MINIMAL AND NEARLY MINIMAL EXPANSIONS IN CONNECTED UNIMODULAR GROUPS

YIFAN JING AND CHIEU-MINH TRAN

Abstract. Let G be a connected unimodular equipped with a (left and hence right) µG, and suppose A, B ⊆ G are nonempty and compact. An inequality by Kemperman gives us µG(AB) ≥ min{µG(A) + µG(B), µG(G)}. Our first result determines the conditions for the equality to hold, providing a complete answer to a question asked by Kemperman in 1964. Our second result characterizes compact and connected G, A, and B that nearly realize equality, with quantitative bounds having the sharp exponent. This can be seen up-to-constant as a (3k − 4)-theorem for this setting, and confirms the connected case of conjectures by Griesmer and by Tao. As an application, we get a measure expansion gap result for connected compact simple Lie groups. The tools developed in our proof include an analysis of the shape of minimally and nearly minimally expanding pairs of sets, a bridge from this to the properties of a certain pseudometric, and a construction of appropriate continuous group homomorphisms to either R or T = R/Z from the pseudometric.

Contents 1. Introduction 2 1.1. Background 2 1.2. Statement of main results 6 1.3. Notation and convention 8 2. Outline of the argument 8 2.1. Overview of the strategy 8 2.2. The first step: Nearly minimal expansions and quotients 11 2.3. The second step: Pseudometrics and homomorphisms 12 2.4. The third step: Nearly minimal expansions and 14 2.5. Structure of the paper 15 arXiv:2006.01824v3 [math.CO] 17 Jun 2021 3. Preliminaries 17 3.1. Locally compact groups and Haar measures 17

2010 Mathematics Subject Classification. Primary 22D05; Secondary 11B30, 05D10, 03C20, 43A05. YJ was supported by Arnold O. Beckman Research Award (Campus Research Board RB21011), by the Department Fellowship, and by the Trjitzinsky Fellowship from UIUC. 1 MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 2

3.2. More on Kemperman’s inequality and the inverse problems 19 4. Reduction to classifying nearly minimally expanding pairs 21 5. Reduction to sets with small measure 23 6. Geometry of minimal and nearly minimal expansions I 25 6.1. Preservation of minimal expansion under quotient 26 6.2. Coarse versions of the main theorems 30 6.3. Structure control on the nearly minimally expanding pairs 33 7. Pseudometrics and group homomorphisms onto tori 36 7.1. Linear pseudometrics 37 7.2. Almost linear pseudometrics: relative sign and total weight functions 41 7.3. Almost linear pseudometrics: group homomorphisms onto tori 46 8. Geometry of minimal and nearly minimal expansions II 55 8.1. Toric transversal intersection in measure 55 8.2. Linear pseudometric from minimal expansions 63 8.3. Almost linear pseudometric from near minimal expansions 68 9. Proof of the main theorems 75 9.1. Minimal expansions in noncompact groups 76 9.2. Nearly minimal expansions in compact groups 77 Acknowledgements 81 References 81

1. Introduction 1.1. Background. The Cauchy–Davenport theorem asserts that if X and Y are nonempty of the group Z/pZ of prime order p, then |X + Y | ≥ min{|X| + |Y | − 1, p}, where we X + Y := {x + y : x ∈ X, y ∈ Y }. The condition for X and Y to have minimal expansion (i.e. equality happens in the above inequality) is essentially given by Vosper’s theorem [Vos56], which states that if 1 < |X|, |Y |, and |X + Y | = |X| + |Y | − 1 < p − 1 then X and Y must be arithmetic progressions with the same common difference. When X and Y have nearly minimal expansion (i.e. the equality nearly happens), one might expect that X and Y are instead contained in arithmetic progressions with slightly larger cardinalities. This was confirmed by Freiman [Fre73] for small X = Y even though the optimal statement, known as the (3k − 4)-conjecture for Z/pZ, remains wide open. Similar results were also obtained for other abelian groups; see e.g. [Kne56, Kem60, DF03, GR06, Tao18, Gri19, Lev20]. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 3

This paper considers Kemperman’s analog of the Cauchy–Davenport theorem to connected unimodular locally compact nonabelian groups. We will determine the conditions for equality to happen and the condition for equality nearly hap- pen when the group is compact. Our work is inspired by recent progresses in the study of small expansions in the nonabelian settings; see [BGT12], in partic- ular, for the classification of approximate groups by Breuillard, Green, and Tao; see also [Hel08, BGS10, BV12, Hru12, PS16, BH18, Hru20]. Throughout, let G be a connected locally , and µG a left Haar measure on G. We further assume that G is unimodular (i.e., the measure µG is invariant under right translation), so µG behaves like an appropriate notion of size. This assumption holds in many situations of interest (e.g, when G is compact, discrete, a nilpotent , a semisimple Lie group, etc). As usual, for A, B ⊆ G, we set AB := {ab : a ∈ A, b ∈ B} and let An be the n-fold product of A for n ∈ N>0. In [Kem64], Kemperman proved that if A, B ⊆ G are nonempty and compact, then

µG(AB) ≥ min{µG(A) + µG(B), µG(G)}. This generalizes earlier results for one-dimensional tori, n-dimensional tori, abelian groups, and second countable compact groups by Raikov [Rai39], Macbeath [Mac53], Kneser [Kne56], and Shields [Shi55]. The problem of determining when equality happens for this inequality was pro- posed in the same paper [Kem64]. After handling a number of easy cases, the problem can be reduced to classifying all connected and unimodular group G and pairs (A, B) of compact subsets of G such that

0 < µG(A), µG(B), and µG(AB) = µG(A) + µG(B) < µG(G). We call such (A, B) a minimally expanding pair on G. It is easy to see that, if I and J are closed intervals in T = R/Z, such that I and J have positive measures and the total of their measures is strictly smaller than µT(T), then I + J is an interval with length the total lengths of I and J. Hence, such (I,J) is a minimally expanding pair on T. More generally, when G is a compact group, χ : G → T is a continuous surjective group homomorphism, I and J are as before, A = χ−1(I) and B = χ−1(J), we can check by using the Fubini theorem that (A, B) is a minimally expanding pair. Note that an arithmetic progression on Z/pZ is the inverse image under a group homomorphism φ : Z/pZ → T of an interval on T, so this example is the counterpart of Vosper’s classification. Another obvious example is when G is a noncompact group, χ : G → R is a continuous surjective group homomorphism with compact kernel, I and J are nonempty compact intervals in R, A = χ−1(I), and B = χ−1(J). MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 4

One might optimistically conjecture, in analogy with Vosper’s theorem, that there are no other G, A, and B such that (A, B) is a minimally expanding pair on G. In view of the earlier discussions, for compact A, B ⊆ G, we say that (A, B) is a δ-nearly minimally expanding pair on G if

0 < µG(A), µG(B), and µG(AB) < µG(A)+µG(B)+δ min{µG(A), µG(B)} < µG(G). We interpret the problem of determining when equality nearly happens in the Kem- perman inequality as classifying all connected and unimodular groups G and δ-nearly minimally expanding pairs (A, B) on G. In analogy with the discussion for the Cauchy–Davenport theorem, we hope for an answer along the following line: If G is compact, and (A, B) is a δ-nearly minimally expanding pair on G with small δ, then there is a continuous and surjective group homomorphism χ : G → T, compact interval I,J ⊆ T , and small ε, such that

−1 −1 −1 −1 (1) A ⊆ χ (I),B ⊆ χ (J), µG(χ (I) \ A) < εµG(A), µG(χ (I) \ B) < εµG(B).

The optimistic conjecture for noncompact groups is similar, but with T replaced by R and an extra condition that χ has compact kernel. Under the extra assumption that G is abelian, the optimistic conjectures for both classification problems were more or less confirmed before our work. In the same paper [Kne56] mentioned earlier, Kneser solved the classification problem for equality with the answer we hope for. For the near equality problem, when G = Td, the desired classification was ob- tained by Bilu [Bil98], and later improved by Candela and De Roton [CDR19] for a special case when d = 1. When G is a general , a classifica- tion result was obtained by Tao [Tao18] for compact G, and by Griesmer [Gri19] when G is noncompact. Griesmer also proved more general results for disconnected groups [Gri14, Gri19]. The results by Griesmer [Gri14, Gri19] and by Tao [Tao18] used nonstandard analysis methods, and did not provide how ε quantitative depends on δ in (1). A sharp exponent classification result (i.e., ε = O(δ)) for compact abelian groups was obtained very recently by Christ and Iliopoulou [CI21]. Results with sharp exponent bounds are likely the best that one can achieve without solving the (3k − 4)-conjecture for Z/pZ. For nonabelian G, not much was known earlier than this paper. In closest proxim- ity to the current work, Bj¨orklundconsidered in [Bj¨o17]a variation of Kemperman’s inequality and the equality classification problem without assuming that G is con- nected while assuming additionally that G is compact, second countable, and has abelian identity component, and the sets A and B are “spread out” (i.e., far away from being subgroups). The only common case to this and our current setting hap- pens when G is abelian and connected. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 5

