Startseite Regularity and inverse theorems for uniformity norms on compact abelian groups and nilmanifolds
Artikel
Lizenziert
Nicht lizenziert Erfordert eine Authentifizierung

Regularity and inverse theorems for uniformity norms on compact abelian groups and nilmanifolds

  • Pablo Candela EMAIL logo und Balázs Szegedy
Veröffentlicht/Copyright: 25. Mai 2022

Abstract

We prove a general form of the regularity theorem for uniformity norms, and deduce an inverse theorem for these norms which holds for a class of compact nilspaces including all compact abelian groups, and also nilmanifolds; in particular we thus obtain the first non-abelian versions of such theorems. We derive these results from a general structure theorem for cubic couplings, thereby unifying these results with the Host–Kra Ergodic Structure Theorem. A unification of this kind had been propounded as a conceptual prospect by Host and Kra. Our work also provides new results on nilspaces. In particular, we obtain a new stability result for nilspace morphisms. We also strengthen a result of Gutman, Manners and Varjú, by proving that a k-step compact nilspace of finite rank is a toral nilspace (in particular, a connected nilmanifold) if and only if its k-dimensional cube set is connected. We also prove that if a morphism from a cyclic group of prime order into a compact finite-rank nilspace is sufficiently balanced (i.e. equidistributed in a certain quantitative and multidimensional sense), then the nilspace is toral. As an application of this, we obtain a new proof of a refinement of the Green–Tao–Ziegler inverse theorem.

Funding statement: The first-named author received funding from Spain’s MICINN project MTM2017-83496-P. The second-named author received funding from the European Research Council under the European Union’s Seventh Framework Programme (FP7/2007-2013)/ERC grant agreement 617747. The research was supported partially by the NKFIH “Élvonal” KKP 133921 grant and partially by the Mathematical Foundations of Artificial Intelligence project of the National Excellence Programme (grant no. 2018-1.2.1-NKP-2018-00008).

A Results from nilspace theory

In this appendix our first and main aim is to prove Theorem 1.10. We also gather some results from nilspace theory which are adaptations of results from previous works.

We begin with the following useful description of cfr k-step nilspaces whose k-1 factor is toral, which was stated as Theorem 6.2.

Theorem A.1.

Let X be a k-step cfr nilspace such that the factor Xk-1 is toral. Let G denote the Lie group Θ(X), let G denote the degree-k filtration (Θi(X))i0, and for an arbitrary fixed xX let Γ=StabG(x). Then X is isomorphic as a compact nilspace to the coset space G/Γ with cube sets Cn(X)=(Cn(G)Γn)/Γn, n0.

To prove this, we adapt the proof of [3, Theorem 2.9.17].

Proof.

Fix xX and let Γ=StabG(x).

We first claim that Γ is discrete. Indeed, letting h:Θ(X)Θ(Xk-1) be the natural continuous homomorphism defined by h(α)(y)=πk-1(α(x)) (see [3, Lemma 2.9.3]), note that h(Γ) is a subgroup of the stabilizer of πk-1(x) in Θ(Xk-1), and since Xk-1 is toral, this stabilizer is discrete (see the proof of [3, Theorem 2.9.17]), so h(Γ) is discrete. Then, since h-1(h(Γ)) is a union of cosets of ker(h), it suffices to show that Γker(h) is discrete. This follows from [3, Lemma 2.9.9], since no non-trivial element of τ(Zk) stabilizes x.

By [3, Corollary 2.9.12] the Lie group Θ(X)0 acts transitively on the connected components of X, and since Xk-1 is toral, it follows that Θ(X)0,Zk acts transitively on X. Indeed, if x,yX are in different components, there is gΘ(Xk-1)0 such that gπk-1(x)=πk-1(y). Then there is gΘ(X)0 such that h(g)=g, and since g is path-connected to the identity in G, it follows that gx is in the same component as x. Moreover, by definition of h we have πk-1(gx)=gπk-1(x)=πk-1(y). There is therefore zZk such that zgx=y, which proves the claimed transitivity. Now since GΘ(X)0,Zk, we have that G also acts transitively on X, whence X is homeomorphic to the coset space G/Γ (see [25, Chapter II, Theorem 3.2]). In particular, since X is compact, we have that Γ is cocompact.

