaging over from the appearing in Proposition 2.4. This is done using the following symmetry that these measures enjoy.
Lemma 2.5.
[DS18, equation (2.2) page 154] Denote by the map that sends a lattice to its dual. For , let denote the unique integer satisfying modulo . Then for any and any ,
Proof of Theorem 1.6.
Let be as in the statement of the theorem. Then by Proposition 2.4, for any there is a compact set so that
Set . Then , hence we similarly have a compact set so that
By Lemma 2.5 we have that
Let . Putting it all together, we have
Note that is compact, since is compact and is continuous, hence it follows that any weak* limit of satisfies
As is arbitrary, the theorem follows. ∎
3. Continued fraction expansion of rationals
In this section we deduce Theorems 1.2-1.4 from Corollary 1.7. The first result we prove is Theorem 1.3, and the rest would follow with little work. We use the tight relationship between the -action on and continued fraction expansions, and a bit of the ergodicity of with respect to . We will use the analysis carried out in [DS18, §4] which in turn relies on the presentation of the aforementioned relation as in the book [EW11].
With a slight abuse of notation, for and we will let , where the latter is the probability measure supported on as before. We will be using the following result from [DS18, §4]:
Lemma 3.1 (Lemma 4.11 and Theorem 4.12 from [DS18]).
Let be a sequence of denominators and assume that for we are given satisfying . Then and for any finite string of integers , .
Remark 3.2.
We note that in [DS18, Lemma 4.11] there is no mention of the asymptotic densities . Still, the claim as we stated it follows easily; Let denote the subinterval of the unit interval consisting of those numbers whose starts with . Then, in the notations of [DS18, Lemma 4.11], it follows from the ergodicity of with respect to the Gauss map that (c.f. the introduction of [AS18]), and it follows straight from the definitions that . Since , we have . As we deduce that indeed
as claimed.
Recall that in Corollary 1.7 we showed that given the condition we obtain that . As can be seen in Lemma 3.1, we would like to consider measures instead. This will be done by moving from averages over to a claim about ‘almost every’ . This step is actually quite general in nature. To formulate it in a more general setting, we need the following definition.
Definition 3.3.
For a finite non-empty set of probability Radon measures on , we denote its average by
We also define to be the zero measure.
Definition 3.4.
If is a sequence of sets as in Definition 3.3, we say that they are uniformly almost invariant if any possible weak* limit point of is -invariant, for any choice of non-empty subsets .
Remark 3.5.
The application we have in mind for this last definition is , so that . The claim that is ‘uniformly almost invariant’ follows from the fact that each measure is a diagonal continuous average of a probability measure, over a time interval of length which approaches infinity.
We prove the following proposition, which is an application of the ergodicity of the Haar measure on .
Proposition 3.6.
Let be a sequence of uniform almost invariant sets such that . Then there exist such that and for any choice of a sequence we have that .
Proof.
We first claim that the conclusion of the proposition holds if for any and , there exists an integer such that for all ,
Indeed, since is a separable Banach space, there exists a countable dense subset of . Set and for each define
It follows from the claim that for any and ,
(3.1) |
For any , we set
By (3.1), we have , hence .
Note that for any choice of a sequence , we have for all .
Since is a dense subset of , it follows that . This proves the reduction.
Let us now prove the claim. Assume, for the sake of contradiction, that the claim does not hold. Then there exist , a function , and an unbounded subset of ’s such that for all
(3.2) |
It follows from (3.2) that there is an infinite subsequence such that along , at least one of the following statements hold:
-
(1)
-
(2)
Let us assume without loss of generality that option (1) holds. For , let
Since and are bounded, we may take an infinite subsequence so that and converge to some numbers , respectively. Note that . Next, recall that the set of all Radon measures with mass at most , is sequentially compact in the weak* topology. Therefore, we may assume that was chosen so that and as in , for some measures . As our sets are uniformly almost invariant, these limits are -invariant measures.
Let us now present as a convex combination as follows:
(3.3) |
Taking the limit over the sequence we conclude from (3.3), using the assumed convergence , that
If , then . Otherwise both , so the ergodicity of with respect to the -action also implies that . However, in either case, by definition of we have and hence
in contradiction. This finishes the proof of the claim and with it the proposition. ∎
We can now turn to prove the main results of this paper concerning the continued fraction expansions of rationals. We start from Theorem 1.3, from which Theorem 1.2 would then easily follow.
Proof of Theorem 1.3.
Let be as in the statement of the theorem. By Corollary 1.7, . Hence we can apply Proposition 3.6 and find a set of size , so that for any choices of we have that . Consequently, by Lemma 3.1, for every such sequence we have that
(3.4) |
and for any finite string of integers ,
(3.5) |
Assume by way of contradiction that for some we have
where is chosen uniformly at random. This means that along a subsequence we have that the set
occupies a positive proportion of (that is for some positive ), and in particular, for all large . We choose and arrive to a contradiction arising from (3.4) and the definition of .
The argument for the asymptotic frequency is identical, hence is omitted. ∎
Next, we deduce Theorem 1.2 from Theorem 1.3. We leave it to the reader to prove that the inverse deduction holds as well, so the two results are in fact equivalent.