Toward showing that appropriate versions of the optimistic conjectures also hold for the nonabelian classification problem, there is an important new challenge: While the desired conclusions for the abelian setting are mainly about the structure of (A, B), the structure of G is also highly involved for the nonabelian setting. If G = SO3(R), for example, one would not be able to find a minimally expanding pair according to the optimistic answers because there is no continuous surjective group homomorphism from SO3(R) to T. On the other hand, one can always find a continuous and surjective group homomorphism from a compact connected nontrivial abelian group to T and use this to construct minimally expanding pairs. The above challenge connects our problem to the subject of small expansions in nonabelian groups, a fascinating topic that brings together ideas from different areas of mathematics. The phenomenon that expansion rate encodes structural informa- tion about the group can already be seen through the following famous theorem by Gromov [Gro81] in geometric group theory: If G is a group generated by a finite set X = X−1, and |Xn| grows polynomially as a of n, then G must be virtually nilpotent. A more recent result by Breuillard indicates that some of the analysis go through for locally compact groups [Bre14]. Even more suggestive is the classifications of approximate groups in [BGT12] mentioned earlier (see the defini- tion in Section 6.2). In our proof, we will use the continuous version of the result in [BGT12]; this was proven in the thesis of Carolino [Car15] and can also be deduced from the result in [BGT12] using a result of Massicot–Wagner [MW15]. The ideas in the proof of these results can be traced back to the solution of Hilbert’s Fifth prob- lem by Montgomery–Zippin [MZ52], Gleason [Gle52], and Yamabe [Yam53], which we will also use later on. Finally, let us mention that these stories are also closely tied to the study of definable groups in model theory. This is the natural habitat of the aforementioned result by Massicot–Wagner [MW15], and also of Hrushovski’s Lie model theorem [Hru12], a main ingredient for the proof of the main theorem in [BGT12]. Before getting to the results, we briefly survey a number of works for nonabelian groups which are thematically relevant but use different techniques. When A, B ⊆ G are finite and nonempty, DeVos [DeV13] classified all situations where |AB| < |A| + |B|. In [BF19], Bj¨orklund and Fish studied an expansion problem with respect to upper Banach density in amenable nonabelian groups and obtained conclusions with similar flavor. Under the assumption that A is a finite of a group G such that the relation xy ∈ A has finite VC-dimension (or NIP), Terry, Conant, and Pillay [CPT21, CP20] shows that A must have a structure which is surprisingly similar to the optimistic conjecture mentioned earlier. It would also be interesting to study a different minimal and nearly minimal mea- sure expansion problem where we fix a connected unimodular group G instead of letting G range over all connected unimodular group G as we are doing here. When MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 6

G is Rn, Kemperman inequality is a consequence of the Brunn–Minkowski inequality 1/n 1/n 1/n µG(AB) ≥ µG(A) + µG(B) . This inequality also holds for nilpotent G [McC69, Gro03, Tao11]. The equality holds in the Brunn–Minkowski inequality for Rn if and only if A and B are homothetic convex subsets of Rn. This was a result by Brunn and Minkowski when A and B are further assumed to be convex, and a result by Lyusternik [Lyu35], Henstock and Macbeath [HM53] in the general case. A qualitative answer for the near equality Brunn–Minkowski problem for Rn is obtained by Christ [Chr12], and a quantitative version is obtained by Figalli and Jerison [FJ17]. We do not pursue this direction further here.

1.2. Statement of main results. Our first main result determines the conditions for equality to happen in the Kemperman inequality answering a question by Kem- perman in [Kem64]. Scenario (v) and (vi) in the theorem is a classification of the groups G and minimally expanding pairs (A, B) on G. Theorem 1.1. Let G be a connected unimodular group, and let A, B be nonempty compact subsets of G. If

µG(AB) = min{µG(A) + µG(B), µG(G)}. then we have the following:

(i) µG(A) + µG(B) = 0 implies µG(AB) = 0; (ii) µG(A) + µG(B) ≥ µG(G) implies AB = G; (iii) µG(A) = 0 and 0 < µG(B) < µG(G) imply there is compact H ≤ G and compact B1 ⊆ B such that with B2 = B \ B1, we have HB1 = B1, µG(AB2) = 0, and A ⊆ gH for some g ∈ G; (iv) µG(B) = 0 and 0 < µG(A) < µG(G) imply there is compact H ≤ G and compact A1 ⊆ A such that, with A2 = A \ A1, we have A1H = A1, µG(A2B) = 0, and B ⊆ Hg for some g ∈ G; (v)0 < min{µG(A), µG(B), µG(G) − µG(A) − µG(B)}, and G is compact together imply that there is a surjective continuous group homomorphism χ : G → T and compact intervals I and J in T with I + J 6= T such that A = χ−1(I) and B = χ−1(J); (vi)0 < min{µG(A), µG(B)}, and G is not compact together imply that there is a surjective continuous group homomorphism χ : G → R with compact kernel and compact intervals I and J in R such that A = χ−1(I) and B = χ−1(J).

Moreover, µG(AB) = min{µG(A) + µG(B), µG(G)} holds if and only if we are in exactly one of the implied scenarios in (i-vi). MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 7

Next we obtain a classification of nearly minimally expanding pairs. This answers questions by Griesmer [Gri19] and confirms a conjecture by Tao [Tao18, Conjecture 5.1], under the extra assumption of connectedness. Theorem 1.2. Let G be a connected compact group, and let A, B be compact subsets of G with 0 < λ = min{µG(A), µG(B), 1 − µG(A) − µG(B)}. There is a constant K = K(λ), not depending on G, such that for any 0 ≤ ε < 1, whenever we have δ ≤ Kε and

µG(AB) < µG(A) + µG(B) + δ min{µG(A), µG(B)}, there is a surjective continuous group homomorphism χ : G → T together with two compact intervals I,J ∈ T with

µT(I) < (1 + ε)µG(A), µT(J) < (1 + ε)µG(B), and A ⊆ χ−1(I), B ⊆ χ−1(J). It is worth noting that the linear dependence between ε and δ is the best possible up to a constant factor. The conclusions in Theorems 1.1 and 1.2, with suitable modifications, hold for arbitrary A, B ⊆ G with inner measures; see Section 9. The proof Theorem 1.2 yields a number of auxiliary results. One of them is a short proof of the main result in [Tao18] and its sharp-exponent improvement in [CI21]; see Theorem 6.14. We also showed quantitatively that measure approximate groups of a compact Lie group cannot be a “Kakeya set” with respect to the “torus directions”; see Theorem 8.12. Most notably, we obtain the following measure expansion gap result for sets in connected compact simple Lie groups. Theorem 1.3 (Expansion gaps in compact simple Lie groups). There is a constant η ≥ 10−12 such that the following holds. Let d > 0 be an integer. There is a constant C > 0 depending only on d such that if G is a connected compact simple Lie group of dimension at most d, and A is a compact set of G with 0 < µG(A) < C. Then 2 µG(A ) > (2 + η)µG(A). We did not try to optimize the constant η of Theorem 1.3 in this paper. One may compare Theorem 1.3 with expansion gaps for finite sets. The latter problem is well studied, which is initialed by Helfgott [Hel08] where he proved an expansion gap in SL2(Z/pZ). Results on the expansions for finite sets are one of the main ingredients in proving many of spectral gap results. For example, the result by Helfgott is largely used in the proof by Bourgain and Gamburd [BG08, BG12]. De Saxc´eproved in [dS15] an expansion gap results in simple Lie groups, which is used in the later proof of spectral gap results [BdS16, BIG17]. For more background in this direction, we refer the reader to [BL18, Tao15]. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 8

1.3. Notation and convention. Throughout, let k and l range over the set Z of integers, and m and n range over the set N = {0, 1,...} of natural numbers. A constant in this paper is always a positive . For real valued quantities r and s, we will use the standard notation r = O(s) and s = Ω(r) to denote the statement that r < Ks for an absolute constant K independent of the choice of the parameters underlying r and s. If we wish to indicate dependence of the constant on an additional parameter, we will subscript the notation appropriately. We let G be a locally compact group, and µG a left Haar measure on G. We normalize µG, i.e., scaling by a constant to get µG(G) = 1, when G is compact. For µG-measurable A ⊆ G and a constant ε, we set ε <ε StabG(A) = {g ∈ G : µG(A 4 gA) ≤ ε} and StabG (A) = {g ∈ G : µG(A 4 gA) < ε}.

Suppose A, B, and AB are µG-measurable sets in G. The discrepancy of A, B in G is defined by

dG(A, B) = µG(AB) − µG(A) − µG(B).