Recall from [4, Definition 3.2.38] that two cubes c1,c2Cn(X) are said to be translation equivalent if there is an element cCn(G) such that c2(v)=c(v)c1(v). We now show that Cn(X)=πΓn(Cn(G)), i.e. that every cube on X is translation equivalent to the constant x cube. First we claim that for every cube cCn(X) there is a cube cCn(X) that is translation equivalent to the constant x cube and such that πk-1c=πk-1c. Indeed, given cCn(X), we have

πk-1cCn(Xk-1),

and since X is toral the latter cube is translation equivalent to the cube with constant value x=πk-1(x), i.e. πk-1c=c~x for some cube c~ on the group Θ(Xk-1)0 with the filtration (Θi(Xk-1)0)i0. By the unique factorization result for these cubes [4, Lemma 2.2.5], we have

c~=g~0F0g~2n-1F2n-1,

where g~jΘcodim(Fj)(Xk-1)0. By [3, Theorem 2.9.10 (ii)], for each j[0,2n) there is gjΘcodim(Fj)(X)0 such that h(gj)=g~j. Let c* be the cube in Cn(Θ(X)0) defined by

c*=g0F0g2n-1F2n-1.

Let c=c*x. This is in Cn(X), and is translation equivalent to the constant x cube. Moreover, by construction πk-1c equals

πk-1n(c*x)=(jh(gj)Fj)x=(jg~jFj)x=c~x=πk-1c.

This proves our claim.

It follows from [4, Theorem 3.2.19] and the definition of degree-k bundles (in particular [4, (3.5)]) that c-cCn(𝒟k(Zk)). But then, using translations from τ(Zk)=Θk(X), we can correct c further to obtain c, thus showing that c is itself a translation cube with translations from Θ(X). (Such a correction procedure has been used in previous arguments, see for instance the proof of [4, Lemma 3.2.25].)

We have thus shown that Cn(X)πΓn(Cn(G)). The opposite inclusion is clear, by definition of the groups Θi(X). ∎

We can now prove Theorem 1.10, which we restate here.

Theorem A.2.

Let X be a k-step cfr nilspace. If Ck(X) is connected, then X is toral.

Proof.

We argue by induction on k. For k=1 the statement is clear. For k>1, first note that Ck(Xk-1) is connected (by continuity of πk-1), and so (since projection to a k-1 face of a k cube is a continuous map) we have also that Ck-1(Xk-1) is connected, so by induction we have that Xk-1 is toral. Now suppose for a contradiction that X is not toral. Then the last structure group Zk must be a disconnected compact abelian Lie group. By quotienting out the torus factor of Zk if necessary, we can assume that X now has k-th structure group Zk being a finite abelian group of cardinality greater than 1. We shall now deduce that Ck(X) must be disconnected, a contradiction.

By Theorem A.1 we have that X is isomorphic to the coset nilspace (G/Γ,G), where G=Θ(X) and Γ=StabG(x) for some fixed point xX. Hence Ck(X)=Ck(G)/Γk. Let σk be the Gray code map on Gk (see [4, Definition 2.2.22]), and recall that restricted to Ck(G) this map takes values in Gk (see [3, Proposition 2.2.25]) and that GkZk (see [4, Lemma 3.2.37]). We know that shifting any value c(v) of a cube cCk(G) by any element of Zk still gives a cube in Ck(G) (see [4, Remark 3.2.12]). It follows that σk maps Ck(G) onto Zk. On the other hand, the map σk only takes the value idG on Γk, since ΓGk={idG} (as the action of GkZk is free). Now let C denote the identity component of Cn(G). It is standard that C is normal in Cn(G). We also have σk(CΓk)={idG}. Indeed, since σk is continuous and Zk is discrete, for every element cγCΓk we have σk(γ)=0, and cγ is in the same component as γ, so we must also have σk(cγ)=0. But then the product set CΓk must be a proper subgroup of Cn(G) (otherwise its image under σk would be Gk). Thus we have shown that Cn(G)/CΓk is not the one point space. Hence there are at least two disjoint cosets of CΓk forming a cover of Cn(G). Since the latter group is a Lie group, C is open, and therefore these covering cosets of CΓk are open sets. But then the quotient map q:Cn(G)Cn(G)/Γk (which is open) sends these cosets to disjoint open sets covering Ck(G)/Γk, so Ck(X) is disconnected. ∎

We add the following lemma concerning the Haar measures on cube sets.