Proof of Theorem 1.2.
We only prove item 2 concerning the length of the c.f.e. The argument for the asymptotic frequency is the same, hence is omitted. For any and , let
Note that an equivalent formulation of the theorem is the following inequality, which we now prove:
(3.6) |
We end up this paper with deducing Theorem 1.4, namely our application for rational numbers with prime numerators.
Proof of Theorem 1.4.
For any , let
We will show that
(3.8) |
This would allow us to apply Theorem 1.3 with this choice of , and the resulting statement is precisely Theorem 1.4.
Equation (3.8) is just a simple consequence of the prime number theorem. We include the argument for completeness. Let denote the number of primes . Then
Since the number of prime divisors of is bounded by we get that . In particular,
where the asymptotic relation follows from the fact that by the prime number theorem. So, in order to show (3.8), it is suffices to show that . Indeed, again by the prime number theorem, we know that Taking logarithms we obtain that
Dividing by we see that as needed. ∎
References
- [AKS81] Roy Adler, Michael Keane, and Meir Smorodinsky, A construction of a normal number for the continued fraction transformation, J. Number Theory 13 (1981), no. 1, 95–105. MR 602450
- [AS18] Menny Aka and Uri Shapira, On the evolution of continued fractions in a fixed quadratic field, J. Anal. Math. 134 (2018), no. 1, 335–397. MR 3771486
- [Bi05] V. A. Bykovski˘i, An estimate for the dispersion of lengths of finite continued fractions, Fundam. Prikl. Mat. 11 (2005), no. 6, 15–26. MR 2204419
- [BV05] Viviane Baladi and Brigitte Vallée, Euclidean algorithms are Gaussian, J. Number Theory 110 (2005), no. 2, 331–386. MR 2122613
- [Dix70] John D. Dixon, The number of steps in the Euclidean algorithm, J. Number Theory 2 (1970), 414–422. MR 266889
- [DS18] Ofir David and Uri Shapira, Equidistribution of divergent orbits and continued fraction expansion of rationals, J. Lond. Math. Soc. (2) 98 (2018), no. 1, 149–176. MR 3847236
- [EK12] Manfred Einsiedler and Shirali Kadyrov, Entropy and escape of mass for , Israel J. Math. 190 (2012), 253–288. MR 2956241
- [EKP15] M. Einsiedler, S. Kadyrov, and A. Pohl, Escape of mass and entropy for diagonal flows in real rank one situations, Israel J. Math. 210 (2015), no. 1, 245–295. MR 3430275
- [ELMV12] Manfred Einsiedler, Elon Lindenstrauss, Philippe Michel, and Akshay Venkatesh, The distribution of closed geodesics on the modular surface, and Duke’s theorem, Enseign. Math. (2) 58 (2012), no. 3-4, 249–313. MR 3058601
- [EW11] Manfred Einsiedler and Thomas Ward, Ergodic theory with a view towards number theory, Graduate Texts in Mathematics, vol. 259, Springer-Verlag London Ltd., London, 2011. MR 2723325
- [Hei69] H. Heilbronn, On the average length of a class of finite continued fractions, Number Theory and Analysis (Papers in Honor of Edmund Landau), Plenum, New York, 1969, pp. 87–96. MR 258760
- [Hen89] Doug Hensley, The distribution of badly approximable numbers and continuants with bounded digits, Théorie des nombres (1989), 371–385.
- [Hen90] by same author, The distribution of badly approximable rationals and continuants with bounded digits. II, J. Number Theory 34 (1990), no. 3, 293–334. MR 1049508
- [Hen94] by same author, The number of steps in the euclidean algorithm, J. Number Theory 49 (1994), no. 2, 142–182.
- [Jar29] Vojtĕch Jarník, Zur metrischen theorie der diophantischen approximationen, Prace Matematyczno-Fizyczne 36 (1928-1929), no. 1, 91–106 (pol).
- [KKLM17] S. Kadyrov, D. Kleinbock, E. Lindenstrauss, and G. A. Margulis, Singular systems of linear forms and non-escape of mass in the space of lattices, J. Anal. Math. 133 (2017), 253–277. MR 3736492
- [Mor22] Ron Mor, Excursions to the cusps for geometrically finite hyperbolic orbifolds and equidistribution of closed geodesics in regular covers, Ergodic Theory Dynam. Systems 42 (2022), no. 12, 3745–3791. MR 4504680
- [Mor23] by same author, Bounding entropy for one-parameter diagonal flows on using linear functionals, 2023, arXiv:2302.07122.
- [Por75] J. W. Porter, On a theorem of Heilbronn, Mathematika 22 (1975), no. 1, 20–28. MR 498452
- [Ust08] A. V. Ustinov, On the number of solutions of the congruence under the graph of a twice continuously differentiable function, Algebra i Analiz 20 (2008), no. 5, 186–216. MR 2492364
- [Van16] Joseph Vandehey, New normality constructions for continued fraction expansions, J. Number Theory 166 (2016), 424–451. MR 3486285