When G is connected, we always have dG(A, B) ≥ 0. Let H range over closed subgroups of G. We let µH denote a left Haar measure of H, and normalize µH when H is compact. Let G/H and H\G be the left coset space and the right coset space with quotient maps

π : G → G/H and πe : G → H\G Given a coset decomposition of G, say G/H, a left-fiber of a set A ⊆ G refers to A ∩ xH for some xH ∈ G/H. We also use µH to denote the fiber lengths in the −1 paper, that is, we sometimes write µH (A ∩ xH) to denote µH (x A ∩ H). By saying that G is Lie group, we mean G is a real Lie group with finite dimension, and we denote dim(G) the real dimension of G.

2. Outline of the argument In this section, we informally explain some of the main new ideas of the proofs. We decided to write a slightly longer outline as some of the later computations are rather technical. 2.1. Overview of the strategy. After handling a number of easy cases, the proofs of Theorem 1.1 and Theorem 1.2 require constructing appropriate continuous and surjective group homomorphisms into either R or T under the given data of a mini- mally or nearly minimally expanding pair (A, B) on a connected unimodular group G. The key difficulty of the problem is that many methods in the abelian settings (e.g., fourier analysis) has no generalization to the nonabelian settings which is obvi- ously useful for our purpose. We will get around this, essentially using an induction on dimension strategy to reduce to the abelian settings. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 9

Our argument can be broadly divided into three steps and a small preparation. For the preparation, we use a submodularity argument also used in [Tao18] and [CI21] to arrange that µG(A) = µG(B) are rather small when G is compact. In the first step, using ideas from the solution of Hilbert’s Fifth problem, arith- metic combinatorics, and model theory, we choose a compact and connected normal H of G such that G/H is a Lie group of bounded dimension and H is “in roughly the same direction” as A and B (i.e., the measures of the images of A and B in G/H are not too large compared to A and B). Then, by studying the geometric shape of A and B with respect to H, we prove a coarse version of our theorem: the minimally or nearly minimally expanding pair (A, B) arises from a minimally or nearly minimally expanding pair (A0,B0) on G/H. This also essentially reduces the problem to the case of Lie group, where we are aided by a powerful structure theory. The second step reduces the problems of constructing the desired group homomor- phism onto T or R to showing that dA(g1, g2) := µG(A) − µG(g1A ∩ g2A) satisfies 2 (2) dA(g1, g3) = |dA(g1, g2) ± dA(g2, g3)| + ε and dA(idG, g ) = 2dA(idG, g) + ε, for some error ε, and assuming dA(gi, gj) < µG(A)/4 for gi, gj ∈ {idG, g, g1, g2, g3}. Note that dA is a pseudometric (i.e., it satisfies all the properties of a metric except dA(g1, g2) = 0 implying g1 = g2). We remark that similar pseudometrics are also used in the solution of Hilbert’s Fifth problem. Moreover, <ε StabG (A) = {g ∈ G : dA(idG, g) < ε/2}.

Hence, looking at dA can be seen as an alternative way of considering approximate stabilizers, a recurring theme in the study of definable groups in model theory and approximate groups. In the third step, we assume that G is a Lie group. Using probabilistic and Lie- theoretic arguments, we choose a connected closed subgroup H of G with smaller dimension such that all the cosets of H intersect A and B “transversally in measure” (i.e. the intersection of A or B with each coset of H has small measure). When G is compact, H can be chosen to be a one-dimensional torus subgroups, so this can be seen as showing that A and B cannot be “Kakeya sets” in “torus subgroups directions”. We can assume H already satisfy the conclusion of Theorem 1.1 and Theorem 1.2 as an induction hypothesis. We then obtain a description of the geo- metric shapes of A and B relative to the cosets of H. Using that, we show that dA satisfies (2) which completes the proof. The remaining part of Section 2.1 will further elaborate ideas from the steps above. Sections 2.2, 2.3, and 2.4 and will explain further a number of technical innovations. The bulk of the second step is to produce a continuous group homomorphism onto R or T from the data of a pseudometric as in (2). Let us assume we already had a continuous surjective group homomorphism χ : G → T, and dT is the Euclidean MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 10

H xH T χ G

g1A ∃ H χ(g1A ∩ xH) g2A

g1A χ(g2A ∩ xH)

G g2A G/H

dA(g1, g2) = µG(g1A \ g2A) δx,A(g1, g2) = µT (χ(g1A ∩ xH) \ χ(g2A ∩ xH))

Figure 1. The third step

metric on T. Set d(g1, g2) = dT(χ(g1), χ(g2)). It is easy to see that d is a pseudometric on G with the “linear” property in (2) with ε = 0. In this case χ can be recovered from d by noticing that

ker χ = {g ∈ G : d(g, idG) = 0}. It turns out that it is also possible for ε > 0 as in (2), but it requires developing some nontrivial machinery, especially when ε grows linearly on the radius of the pseudometric; see Section 2.3 for details. Let us next consider the geometrical description of (A, B) in the third step and how it can be used to show (2). We suppose a “transversally in measure” H is already obtained. We visualize G in two ways: a rectangle with the horizontal side representing G/H and each of the vertical section representing a left coset of H, and a similar dual picture for H\G; see the middle figure of Figure 1. The main idea is to show that g1A and g2A geometrically (under this coset decomposition) look like in the picture with g1, g2 ∈ G in a suitable neighborhood of idG, and behave rigidly under translations. (For instance, we want the “fibers” of A and B, i.e. intersections of A or B with coset of H, to be preimages of intervals of T , and all the nonempty “fibers” of g1A to have similar “lengths”.) By induction hypothesis, we can assume that H satisfies the conclusions in the main theorems. In particular, the Euclidean metric on T induces a pseudometric δx,A on a generic fiber xH. The key point is that, the nice geometric shape of A can pass some of the properties of δx,A to the global pseudometric dA. The geometric idea in the first step is somewhat similar to what described above for the third step, so to end Section 2.1, let us explain how ideas from arithmetic combinatorics and model theory comes into play. Using an argument from [Tao08], we can produce from (A, B) a commensurable open approximate groups S. Then, by a result from Carolino’s thesis [Car15], which can be regarded as the continuous MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 11 version of the Lie model theorem from [Hru12] and [BGT12], we produce a continuous surjective group homomorphism to a Lie group with bounded dimension. The bound on dimension will play a role in determining what it means for (A, B) to have small measure in the third step, and contribute in the error bound in Theorem 1.2. 2.2. The first step: Nearly minimal expansions and quotients. Suppose (A, B) is a nearly minimally expanding pair on G, H is a connected, compact and normal subgroup of G, π : G → G/H is the quotient map, and

µG/H (π(A)) + µG/H (π(B) < 1. The goal of this step is to show the following quotient domination result: There is 0 0 −1 0 a nearly minimally expanding pair (A ,B ) on G/H such that µG(A 4 π (A )) and −1 0 µG(B 4 π (B )) are both small.

A2 H 0 (A ∪ A )2 \ A2 A2 1 0 1 0 A A0 2 2 (A0 ∪ A1 ∪ A2) \ (A0 ∪ A1) 2 1 1 A 2 2 A1 −→ ... A2 0 G/H w0 w1 w2 ... ≥ 2w0 π(A) 2 ≥ 2w0 + 2w1 π(A )

2 Figure 2. Lower bound for µG(A ).

To illustrate the idea, we focus on the special case with A = B. Employing the geometric language in Section 2.1, we call µG/H (π(A)) the width of A, for each g in −1 G, we call A∩gH a fiber of A, and refer to µH (g A∩H) as its length. We consider a SN further special case assuming that A can be partitioned into N +1 parts A = i=0 Ai such that the images under π of the Ai’s are compact and pairwise disjoint, Ai has width wi, the fibers in A0 all have length ≥ 1/2, the fibers in Ai all have the same length li ≤ 1/2 for each i ≥ 1, and li ≥ li+1 for all i < N. This further special case is, in fact, quite representative, as we can reduce the general problem to it using approximation techniques. The proof of this step can be seen as the following “spillover” argument. Applying 2 the Kemperman inequalities for H and G/H, we learn that all the fibers in A0 has 2 2 length 1, and the width of A0 is at least 2w0. By Fubini’s theorem, µG(A0) is at 2 least 2w0. Next, consider (A0 ∪ A1) . A similar argument gives us that all the fibers MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 12

2 2 in (A0 ∪ A1) has length at least 2l1, and the width of (A0 ∪ A1) in G/H is at least 2 2w0 + 2w1. Note that 2l1(2w0 + 2w1) is a weak bound for µG((A0 ∪ A1) ) since the 2 fibers in A0 are “exceptionally long”. Taking all of these into account, a stronger lower bound for µG(A0(A0 ∪ A1)) is

2w0 + 4l1w1. 2 Note that µG(A) = l0w0 + ... + lN wN . Hence, µG(A ) is nearly 2µG(A) implies that we must nearly have w0 = 1, w1 = ... = wN = 0, and l0 = 1. From this, one can deduce the conclusion that we want for this step. 2.3. The second step: Pseudometrics and homomorphisms. We assume in this section that G is a connected and compact Lie group, d is a left-invariant con- tinuous pseudometric on G with the following properties: (i) (Locally almost linearity) There is λ ∈ R>0 such that with

N(λ) := {g ∈ G : d(idG, g) < λ}, −6 there is ε < 10 µG(N(λ)) such that for all g1, g2, g3 ∈ N(λ)

d(g1, g3) ∈ |d(g1, g2) ± d(g2, g3)| + (−ε, ε).