Lemma A.3.

Let X be a k-step cfr nilspace such that Xk-1 is toral. Then for every integer n0 the connected components of Cn(X) have equal positive Haar measure.

Proof.

Recall that Cn(X) is a compact abelian bundle with base Cn(Xk-1), bundle projection π:=πk-1n, and structure group Z~k:=Cn(𝒟k(Zk)), where Zk is the k-th structure group of X (see[3, Lemma 2.2.12]). The Haar measure μ on Cn(X) is invariant under the continuous action of Z~k, by construction (see [3, Proposition 2.2.5]). Assuming that there is more than one component of Cn(X), let c1,c2 be any points in distinct components C1, C2, respectively. Then, since Xk-1 is toral, by [3, Theorem 2.9.17] there is a cube cCn(Θ(Xk-1)0) such that cπ(c1)=π(c2). By [3, Theorem 2.9.10] there is a cube c~Cn(Θ(X)0) such that π(c~c1)=π(c2). There is therefore zZ~k such that c~c1+z=c2. We also have c~c1 still in C1, because the map c1c~c1 is a composition of multiplications by face-group elements of the form gF where F is a face in n and g is in the connected Lie group Θcodim(F)(X)0. Hence (C1+z)C2 is non-empty (containing c2), so C1+zC2 (since C1+z is connected and C2 is a maximal connected set), whence μ(C1)=μ(C1+z)μ(C2). Similarly μ(C2)μ(C1). ∎

Next, we prove the properties of the Ud-seminorms from Definition 1.4.

Lemma A.4.

For every k-step compact nilspace X and every d2, the function defined by ffUd is a seminorm on L(X).

The case of this lemma for compact abelian groups is given in several sources, all based essentially on the original argument of Gowers in [14, Lemma 3.9]. The case of nilmanifolds appears in [28, Chapter 12, Proposition 12]. These two cases already yield (via inverse limits) the result for the class of nilspaces concerned in our main results. Below we recall another proof from [7], which works at the more general level of cubic couplings. Let us mention also that Ud is non-degenerate (and is therefore a norm on L(X)) when the step k of X is less than d. For compact abelian groups this follows from the fact that we always have fUdfU2=f^4, and for nilmanifolds it is given in [28, Chapter 12, Theorem 17]. For general compact nilspaces, this non-degeneracy property can be proved using known results from nilspace theory; as this is not needed in this paper, we omit the details.

Proof of Lemma A.4.

The lemma follows from results in [7], namely from [7, Proposition 3.6], which shows that the Haar measures μn on Cn(X) form a cubic coupling, and [7, Corollary 3.17], which yields the seminorm properties for a general cubic coupling. ∎

We close this appendix with a proof of Proposition 6.3. Recall the following basic useful description of polynomial sequences (see for instance [6, Lemma 2.8]).

Lemma A.5 (Taylor expansion).

Let gpoly(Z,G), where G has degree at most s. Then there are unique Taylor coefficientsgiGi such that for all nZ we have

g(n)=g0g1ng2(n2)gs(ns).

Conversely, every such expression defines a map gpoly(Z,G). Moreover, if HG and g is H-valued, then giH for each i.

Proof of Proposition 6.3.

Since ϕβ is a morphism G/Γ, it suffices to prove the following statement: for every morphism ϕ:G/Γ, there is a morphism ψ:G (whence ψpoly(,G)) such that πΓψ=ϕ. We prove this by descending induction on j[k+1], showing that the statement holds for maps ϕ taking values in (GjΓ)/Γ. For j=k+1, since Gk+1={idG}, the map ϕ is constant and the statement is trivially verified letting ψ be a constant Γ-valued map. For j<k+1, suppose that the statement holds for j+1 and that ϕ takes values in (GjΓ)/Γ. It follows from the filtration property that Gj+1Γ is a normal subgroup of GjΓ and that the quotient GjΓ/(Gj+1Γ) is an abelian group. Denoting this abelian group by Aj, let qj:(GjΓ)/ΓAj be the quotient map for the action of Gj+1 on (GjΓ)/Γ. Note that qj is a nilspace morphism. More precisely, for every cube cΓn on (GjΓ)/Γ (where cGjnCn(G)), we have qj(cΓn)=(q~jc)Γn, where q~j is the quotient homomorphism GjGj/Gj+1; this implies that every (j+1)-face of qj(cΓn) has value 0 under the Gray-code map σj+1, so qj is a morphism into 𝒟j(Aj). It follows that qjϕ is a morphism 𝒟j(Aj), and is in particular a polynomial map of degree at most k, so by Lemma A.5 we have qjϕ(x)==0ka(x) for x, for some aAj, and binomial coefficients (x). Since qj is surjective, there exist elements b0,b1,,bk in Gj such that qj(bΓ)=a for each . Let α:G be the polynomial map