(ii) (Locally almost monotonicity) With the same λ in (2), for all g ∈ Nλ, 2 |d(idG, g ) − 2d(idG, g)| ∈ (−ε, ε). We now sketch how to construct a continuous and surjective group homomorphism to T from these data. A crucial argument we will not be able to get into details here is to show that the local monotonicity condition can be deduced from a weaker property of path monotonicity obtained from the third step (Section 2.4). When we are in the special case with ε = 0 in property (i), there is a relatively easy argument which also works for noncompact Lie groups. Set

ker d = {g ∈ G : d(idG, g) = 0}. Using the left invariance, continuity, and triangle inequality, one can show that ker d is a closed subgroup of G. Moreover, in this case, G/ ker d must be isomorphic to T, and the pseudometric d locally must agree with a constant multiple of the pullback of the Euclidean metric. These are, perhaps, not too surprising as a Lie group equipped with a locally linear pseudometric is, intuitively, a very rigid object which locally looks like a straight line. In fact, property (ii) is not needed, as it is a consequence of property (i) in this case. The general case is much harder, as we no longer have the same type of rigidity. In particular, ker d might not be normal, and G/ ker d might not be T even if ker d is normal. The reader familiar with the proof of Hilbert’s Fifth problem would guess that we might try to slightly modify d to get a locally linear pseudometric MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 13 d0 and use the earlier strategy. This is still true at the conceptual level, but our actual argument is much more explicit, allowing error control. We will construct a multi-valued function Ω from G to R with additive properties. From this we get a multi-valued almost group homomorphism Ψ from G to T by quotienting ωZ for a value of ω ∈ R>0 to be described later. The desired group homomorphism χ can be obtained from Ψ using descriptive set theory and Riemannian geometry. For an element g ∈ G, we represent it in “the shortest way” as a product of elements in N(λ). The following notion capture this idea.

Definition 2.1. A sequence (g1, . . . , gm) with gi ∈ N(λ) is irreducible if gi+1 ··· gi+j ∈/ N(λ) for 2 ≤ j ≤ 4.

Now, we set Ωm(g) be the subset of R consists of sums of the form

m X (3) sgn(gi)d(gi, idG), i=1 where (g1, . . . , gm) is an irreducible sequence with g = g1 ··· gm, and sgn(gi) is the relative sign defined in Section 7. Heuristically, sgn(gi) specifies the “direction” of SM the translation by gi relative to a fixed element. Finally, we set Ω(g) = m=1 Ωm(g), where we will describe how to choose M in the next paragraph. We do not quite have Ω(gg0) = Ω(g) + Ω(g0), but a result along this line can be proven. Call the maximum distance between two elements of Ω(g) the error of Ω(g). We want this error to be small so as to be able to extract χ(g) from it eventually. In the construction, each d(gi, idG) will increase the error of the image of Ω(g) by at least ε. Since the error propagate very fast, to get a sharp exponent error bound, we cannot choose very large M. To show that a moderate value of M suffices, we construct a “monitoring system”, an irreducible sequence of bounded length, and every element in G is “close to” one of the elements in the sequence. The knowledge required for the construction amounts to an understanding of the size and expansion rate of small stabilizers of A, which can also be seen as a refined Sanders–Croot–Sisask-type result for nearly minimally expanding sets. In the aforementioned case where d is linear (i.e. ε = 0), the metric induced by d in G/ ker d is a multiple of the standard Euclidean metric in R/Z by a constant ω. If we apply the machinery of irreducible sequence to this special case that ε = 0, Ω(g) will be a single-valued function, and we will get for each g ∈ G that Ω(g) = ψ(g)+ωZ for ψ(g) ∈ [0, ω). In particular, Ω(idG) = ωZ, hence ω = inf Ω(idG) \{0} . In general case when ε > 0, we can define

ω = inf Ω(idG) \ (−Mε, Mε) . MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 14

By further applying properties of irreducible sequences and the monitor lemma, we show that for each g, Ω(g) ⊆ ψ(g)ωZ + (−Mε, Mε). For each g ∈ G, we then set Ψ(g) ⊆ T to be Ω(g)/ωZ. Note that Ψ is “continuous” from the way we construct it. We obtain χ from Ψ by first extracting from Ψ a universal measurable single valued almost-homomorphism ψ, then modifying ψ to get a universal measurable group homomorphism χ, and show that χ is automatically continuous. Many elements of our proof also work for noncompact group. However, the modifying step to get a group homomorphism from an almost group homomorphism does not go through due to the existence of quasi-morphism which are not close to group homomorphism in the noncompact case. 2.4. The third step: Nearly minimal expansions and subgroups. We focus on the case where G is a connected and compact Lie group, and (A, B) is a nearly minimally expanding pair on G with sufficiently small measure. Recall that we want −1 to find a closed and connected subgroup H of G such that the “length” µH (g A∩H) of each “left fiber” A∩gH of A is small, and a similar condition hold for “right fibers” of B. In this case, H can be chosen to be a one-dimensional torus subgroup. For simplicity, we assume A = B. Suppose A is a “Kakeya set”, i.e, it has a long fiber A ∩ gH for every choice of “direction” H of G. In Section 8.1 we will show that 2 under the condition when µG(A ) < MµG(A) for some constant M, for every H, µG(AH)/µG(A) is not too large. This motivates us to define the following notion: Definition 2.2. A is a toric K-expander if there is a one dimensional torus subgroup H of G such that µG(AH) > KµG(A). We will show that every nonempty compact subset of G with sufficiently small measure must be a toric K-expander. Let us present here a pseudo-argument, which nevertheless illustrate the idea. Assume A is not a toric K-expander. Obtain finitely many torus subgroups H1,...,Hn of G such that

G = H1 ··· Hn.

Let us pretend that using the assumption µG(AH1) ≤ KµG(A) we can cover AH1 with (K + 1) right translations of A. It can be then shown that AH1 is not a toric K(K + 1)-expander. Next, we further pretend that AH1H2 can be covered with K(K + 1) + 1 right translations of AT1 which can then be covered by K(K + 1)2 + (K + 1) right translations of A. Continuing the procedure, we get C(K) such that AT1 ··· Tn = G can be covered by C(K) right translations of A. Thus, µG(A) > 1/C(K), contradicting the assumption that µG(A) is very small. The pseudo-argument in the preceding paragraph does not work in most of the cases. In particular, one cannot deduce from µG(AH1) < KµG(A) that AH1 can be covered by finitely many right translations of A. However, it does contain some MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 15 truth, and we will be able to use a probabilistic argument to approximate this pseudo- argument. Now choosing a one-dimensional torus subgroup H of G such that for all x ∈ G and y ∈ G, the fibers xH ∩ A and B ∩ Hy are both short. We will show that the set A and B have the shape as described in Section 2.1. Without loss of generality, we can arrange that the width µG(AH) of A in G/H is at least the width µG(HB) of B in H\G. Choose uniformly at random xH ∈ AH, and applying the Kemperman inequality for H, we have

µG(AB) ≥ ExH∈AH µG((A ∩ xH)B) ≥ ExH∈AH µH (A ∩ xH)µH\G(HB) + µG(B)

µH\G(HB) = µG(A) + µG(B) µG/H (AH)

≥ µG(A) + µG(B);

As (A, B) is nearly minimally expanding, we have µG(AB) is nearly the same as µG(A) + µG(B). The fourth line then gives us that µG(HB) is nearly the same as µG(AH). The second line now gives us that for most of xH ∈ AH, the fiber (xH ∩A) is nearly an interval up to an endomorphism of H by using the induction hypothesis on H. From the first line, µH (A∩xH) is almost constant as xH ranges through AH. Using the above geometric properties of A and B together with some probabilistic arguments, in Section 8 we show that A also behaves rigidly under translations: A look like a horizontal “strip” under the coset decomposition, and the intersection of A and gA also behaves like a “strip”, where g is from a certain neighborhood of identity. Using this, we are able to pass the almost linearity of the pseudometric in H to the pseudometric dA defined in Section 2.1, as well as the path monotonicity which is a necessary ingredient to show assumption (2) in Step 2.