α(x)==0kb(x),

and note that qj(α(x)Γ)=qjϕ(x) for all x. It follows that the map α-1ϕ is a morphism (Gj+1Γ)/Γ, so by induction there is a map ψpoly(,G) such that

α-1(x)ϕ(x)=ψ(x)Γfor all x.

Then ψ(x):=α(x)ψ(x) is a map in poly(,G) with the required property. ∎

B Miscellaneous measure-theoretic results

Lemma B.1.

Let (Ω,A,λ) be a probability space, let B be a sub-σ-algebra of A, and suppose that SA satisfies 1S-E(1S|B)L2ϵ. Then S={xΩ:E(1S|B)(x)>ϵ12} satisfies λ(SΔS)<5ϵ12.

Proof.

We first observe that

λ(SS)ϵ12<Ω(1-1S)𝔼(1S|)dλ=Ω𝔼(1S|)-1S𝔼(1S|)dλ
=λ(S)-𝔼(1S|)L22.

Moreover, from the assumption and the triangle inequality we have

𝔼(1S|)L21SL2-ϵ,

whence 𝔼(1S|)L221SL22-2ϵ=λ(S)-2ϵ. Therefore λ(SS)<2ϵ12.

On the other hand, we have

λ(S)-2ϵ𝔼(1S|)L22=𝔼(1S|),𝔼(1S|)=1S,𝔼(1S|)
SS𝔼(1S|)dλ+SS𝔼(1S|)dλλ(SS)+ϵ12,

so λ(SS)λ(S)-3ϵ12, whence λ(SS)3ϵ12.

Combining the main two inequalities above, the result follows. ∎

We use this lemma to prove the following fact about mod 0 intersections of conditionally independent σ-algebras.

Lemma B.2.

Let (Ω,A,λ) be a probability space, let B0,B1 be sub-σ-algebras of A such that B0λB1, let SiBi, i=0,1, and suppose that λ(S0ΔS1)ϵ. Then there exists CB0B1 such that λ(CΔSi)10ϵ14 for i=0,1.

Proof.

The assumption 1S0-1S1L22ϵ implies

1S0-𝔼(1S0|1)L21S0-1S1L2+1S1-𝔼(1S0|1)L2
ϵ12+𝔼(1S1-1S0|1)L22ϵ12.

The assumption 0λ1 implies that 𝔼(1S0|1) is 01-measurable (in particular, we have 𝔼(1S0|1)=𝔼(1S0|01)). By Lemma B.1 with =01 and 𝒜=0, the set

C={xΩ:𝔼(1S0|1)>(2ϵ12)12}

is in 01 and satisfies λ(CΔS0)5(2ϵ12)1210ϵ14. Similarly, by Lemma B.1 with 𝒜=1 instead of 𝒜=0, this set C satisfies λ(CΔS1)10ϵ14. ∎

We can use this lemma in turn to prove the following fact about ultraproducts of conditionally independent σ-algebras.

Lemma B.3.

Let (X,A,λ) be the ultraproduct of probability spaces (Xi,Ai,λi). For each i let Bi,0,Bi,1 be sub-σ-algebras of Ai such that Bi,0λiBi,1. For j=0,1 let Bj be the Loeb σ-algebra corresponding to the sequence (Bi,j)iN, and let C be the Loeb σ-algebra corresponding to (Bi,0λiBi,1)iN. Then B0λB1=λC and B0λB1.

Proof.

Note that the inclusion 0λ1λ𝒞 is clear, for if A𝒞, then there are sets Aii,0λii,1 such that A=λiωAi, so iωAi is in j up to a null set, j=0,1, whence A0λ1. For the opposite inclusion, let Q be in 0λ1, so for j=0,1 there are sets Qi,ji,j for each i such that Q=λiωQi,j. Then

0=λ((iωQi,0)Δ(iωQi,1))=λ(iω(Qi,0ΔQi,1)),

so letting ϵi=λi(Qi,0ΔQi,1), we have limωϵi=0. By Lemma B.2, for each i there is a set Cii,0λii,1 such that

λ(CiΔQi,j)10ϵi14for j=0,1.