2.5. Structure of the paper. The paper is organized as follows. Section 3 includes some facts about Haar measures and unimodular groups, which will be used in the subsequent part of the paper. A general version of the Kemperman inequality and its inverse theorem in tori are also included in Section 3. Section 4 deals with the more immediate parts of Theorem 1.1 and hence sets up the stage for the main part of the argument. Section 5 allows us to arrange that in a minimally or a nearly minimally expanding pair (A, B), the sets A and B have small measure (Lemma 5.3). Sections 6, 7, and 8 contain main new technical ingredients of the proof, which will be put together in Section 9 to complete the proofs of Theorem 1.1, Theorem 1.2, and Theorem 1.3. Steps 1, 2, and 3 discussed in Sections 2.4, 2.3, and 2.2, corresponds to Sections 6, 7, and 8 respectively. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 16

In Section 6.1, we proved the quotient domination theorem (Theorem 6.5), which allow us to transfer the problem into certain quotient groups. In Section 6.2 we ob- tained a coarse version of the main theorems (Proposition 6.11). These two sections reduce the problem to a bounded dimension Lie group. Section 6.3 contains struc- tural results assuming we have an appropriate homomorphism (Proposition 6.15). We also give a new proof of the inverse Kneser’s inequality [Tao18] with a sharp ex- ponent bound (Theorem 6.14). In the next two sections, we will focus on constructing a suitable homomorphism. In Section 7.1, we showed that a locally linear pseudometric on G would induce a continuous surjective homomorphism to either R or T, with compact kernel (Propo- sition 7.6). Sections 7.2 and 7.3 study the locally almost linear pseudometric in compact Lie groups. In particular, we proved that path monotonicity implies mono- tonicity (Proposition 7.8), and for almost monotone almost linear pseudometric, one can also find a homomorphism mapping to T (Theorem 7.25). In Section 8.1, we show that given a small measure expansion set with small mea- sure, one can find a torus subgroup that is transversal in measure (Theorem 8.12). We construct the pseudometric from geometric properties of nearly minimal expan- sion sets in Sections 8.2 and 8.3. Section 8.2 provides a locally linear pseudometric from minimally expansion sets (Proposition 8.19). In Section 8.3, we construct a path monotone locally almost linear pseudometric (Proposition 8.23). The dependency diagram of the paper is as below.

Proposition 6.11 Proposition 6.15

Theorem 6.14 Theorem 8.12 Proposition 4.1

Lemma 5.3 Theorem 1.2 Theorem 1.1 Proposition 4.2

Theorem 6.5 Proposition 4.4 Proposition 8.23 Proposition 7.6 Proposition 8.19

Proposition 7.6 Theorem 7.25 Theorem 1.3 MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 17

3. Preliminaries Throughout this section, we assume that G is a connected locally compact group (in particular, Hausdorff) equipped with a left Haar measure µG, and A, B ⊆ G are nonempty. 3.1. Locally compact groups and Haar measures. Below are some basic facts about µG that we will use; see [DE09, Chapter 1] for details:

Fact 3.1. Suppose µG is either a left or a right Haar measure on G. Then:

(i) If A is compact, then A is µG-measurable and µG(A) < ∞. (ii) If A is open, then A is µG-measurable and µG(A) > 0. (iii) (Outer regularity) If A is µG-measurable, then there is a decreasing sequence (Un) of open subsets of G with A ⊆ Un for all n, and µG(A) = limn→∞ µG(Un). (iv) (Inner regularity) If A is µG-measurable, then there is an increasing sequence (Kn) of compact subsets of A such that µG(A) = limn→∞ µG(Kn). (v) (Measurability characterization) If there is an increasing sequence (Kn) of com- pact subsets of A, and a decreasing sequence (Un) of open subsets of G with A ⊆ Un for all n such that limn→∞ µG(Kn) = limn→∞ µG(Un), then A is mea- surable. 0 (vi) (Uniqueness) If µG is another measure on G satisfying the properties (1-5), >0 0 then there is C ∈ R such that µG = CµG. (vii) (Continuity of measure under symmetric difference) Suppose A ⊆ G is measur- able, then the function G → R, g 7→ µG(A 4 gA) is continuous. We remark that the assumption that G is connected implies that every measurable set is σ-finite (i.e., countable union of sets with finite µG-measure). Without the con- nected assumption, we only have inner regularity for σ-finite sets. From Fact 3.1(vii), we get the following easy corollary: ε Corollary 3.2. Suppose A is µG-measurable and ε is a constant. Then StabG(A) <ε 0 is closed in G, while StabG (A) is open in G. In particular, StabG(A) is a closed subgroup of G.

We say that G is unimodular if µG (and hence every left Haar measure on G) is also a right Haar measure. The following is well known and can be easily verified:

−1 Fact 3.3. If G is unimodular, A is µG-measurable, then A is also µG-measurable −1 and µG(A) = µG(A ). We use the following isomorphism theorem of topological groups. Fact 3.4. Suppose G is a locally compact group, H is a closed normal subgroup of G. Then we have the following. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 18

(i) (First isomorphism theorem) Suppose φ : G → Q is a continuous surjective group homomorphism with ker φ = H. Then the exact sequence of groups 1 → H → G → Q → 1 is an exact sequence of topological groups if and only if φ is open; the former condition is equivalent to saying that Q is canonically isomorphic to G/H as topological groups. (ii) (Third isomorphism theorem) Suppose S ≤ G is closed, and H ≤ S. Then S/H is a closed subgroup of G/H. If S C G is normal, then S/H is a normal subgroup of G/H, and we have the exact sequence of topological groups 1 → S/H → G/H → G/S → 1; this is the same as saying that (G/H)/(S/H) is canonically isomorphic to G/S as topological groups. Suppose H is a closed subgroup of G. The following fact allows us to link Haar measures on G with the Haar measures on H for unimodular G and H: Fact 3.5 (Quotient formula). Suppose H is a closed subgroup of G with a left Haar measure µH . If f is a continuous function on G with compact support, then Z xH 7→ f(xh) dµH (x). H defines a function f H : G/H → R which is continuous and has compact support. If both G and H are unimodular, then there is unique invariant Radon measures µG/H on G/H such that for all continuous function f : G → R with compact support, the following integral formula holds Z Z Z f(x) dµG(x) = f(xh) dµH (h) dµG/H (xH). G G/H H A similar statement applies replacing the left homogeneous space G/H with the right homogeneous space H\G. We can extend Fact 3.5 to measurable functions on G, but the function f H in the statement can be only be defined and is µG/H -measurable µG-almost everywhere. So, in particular, this problem applies to indicator function 1A of a measurable set A. This causes problem in our later proof and prompts us to sometimes restrict our attention to a better behaved subcollection of measurable subsets of G. We say that a subset of G is σ-compact if it is a countable union of compact subsets of G. Lemma 3.6. We have the following: (i) σ-compact sets are measurable. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 19

(ii) the collection of σ-compact sets is closed under taking countable union, taking finite intersection, and taking product set. 0 (iii) For all µG-measurable A, we can find a σ-compact subset A of A such that 0 µG(A ) = µG(A). (iv) Suppose G is unimodular, H is a closed subgroup of G with a left Haar measure µH , A ⊆ G is σ-compact, and 1A is the indicator function of A. Then aH 7→ 1H µH (A ∩ aH) defines a measurable function A : G/H → R. If H is unimodular and, µG/H is the given in Fact 3.5, then Z Z µG(A) = µH (A ∩ aH) dµH (h) dµG/H (xH). G/H H A similar statement applies replacing the left homogeneous space G/H with the right homogeneous space H\G. Proof. The verification of (i-iii) is straightforward. We now prove (iv). First consider the case where A is compact. By Baire’s Theorem, 1A is the pointwise limit of a monotone nondecreasing sequence of continuous function of compact support. If f : G → R is a continuous function of compact support, then the function Z f H : G/H → R, aH 7→ f(ax)dx H is continuous with compact support, and hence measurable; see, for example, [DE09, R 1 Lemma 1.5.1]. Noting that µH (A∩aH) = H A(ax)dx, and applying monotone con- 1H vergence theorem, we get that A is the pointwise limit of a monotone nondecreasing sequence of continuous function of compact support. Using monotone convergence 1H theorem again, we get A is integrable, and hence measurable. Also, by monotone convergence theorem, we get the quotient integral formula in the statement. Finally, the general case where A is only σ-compact can be handled similarly, noting that 1A is then the pointwise limit of a monotone nondecreasing sequence of indicator functions of compact sets.  Suppose H is a closed subgroup of G. Then H is locally compact, but not neces- sarily unimodular. We use the following fact in order to apply induction arguments in the later proofs. Fact 3.7. Let G be a unimodular group. If H is a closed normal subgroup of G, then H is unimodular. Moreover, if H is compact, then G/H is unimodular. 3.2. More on Kemperman’s inequality and the inverse problems. We will need a version of Kemperman’s inequality for arbitrary sets. Recall that the inner Haar measure µeG associated to µG is given by µeG(A) = sup{µG(K): K ⊆ A is compact.} MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 20

The following is well known and can be easily verified:

Fact 3.8. Suppose µeG is the inner Haar measure associated to µG. Then we have the following: (i) (Agreement with µG) If A is measurable, then µeG(A) = µ(A). (ii) (Inner regularity) There is σ-compact A0 ⊆ A such that 0 µeG(A) = µeG(A ) = µG(A). (iii) (Superadditivity) If A and B are disjoint, then

µeG(A ∪ B) ≥ µeG(A) + µeG(B). (iv) (Left invariance) For all g ∈ G, µeG(gA) = µe(A). (v) (Right invariance) If G is unimodular, then for all g ∈ G, µeG(Ag) = µe(A). It is easy to see that we can replace the assumption that A and B are compact in Kemperman’s inequality in the introduction with the weaker assumption that A and B are σ-compact. Together with the inner regularity of µeG (Fact 3.8.2), this give us the first part of the following Fact 3.9. The second part of Fact 3.9 follows from the fact that taking product sets preserves compactness, σ-compactness, and analyticity. Note that taking product sets in general does not preserve measurability, so we still need inner measure in this case.

Fact 3.9 (Generalized Kemperman inequality for connected groups). Suppose µeG is the inner Haar measure on G, and A, B ⊆ G are nonempty. Then

µeG(AB) ≥ min{µeG(A) + µeG(B), µeG(G)}. Moreover, if A and B are compact, σ-compact, or analytic, then we can replace µeG with µG. The remaining parts of Theorem 1.1 consist of classifying the minimally expanding pairs (A, B) and show that they match the description in situations (iii) and (iv) of Theorem 1.1. For compact group, our strategy is to reduce the problem to the known situations of one dimensional tori. Hence, we need the following special case of Kneser’s classification result, and the sharp dependence between ε and δ is essentially due to Bilu [Bil98]. Fact 3.10 (Inverse theorem for Td). Let A, B be compact subsets of Td. For every τ > 0, there is a constant c = c(τ) such that if −1 τ µTd (A) ≤ µTd (B) ≤ µTd (A) ≤ c, then either µTd (A + B) ≥ µTd (A) + 2µTd (B), or there are compact intervals I,J in T with µT(I) = µTd (A + B) − µTd (B) and µT(J) = µTd (A + B) − µTd (A), and a continuous surjective group homomorphism χ : Td → T, such that A ⊆ χ−1(I) and B ⊆ χ−1(J). MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 21

When the group is a one dimensional torus T, a sharper result is recently obtained −1549 by Candela and de Roton [CDR19]. The constant cG = 3.1 · 10 in the following fact is from an earlier result in Z/pZ by Grynkiewicz [Gry13].

Fact 3.11 (Inverse theorem for T). There is a constant cG > 0 such that the following holds. If A, B are compact subsets of T, with dT(A, B) < min{µT(A), µT(B), cG}, and 1 − µT(A) − µT(B) ≥ 2dT(A, B). Then there is a continuous surjective group homomorphism χ : T → T, and two compact intervals I,J ⊆ T such that

µT(I) ≤ µT(A) + dG(A, B), µT(J) ≤ µT(B) + dG(A, B), and A ⊆ χ−1(I), B ⊆ χ−1(J). For noncompact group, we reduce the problem to the known situation of additive group of real numbers. The following result can be seen as the stability theorem of the Brunn–Minkowski inequality in Rd when d = 1.

Fact 3.12 (Inverse theorem for R). Let A, B be compact subsets in R with µG(A) ≥ µG(B), and let µR be the Lebesgue measure in R. Suppose we have

µR(A + B) < µR(A) + 2µR(B).

Then there are compact intervals I,J ⊆ R with µR(I) = µR(A + B) − µR(B) and µR(J) = µR(A + B) − µR(A), such that A ⊆ I and B ⊆ J. 4. Reduction to classifying nearly minimally expanding pairs To set the stage for the later discussion, we would like to separate the core part of Theorem 1.1 from the more immediate parts. Throughout G is a connected uni- modular group, and µG is a Haar measure on G. Proposition 4.1. Suppose A, B ⊆ G are nonempty and compact and one of the situation listed in Theorem 1.1 holds, then µG(AB) = min{µG(A) + µG(B), µG(G)}. Proof. We will only consider situation (v) because (i-iv) are immediate and (vi) can be showed in a same way as (v). Suppose we are in situation (v) of Theorem 1.1. As χ is a group homomorphism, we have AB = χ−1(I + J). Note that by quotient integral formula, we have µG(A) = µT(I), µG(B) = µT(J), µG(AB) = µT(I + J). The desired conclusion follows from the easy that µT(I + J) = µT(I) + µT(J).  The following lemma clarifies the second statement in situation (ii) of Theorem 1.1. Note that we do not have the later part of Theorem 1.1(ii) when A, B are not assumed to be compact. For example with A = B = (0, 1/2)+Z, we have µT(A)+µT(B) = 1, but AB = (0, 1) + Z is not the whole T.

Proposition 4.2. If either A, B ⊆ G are measurable and µG(A) + µG(B) > µG(G) or A, B ⊆ G are compact and µG(A) + µG(B) ≥ µG(G), then AB = G. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 22

Proof. The cases where either A or B is empty are immediate, so we assume that A, B are nonempty. Suppose g is an arbitrary element of G. It suffices to show that −1 −1 A g and B has nonempty intersection. As G is unimodular, µG(A) = µG(A ) by −1 −1 Fact 3.3. Hence µG(A g) + µG(B) = µG(G). If µG(A g ∩ B) > 0, then we are −1 −1 done. Otherwise, we have µG(A g ∩ B) = 0, and so µG(A g ∪ B) = µG(G) by the inclusion-exclusion principle. As A and B are compact, A−1g ∪ B is also compact, and the complement of A−1g ∪ B is open. Since nonempty open sets has positive −1 −1 measure, µG(A g ∪ B) = µG(G) implies A g ∪ B = G. Now, since G is connected, −1 we must have A g ∩ B must be nonempty.  The following corollary will be used many times later.

Corollary 4.3. If either A ⊆ G is measurable and nµG(A) > µG(G) or A ⊆ G is n compact and nµG(A) > µG(G), then A = G. Proof. By replacing A with a subset with the same measure if necessary, we can as- sume that A is σ-compact. Then by the generalized Kemperman inequality (Fact 3.9), n−1 we get µG(A ) + µG(A) ≥ µG(G). Applying Proposition 4.2, we get the desired conclusion.  Now we clarify the situation in (iii) of Theorem 1.1, situation (iv) can be proven in the same way.

Proposition 4.4. Suppose A, B ⊆ G are nonempty, compact, and with µG(A) = 0, 0 < µG(B) < µG(G), and

µG(AB) = min(µG(A) + µG(B), µG(G)). Then there is a compact subgroup H of G such that A ⊆ gH for some g ∈ G, and B = B1 ∪ B2 with HB1 = B1 and µG(AB2) = 0.

Proof. Without loss of generality, we can assume that A and B both contain idG. 0 Let H = StabG(B), let B1 be the set of b ∈ B such that whenever U is an open neighborhood of b, we have µG(U ∩ B) > 0, and let B2 = B \ B1. We will now verify that H, B1, and B2 are as desired. We make a number of immediate observations. As µG(AB) = µG(B) and idG is in A, we must have A ⊆ H. Note that B2 consists of b ∈ B such that there is open neighborhood U of B with µG(U ∩ B) = 0. So B2 is open in B. Moreover, if K is n a compact subset of B2, then K has a finite cover (U)i=1 such that µG(Ui ∩ B) = 0, which implies µG(K) = 0. It follows from inner regularity of µG (Fact 3.1(iv)), that µG(B2) = 0. Hence, B1 is a closed subset of B with µG(B1) = µG(B). Since B is compact, B1 is also compact. As H is a closed subset of G, the compactness of H follows immediately if we can show that HB1 = HB MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 23

It remains to verify that HB1 = B1. As µG is both left and right translation invariant, we also have that for all g ∈ G, if U is an open neighborhood of gb ∈ gB1, then µG(U ∩ gB1) > 0. Suppose hb is in HB1 \ B1 with h ∈ H. Set U = G \ B1. Then U is an open neighborhood of hb. From the earlier discussion, we then have µG(U ∩ hB1) > 0. This implies that µG(hB \ B) > 0 contradicting the fact that 0 h ∈ H = StabG.  5. Reduction to sets with small measure

Throughout this section, G is a connected compact group, µG is the normalized Haar measure on G, and A, B ⊆ G are σ-compact sets with positive measure. We will show that if (A, B) is nearly minimally expanding in G, then we can σ-compact A0 and B0 each with smaller measure such that the pair (A0,B0) is also nearly minimally expanding. The similar approach used in this section is introduced by Tao [Tao18] and used to obtain an inverse theorem in the abelian setting. We first prove the following easy fact, which will be used several times later in the paper. Let f, g : G → C be functions. For every x ∈ G, we define the convolution of f and g to be Z −1 f ∗ g(x) = f(y)g(y x) dµG(y). G Note that f ∗ g is not commutative, but associative by Fubini’s Theorem. 2 Lemma 5.1. Let t be any real numbers such that µG(A) ≤ t ≤ µG(A). Then there are x, y ∈ G such that µG(A ∩ (xA)) = µG(A ∩ (Ay)) = t. Proof. Consider the maps:

π1 : x 7→ 1A ∗ 1A−1 (x) = µG(A ∩ (xA)), and π2 : y 7→ 1A−1 ∗ 1A(y) = µG(A ∩ (Ay)).