Let R=iωCi. By construction R𝒞, and by the last inequality we have R=λQ, so the required inclusion holds. Finally, the desired conclusion 0λ1 can be seen to follow from i,0λii,1, i, using the definition of conditional independence [7, Definition 2.9] and basic facts about Loeb probability spaces. More precisely, by [7, Theorem 2.4 and Remark 2.5] it suffices to show that every function f in L(1) satisfies 𝔼(f|0)=λ𝔼(f|0λ1). To show this, we use first that f is λ-almost-surely equal to a measurable function of the form f=limωfi (see [35, Corollary 5.1]), and then we prove the equality

𝔼(f|0)=λ𝔼(f|0λ1),

by deducing it from the fact that, by the assumption i,0λii,1, the analogous equality holds for the fi. This last deduction is enabled by the fact that 𝔼(|0)=limω𝔼(|i,0), a fact which is confirmed in a straightforward way by checking that for any function of the form g=limωgiL1(𝒜) (with each gi measurable) we have that limω𝔼(gi|i,0) satisfies the defining property of the conditional expectation 𝔼(g|0), i.e. that for every hL1(0) we have 𝐗hgdλ=𝐗hlimω𝔼(gi|i,0)dλ. This last equality is seen using an S-integrable lifting of h (see [35, Theorem 6.4]), commuting ultralimit and integrals as afforded by [35, Theorem 6.2, part 4], and basic properties of ultralimits. ∎

We also prove the following approximation result for measure-preserving group actions.

Lemma B.4.

Let G be an amenable group acting on a Borel probability space (Ω,A,λ) by measure-preserving transformations, and let SA be such that for some ϵ>0 we have λ(SΔ(gS))ϵ for every gG. Then there exists a set SA such that gS=λS for all gG and λ(SΔS)5ϵ14.

Proof.

We first suppose that G is countable. Let (Fj)j be a Følner sequence in G and for each j let hj=𝔼gFj1gS. By the mean ergodic theorem for amenable groups [43, Theorem 2.1], letting be the σ-algebra of G-invariant sets in 𝒜, and f be a version of 𝔼(1S|), we have f-hjL20 as j. Note that for every j we have

1S-fL21S-hjL2+hj-fL2hj-fL2+𝔼gFj1S-1gSL2
hj-fL2+ϵ12,

so letting j yields 1S-fL2ϵ12. By Lemma B.1, the set S={xΩ:f(x)>ϵ14} satisfies λ(SΔS)5ϵ14, and since f is G-invariant, we have gS=λS for every gG.

We now reduce the general case to the countable case. It suffices to prove that if G is a group acting on a separable metric space (X,d) by isometries, then there is a countable group G0G such that if xX is a fixed point for G0, then it is a fixed point for G (we then apply this with X the measure algebra of 𝒜). Let (xi)i be a dense sequence in X. For each i, the orbit Gxi is itself separable, so there is a countable set SiG such that Sixi is dense in this orbit. Let G0 be the subgroup of G generated by iSi. Observe that for every i, gG and ϵ>0, there is gSiG0 such that d(gxi,gxi)<ϵ. Suppose for a contradiction that there is xX that is G0-invariant but not G-invariant, so d(gx,x)=ϵ>0. Then by the density of (xi)i there is i such that d(x,xi)<ϵ100, so

d(gxi,xi)d(gxi,x)-d(x,xi)d(gx,x)-d(gxi,gx)-d(x,xi),

which by the isometry property equals d(gx,x)-2d(x,xi)98ϵ100. Hence

d(gxi,xi)98ϵ100.

By the earlier observation, there is gG0 such that d(gxi,gxi)<ϵ100, so

d(gxi,xi)d(gxi,xi)-d(gxi,gxi)97ϵ100.

Combining this last inequality with d(x,xi)<ϵ100 and the triangle inequality and isometry property, we deduce that d(gx,x)d(gxi,xi)-2d(x,xi)95ϵ100, which contradicts that x is G0-invariant. ∎

Lemma B.5.