By Fact 3.1, both π1 and π2 are continuous functions, and equals to µG(A) when x = y = idG. By Fubini’s theorem, 1 1 2 1 1 E ( A ∗ A−1 ) = µG(A) = E ( A−1 ∗ A). Then the lemma follows from the intermediate value theorem, and the fact that G is connected. 

Recall that dG(A, B) = µG(AB) − µG(A) − µG(B) is the discrepancy of A and B on G. The following property is sometimes referred to as submodularity in the literature. Note that this is not related to modular functions in locally compact groups or the notion of modularity in model theory.

Lemma 5.2. Let γ1, γ2 > 0, and A, B1,B2 are σ-compact subsets of G. Suppose that dG(A, B1) ≤ γ1, dG(A, B2) ≤ γ2, and

µG(B1 ∩ B2) > 0, and µG(A) + µG(B1 ∪ B2) ≤ 1. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 24

Then both dG(A, B1 ∩ B2) and dG(A, B1 ∪ B2) are at most γ1 + γ2. Proof. Observe that for every x ∈ G we have 1 1 1 1 AB1 (x) + AB2 (x) ≥ A(B1∩B2)(x) + A(B1∪B2)(x), which implies

(4) µG(AB1) + µG(AB2) ≥ µG(A(B1 ∩ B2)) + µG(A(B1 ∪ B2)).

By the fact that dG(A, B1) ≤ γ1 and dG(A, B2) ≤ γ2, we obtain

µG(AB1) ≤ µG(A) + µG(B1) + γ1, and µG(AB2) ≤ µG(A) + µG(B2) + γ2. Therefore, by equation (4) we have

µG(A(B1 ∩ B2)) + µG(A(B1 ∪ B2))

≤ 2µG(A) + µG(B1 ∩ B2) + µG(B1 ∪ B2) + γ1 + γ2.

On the other hand, as µG(B1 ∩ B2) > 0 and µG(A) + µG(B1 ∪ B2) ≤ 1, and using Kemperman’s inequality, we have

µG(A(B1 ∩ B2)) ≥ µG(A) + µG(B1 ∩ B2), and

µG(A(B1 ∪ B2)) ≥ µG(A) + µG(B1 ∪ B2). This implies

µG(A(B1 ∩ B2)) ≤ µG(A) + µG(B1 ∩ B2) + γ1 + γ2, and

µG(A(B1 ∪ B2)) ≤ µG(A) + µG(B1 ∪ B2) + γ1 + γ2.

Thus we have dG(A, B1 ∩ B2), dG(A, B1 ∪ B2) ≤ γ1 + γ2.  The following lemma is the main result of this section, it says if G admits a small expansion pair, one can another find pair of sets with sufficiently small measures, and still has small expansion.

Lemma 5.3. Let d1, d2 ∈ (0, 1/4) be positive real numbers, and let

0 < λ = min{µG(A), µG(B), 1 − µG(A) − µG(B)}. 0 0 Suppose dG(A, B) ≤ γ. Then there are σ-compact sets A ,B ⊆ G satisfying 0 0 (i) µG(A ) = d1 and µG(B ) = d2, 0 0 0 0 (ii) dG(A ,B), dG(A, B ), and dG(A ,B ) are Od1,d2 (γ/m). MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 25

Proof. We assume µG(A) > d1 and µG(A) ≥ µG(B). The case when µG(A) is less than d1 can be proven in a similar way by replacing taking intersections by taking unions. For every g ∈ G, both dG(gA, B) and dG(A, Bg) are still upper bounded 2 by γ. By Lemma 5.1, for every t with µG(A) ≤ t ≤ µG(A), there is g ∈ G such that µG(A ∩ gA) = t. Assuming that in each step, we can choose g such that 2 µG(A ∩ gA) = µG(A) , and replace A by A ∩ gA. Hence, after O(log log 1/d1) steps, the measure of A will achieve d1. The issue of this simple argument is that we may have µG(A ∪ gA) + µG(B) > 1 when µG(A) ≥ 1/3, so that we cannot apply Lemma 5.2. Thus, in the first few steps, we will choose g such that µG(A ∪ gA) is not too large. 2 We first consider the case when µG(A) ≥ 1/3, and µG(A) − λ ≥ µG(A) . We are going to choose g ∈ G such that

(5) 2µG(A) − µG(A ∩ gA) + µG(B) = µG(A ∪ gA) + µG(B) ≤ 1, and µG(A∩gA) ≥ max{d1, µG(A)−λ}. Such g exists by Lemma 5.1. Let A1 = A∩gA, 2 then µG(A1) ≤ µG(A) − λ, µG(A) . By Lemma 5.2, dG(A1,B) ≤ 2γ. Next we choose 2 g1 ∈ G satisfying (5) with A replaced by A1, and µG(A1 ∩g1A1) ≥ min{d1, µG(A1) }. Let A2 = A1 ∩ g1A1, then 2 µG(A2) ≤ max{µG(A1) − 2λ, µG(A1) }, and dG(A2,B) ≤ 4γ. Repeat this procedure for t1 steps until either µG(At1 ) = d1, or t−1 2 µG(At) − 2 λ ≤ µG(At) . In either case, we have t1 ≤ log(1/3λ). 2 Next, if µG(At1 ) > d1, we choose gt1 in G such that µG(At1 ∩gt1 At1 ) = µG(At1 ) . By the way we define t1, we have µG(At1 ∪gt1 At1 )+µG(B) ≤ 1. Set At1+1 = At1 ∩gt1 At1 .

Repeat this procedure for t2 steps until µG(At1+t2 ) = d1. We have

log d1 1 t2 ≤ log ≤ log log , log µG(At1 ) d1

t1+t2 and dG(At1+t2 ,B) ≤ 2 γ = Od1 (γ/λ). We then apply the same procedures for B to arrange B having measure d2. If we have µG(A) < 1/3 at the beginning, we are able to choose g such that 2 µG(A ∩ gA) = µG(A) and µG(A ∪ gA) + µG(B) ≤ 1. Hence, it only requires at most log log(1/d1) steps to make A having measure d1. 

6. Geometry of minimal and nearly minimal expansions I This section studies the shape of a nearly minimally expanding pairs relative to a connected compact normal subgroup of the ambient such that the images of the pair under the quotient map have small measure. In Section 6.1, we obtain results that will allow us to reduce the Theorem 1.1 and Theorem 1.2 to analogous result about a simpler . Section 6.2 applies Section 6.1 to MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 26 reduce Theorem 1.1 and Theorem 1.2 to the case of Lie groups and also prove a coarse version of these results. Section 6.3 applies Section 6.1 to further reduce Theorem 1.1 and Theorem 1.2 to the problem of constructing suitable group homomorphism onto either T or R. Throughout this section, G is a connected unimodular locally compact group with Haar measure µG, and A and B are σ-compact subsets of G with positive µG-measure. We will assume familiarity with the preliminary Section 3.1 on locally compact group and Haar measure. 6.1. Preservation of minimal expansion under quotient. In this section, H is a connected compact normal subgroup of G, so H and G/H are unimodular by Fact 3.7. Let µH , and µG/H be the Haar measure on G, H, and G/H, and let µeG and µeG/H be the inner Haar measures on G and G/H. Suppose r and s are in R. We set

A(r,s] := {a ∈ A : µH (A ∩ aH) ∈ (r, s]} and

πA(r,s] := {aH ∈ G/H : µH (A ∩ aH) ∈ (r, s]}.

In particular, πA(r,s] is the image of A(r,s] under the map π. We define B(r0,s0] and 0 0 πB(r0,s0] likewise for r , s ∈ R. We have a number of immediate observations. 0 0 >0 Lemma 6.1. Let r, s, r , s be in R . For all aH ∈ πA(r,s], bH ∈ πB(r0,s0], the sets A(r,s] ∩ aH, B(r0,s0] ∩ bH are nonempty σ-compact. For all subintervals (r, s] of (0, 1], A(r,s] is µG-measurable and πA(r,s] is µG/H -measurable.

Proof. The first assertion is immediate from the definition. Let 1A be the indicator function of A. Then the function 1H A : G/H → R

aH 7→ µH (A ∩ aH) 1H −1 is well-defined and measurable by Lemma 3.6. As πA(r,s] = ( A ) (r, s] and −1 A(r,s] = A ∩ π (πA(r,s]), we get the second assertion. 