Let Y be a compact Polish space, let d be a metric compatible with the weak topology on P(Y), and let (Xi,λi)iN be a sequence of Borel probability spaces. For each iN let fi:XiY be a Borel function, and let ω be a non-principal ultrafilter on N. Then, letting f=limωfi, we have limωd(λifi-1,λf-1)=0.

Proof.

As shown in [29, Theorem (17.19)], one can always metrize this space of probability measures with a metric of the form d(μ,ν)=r12r|hrdμ-hrdν|, for a sequence of continuous functions hr:Y with hr1, r. Since d and d metrize the same topology, it suffices to prove that limωd(λifi-1,λf-1)=0.

Suppose for a contradiction that for some b(0,1) and some set Sω, for every iS we have d(λifi-1,λf-1)>b. Then, for each iS, a short argument by contradiction shows that there exists r=r(i)[1,2log2(2b)] such that

|Xihrfidλi-Xhrfdλ|b2.

Using the ultrafilter properties, we then deduce that for some fixed integer r there is a set SS with Sω such that for all iS we have

|Xihrfidλi-Xhrfdλ|b2.

Now we have two exhaustive possibilities. The first one is that some S′′S with S′′ω satisfies XihrfidλiXhrfdλ+b2 for all iS′′; but then, commuting ultralimit and integrals (as in the proof of Lemma B.3), we obtain

Xhrfdλ=limωXihrfidλiXhrfdλ+b2>Xhrfdλ,

a contradiction. The other option is that some S′′S with S′′ω satisfies

XhrfdλXihrfidλi+b2for all iS′′;

then we deduce similarly that

Xhrfdλ=limωXihrfidλiXhrfdλ-b2<Xhrfdλ,

obtaining again a contradiction. ∎

We finish with a lemma concerning the interaction of the Loeb-measure construction with products, when the underlying measures are couplings on Borel probability spaces.

Lemma B.6.

Let (Xi)iN, (Yi)iN be sequences of Polish spaces, and for each iN let μi be a Borel probability measure on B(Xi) and let νi be a Borel probability measure on B(Xi)B(Yi). Let (X,LX,μ), (X×Y,LX×Y,ν) be the corresponding Loeb probability spaces. Suppose that the projection πi:Xi×YiXi, (x,y)x is measure preserving for every iN. Then the projection π:X×YX, (x,y)x is measurable with respect to LX, LX×Y, and is measure-preserving with respect to μ,ν.

Proof.

The preimage under π of any internal measurable set in 𝐗 is an internal measurable set in 𝐗×𝐘, and it is also clear that if A is an internal measurable subset of 𝐗, then νπ-1(A)=μ(A). (These claims follow from the fact the projections πi are measure-preserving maps and that taking ultraproducts commutes with taking preimages under the projections.) Now 𝐗 consists precisely of sets S such that for every ϵ>0 there exist internal measurable sets Ai,Ao𝐗 with AiSAo and μ(AoAi)<ϵ (see [35, Section 2.1]). This combined with the properties already established for π for internal sets implies that π-1(𝐗)𝐗×𝐘 and μπ-1=ν, as required. ∎

Acknowledgements

We thank Terence Tao for useful feedback. We also thank the anonymous referee for valuable feedback helping to improve this paper.

References

[1] V. Bergelson, T. Tao and T. Ziegler, An inverse theorem for the uniformity seminorms associated with the action of 𝔽p, Geom. Funct. Anal. 19 (2010), no. 6, 1539–1596. 10.1007/s00039-010-0051-1Suche in Google Scholar

[2] O. A. Camarena and B. Szegedy, Nilspaces, nilmanifolds and their morphisms, preprint (2010), http://arxiv.org/abs/1009.3825. Suche in Google Scholar

[3] P. Candela, Notes on compact nilspaces, Discrete Anal. 2017 (2017), Paper No. 16. 10.19086/da.2106Suche in Google Scholar

[4] P. Candela, Notes on nilspaces: Algebraic aspects, Discrete Anal. 2017 (2017), Paper No. 15. Suche in Google Scholar

[5] P. Candela, D. González-Sánchez and B. Szegedy, On nilspace systems and their morphisms, Ergodic Theory Dynam. Systems 40 (2020), no. 11, 3015–3029. 10.1017/etds.2019.24Suche in Google Scholar

[6] P. Candela and O. Sisask, Convergence results for systems of linear forms on cyclic groups and periodic nilsequences, SIAM J. Discrete Math. 28 (2014), no. 2, 786–810. 10.1137/130935677Suche in Google Scholar

[7] P. Candela and B. Szegedy, Nilspace factors for general uniformity seminorms, cubic exchangeability and limits, preprint (2018), http://arxiv.org/abs/1803.08758; to appear in Mem. Amer. Math. Soc. Suche in Google Scholar

[8] N. J. Cutland, Nonstandard measure theory and its applications, Bull. Lond. Math. Soc. 15 (1983), no. 6, 529–589. 10.1112/blms/15.6.529Suche in Google Scholar

[9] G. Elek and B. Szegedy, A measure-theoretic approach to the theory of dense hypergraphs, Adv. Math. 231 (2012), no. 3–4, 1731–1772. 10.1016/j.aim.2012.06.022Suche in Google Scholar

[10] D. H. Fremlin, Measure theory. Vol. 2. Broad foundations, Torres Fremlin, Colchester 2003. Suche in Google Scholar

[11] D. H. Fremlin, Measure theory. Vol. 3. Measure algebras, Torres Fremlin, Colchester 2004. Suche in Google Scholar

[12] G. Georganopoulos, Sur l’approximation des fonctions continues par des fonctions lipschitziennes, C. R. Acad. Sci. Paris Sér. A-B 264 (1967), A319–A321. Suche in Google Scholar

[13] E. Glasner, Y. Gutman and X. Ye, Higher order regionally proximal equivalence relations for general minimal group actions, Adv. Math. 333 (2018), 1004–1041. 10.1016/j.aim.2018.05.023Suche in Google Scholar

[14] W. T. Gowers, A new proof of Szemerédi’s theorem, Geom. Funct. Anal. 11 (2001), no. 3, 465–588. 10.1007/s00039-001-0332-9Suche in Google Scholar

[15] W. T. Gowers, Generalizations of Fourier analysis, and how to apply them, Bull. Amer. Math. Soc. (N. S.) 54 (2017), no. 1, 1–44. 10.1090/bull/1550Suche in Google Scholar

[16] W. T. Gowers and L. Milićević, An inverse theorem for Freiman multi-homomorphisms, preprint (2020), http://arxiv.org/abs/2002.11667. Suche in Google Scholar

[17] B. Green and T. Tao, The primes contain arbitrarily long arithmetic progressions, Ann. of Math. (2) 167 (2008), no. 2, 481–547. 10.4007/annals.2008.167.481Suche in Google Scholar

[18] B. Green and T. Tao, An arithmetic regularity lemma, an associated counting lemma, and applications, An irregular mind, Bolyai Soc. Math. Stud. 21, János Bolyai Mathematical Society, Budapest (2010), 261–334. 10.1007/978-3-642-14444-8_7Suche in Google Scholar

[19] B. Green and T. Tao, The quantitative behaviour of polynomial orbits on nilmanifolds, Ann. of Math. (2) 175 (2012), no. 2, 465–540. 10.4007/annals.2012.175.2.2Suche in Google Scholar

[20] B. Green, T. Tao and T. Ziegler, An inverse theorem for the Gowers Us+1[N]-norm, Ann. of Math. (2) 176 (2012), no. 2, 1231–1372. 10.4007/annals.2012.176.2.11Suche in Google Scholar

[21] Y. Gutman and Z. Lian, Strictly ergodic distal models and a new approach to the Host–Kra factors, preprint (2019), http://arxiv.org/abs/1909.11349. 10.1016/j.jfa.2022.109779Suche in Google Scholar

[22] Y. Gutman, F. Manners and P. P. Varjú, The structure theory of nilspaces II: Representation as nilmanifolds, Trans. Amer. Math. Soc. 371 (2019), no. 7, 4951–4992. 10.1090/tran/7503Suche in Google Scholar

[23] Y. Gutman, F. Manners and P. P. Varjú, The structure theory of nilspaces I, J. Anal. Math. 140 (2020), no. 1, 299–369. 10.1007/s11854-020-0093-8Suche in Google Scholar

[24] Y. Gutman, F. Manners and P. P. Varjú, The structure theory of nilspaces III: Inverse limit representations and topological dynamics, Adv. Math. 365 (2020), Article ID 107059. 10.1016/j.aim.2020.107059Suche in Google Scholar

[25] S. Helgason, Differential geometry, Lie groups, and symmetric spaces, Pure Appl. Math. 80, Academic Press, New York 1978. Suche in Google Scholar

[26] B. Host and B. Kra, Nonconventional ergodic averages and nilmanifolds, Ann. of Math. (2) 161 (2005), no. 1, 397–488. 10.4007/annals.2005.161.397Suche in Google Scholar

[27] B. Host and B. Kra, Parallelepipeds, nilpotent groups and Gowers norms, Bull. Soc. Math. France 136 (2008), no. 3, 405–437. 10.24033/bsmf.2561Suche in Google Scholar

[28] B. Host and B. Kra, Nilpotent structures in ergodic theory, Math. Surveys Monogr. 236, American Mathematical Society, Providence 2018. 10.1090/surv/236Suche in Google Scholar

[29] A. S. Kechris, Classical descriptive set theory, Grad. Texts in Math. 156, Springer, New York 1995. 10.1007/978-1-4612-4190-4Suche in Google Scholar

[30] A. Leibman, Polynomial mappings of groups, Israel J. Math. 129 (2002), 29–60. 10.1007/BF02773152Suche in Google Scholar

[31] G. W. Mackey, Point realizations of transformation groups, Illinois J. Math. 6 (1962), 327–335. 10.1215/ijm/1255632330Suche in Google Scholar

[32] F. Manners, Periodic nilsequences and inverse theorems on cyclic groups, preprint (2014), http://arxiv.org/abs/1404.7742. Suche in Google Scholar

[33] F. Manners, Quantitative bounds in the inverse theorem for the Gowers Us+1-norms over cyclic groups, preprint (2018), https://arxiv.org/pdf/1811.00718.pdf. Suche in Google Scholar

[34] R. S. Palais, The classification of G-spaces, Mem. Amer. Math. Soc. 36 (1960), 1–72. 10.1090/memo/0036Suche in Google Scholar

[35] D. A. Ross, Loeb measure and probability, Nonstandard analysis (Edinburgh 1996), NATO Adv. Sci. Inst. Ser. C: Math. Phys. Sci. 493, Kluwer Academic, Dordrecht (1997), 91–120. 10.1007/978-94-011-5544-1_4Suche in Google Scholar

[36] B. Szegedy, Higher order Fourier analysis as an algebraic theory I, preprint (2009), http://arxiv.org/abs/0903.0897. Suche in Google Scholar

[37] B. Szegedy, Gowers norms, regularization and limits of functions on abelian groups, preprint (2010), http://arxiv.org/abs/1010.6211. Suche in Google Scholar

[38] B. Szegedy, On higher order Fourier analysis, preprint (2012), http://arxiv.org/abs/1203.2260. Suche in Google Scholar

[39] T. Tao, Hilbert’s fifth problem and related topics, Grad. Stud. Math. 153, American Mathematical Society, Providence 2014. 10.1090/gsm/153Suche in Google Scholar

[40] T. Tao and T. Ziegler, The inverse conjecture for the Gowers norm over finite fields via the correspondence principle, Anal. PDE 3 (2010), no. 1, 1–20. 10.2140/apde.2010.3.1Suche in Google Scholar

[41] T. Tao and T. Ziegler, The inverse conjecture for the Gowers norm over finite fields in low characteristic, Ann. Comb. 16 (2012), no. 1, 121–188. 10.1007/s00026-011-0124-3Suche in Google Scholar

[42] E. Warner, Ultraproducts and the foundations of higher order Fourier analysis, PhD thesis, Princeton University, Princeton 2012, https://www.math.columbia.edu/~warner/notes/UndergradThesis.pdf. Suche in Google Scholar

[43] B. Weiss, Actions of amenable groups, Topics in dynamics and ergodic theory, London Math. Soc. Lecture Note Ser. 310, Cambridge University, Cambridge (2003), 226–262. 10.1017/CBO9780511546716.012Suche in Google Scholar

Received: 2020-12-21
Revised: 2022-02-24
Published Online: 2022-05-25
Published in Print: 2022-08-01

© 2022 Walter de Gruyter GmbH, Berlin/Boston

Heruntergeladen am 26.10.2025 von https://www.degruyterbrill.com/document/doi/10.1515/crelle-2022-0016/html
Button zum nach oben scrollen