Note that πA(r,s]πB(r0,s0] is not necessarily µG/H -measurable, so Lemma 6.2(ii) does require the inner measure µeG/H . Lemma 6.2. We have the following: (i) For every aH ∈ πA and bH ∈ πB,  µH (A ∩ aH)(B ∩ bH) ≥ min{µH (A ∩ aH) + µH (B ∩ bH), 1}. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 27

(ii) If A(r,s] and B(r0,s0] are nonempty, then

µeG/H (πA(r,s]πB(r0,s0]) ≥ min{µG/H (πA(r,s]) + µG/H (πB(r0,s0]), µG/H (G/H)}. Proof. Note that both H and G/H are connected. So (i) is a consequence of the generalized Kemperman inequality for H (Fact 3.9) and (ii) is a consequence of the generalized Kemperman inequality for G/H (Fact 3.9).  As the functions we are dealing with are not differentiable, we will need Riemann– Stieltjes integral which we will now recall. Consider a closed interval [a, b] of R, and n functions f :[a, b] → R and g : R → R.A partition P of [a, b] is a sequence (xi)i=0 of real numbers with x0 = a, xn = b, and xi < xi+1 for i ∈ {0, . . . , n − 1}. For such n−1 P , its norm kP k is defined as maxi=0 |xi+1 − xi|, and a corresponding partial sum Pn is given by S(P, f, g) = i=0 f(ci+1)(g(xi+1) − g(xi)) with ci+1 ∈ [xi, xi + 1]. We then define Z b f(x) dg(x) := lim S(P, f, g) a kP k→0 if this limit exists where we let P range over all the partition of [a, b] and S(P, f, g) ranges over all the corresponding partial sums of P . The next fact records some basic properties of the integral. Fact 6.3. Let [a, b], f(x), and g(x) be as above. Then we have: (i) (Integrability) If f(x) is continuous on [a, b], and g(x) is monotone and bounded on [a, b], then f(x) dg(x) is Riemann–Stieltjes integrable on [a, b]. (ii) (Integration by parts) If f(x) dg(x) is Riemann–Stieltjes integrable on [a, b], then g(x) df(x) is also Riemann–Stieltjes integrable on [a, b], and Z b Z b f(x) dg(x) = f(b)g(b) − f(a)g(a) − g(x) df(x). a a The next lemma uses “spillover” estimate, which gives us a lower bound estimate on µG(AB) when the projection of A and B are not too large.

Lemma 6.4. Suppose µG/H (πA) + µG/H (πB) < 1. Set α = supa∈A µH (A ∩ aH), β = supb∈B µH (B ∩ bH), and γ = max{1, α + β}. Then α + β µ (AB) ≥ µ (πA ) + µ (πB ) G γ G/H (α/γ,α] G/H (β/γ,β] α + β α + β + µ (A ) + µ (B . α G (0,α/γ] β G (0,β/γ]) −1 Proof. For x ∈ (0, 1], set Cx = AB ∩ π (πA(xα,α]πB(xβ,β]). One first note that

µG(AB) ≥ µeG(C0). MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 28

By Fact 6.3(1), dµeG(Cx) is Riemann–Stieltjes integrable on any closed subinterval of [0, 1]. Hence, 1 Z γ µeG(C0) = µeG(C1/γ) − dµeG(Cx). 0 Lemma 6.1 and Lemma 6.2(1) give us that

µeG(C1/γ) ≥ µeG/H (πA(α/γ,α]πB(β/γ,β]). >0 Likewise, for x, y ∈ R with x < y ≤ 1/γ, µeG(Cx) − µeG(Cy) is at least  r(α + β) µeG/H (πA(xα,α]πB(xβ,β]) − µeG/H (πA(yα,α]πB(yβ,β]) . Therefore, 1 Z γ µeG(C0) ≥ µeG/H (πA(α/γ,α]πB(β/γ,β]) − (α + β)x dµeG/H (πA(xα,α]πB(xβ,β]). 0 Using integral by parts (Fact 6.3.2), we get 1 Z γ µeG(C0) ≥ µeG/H (πA(xα,α]πB(xβ,β]) d(α + β)x. 0

Applying Lemma 6.2.2 and using the assumption that µG/H (πA) + µG/H (πB) < 1 , we have 1 Z γ µeG(C0) ≥ (µG/H (πA(xα,α]) + µG/H (πB(xβ,β])) d(α + β)x. 0 Using integral by parts (Fact 6.3.2), we arrive at α + β µ (C ) ≥ µ (πA ) + µ (πB ) eG 0 γ G/H (α/γ,α] G/H (β/γ,β] 1 Z γ − (α + β)x d(µG/H (πA(xα,α]) + µG/H (πB(xβ,β])). 0

As d(µG/H (πA(xα,α]) + µG/H (πB(xβ,β])) = − d(µG/H (πA(0,xα]) + µG/H (πB(0,xβ])), α + β µ (C ) ≥ µ (πA ) + µ (πB ) eG 0 γ G/H (α/γ,α] G/H (β/γ,β] 1 Z γ + (α + β)x d(µG/H (πA(0,xα]) + µG/H (πB(0,xβ])). 0 Finally, recall that Z 1/γ Z 1/γ xα dµG/H (πA(0,xα]) = µG(A(0,α/γ]) and βx dµG/H (πB(0,xβ]) = µG(B(0,β/γ]). 0 0 MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 29

Thus, we arrived at the desired conclusion.  The next result in the main result in this subsection. It says if the projections of A and B are not too large, the small expansion properties will be kept in the quotient group.

Theorem 6.5 (Quotient domination). Suppose µG/H (πA)+µG/H (πB) < µG/H (G/H) 0 0 and dG(A, B) < min{µG(A), µG(B)}. Then there are σ-compact A ,B ⊆ G/H such that 0 0 dG/H (A ,B ) < 7dG(A, B) −1 0 −1 0 and max{µG(A 4 π A ), µG(B 4 π B )} < 3dG(A, B). Proof. Let α and β be as in Lemma 6.4. We first show that α + β ≥ 1. Suppose to the contrary that α + β < 1. Then Lemma 6.4 gives us α + β α + β µ (AB) ≥ µ (A) + µ (B) G α G β G

It follows that µG(AB) > µG(A) + µG(B) + min{µG(A), µG(B)}, a contradiction. Now we have α + β ≥ 1. Hence, Lemma 6.4 yields

µG(AB) ≥ µG/H (πA(α/(α+β),α]) + µG/H (πB(β/(α+β),β]) α + β α + β + µ (A ) + µ (B . α G (0,α/γ] β G (0,β/(α+β)]) 0 0 Choose σ-compact A ⊆ πA(α/(α+β),α] and B ⊆ πB(β/(α+β),β]) σ-compact such that 0 0 µG/H (A ) = µG/H (πA(α/(α+β),α]) and µG/H (B ) = µG/H (πB(β/(α+β),β]). We will verify that A0 and B0 satisfy the desired conclusion. 0 0 Since µG/H (A ) ≥ (1/α)µG(A(α/(α+β),α]), µG/H (B ) ≥ (1/β)µG(B(β/(α+β),β]) and α + β > 1, we have 1 1 µ (AB) ≥ µ (A) + µ (B). G α G β G

From µG(AB) − µG(A) − µG(B) = dG(A, B) ≤ min{µG(A), µG(B)}, we deduce that α, β ≥ 1/2. By our assumption µG(AB) < µG(A) + µG(B) + dG(A, B). Hence, 0 0 dG(A, B) ≥ µG/H (A ) − µG(A(α/(α+β),α]) + µG/H (B ) − µG(B(β/(α+β),β]) β α + µ (A ) + µ (B . α G (0,α/γ] β G (0,β/(α+β)]) 0 Therefore, µG/H (A ) − µG(A(α/(α+β),α]) and (β/α)µG(A(0,α/γ]) are at most dG(A, B). −1 0 Noting also that β/α ≤ 1/2, we get µG(A 4 π (A ) ≤ 3dG(A, B). A similar argu- −1 0 ment yield µG(B 4 π (B ) ≤ 3dG(A, B). MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 30

−1 0 0 Finally, note that π (A B ) is equal to A(α/(α+β),α]B(β/(α+β),β], which is a subset of AB. Combining with µG(AB) < µG(A) + µG(B) + dG(A, B), we get 0 0 0 0 µG/H (A B ) ≤ µG(A) + µG(B) + dG(A, B) ≤ µG/H (A ) + µG/H (B ) + 7dG(A, B), which completes the proof.  The next corollary of the proof of Theorem 6.5 gives a complementary result when without the assumption that µG/H (πA) + µG/H (πB) < µG/H (G/H).

Corollary 6.6. Suppose G is noncompact and dG(A, B) = 0. Then there are 0 0 0 0 −1 0 σ-compact A ,B ⊆ G/H such that dG/H (A ,B ) = 0, µG(A 4 π A ) = 0, and −1 0 µG(B 4 π B ) = 0.

Proof. Choose an increasing sequence (An) of compact subsets of A and an increasing S∞ S∞ sequence (Bn) of compact subsets of B such that A = n=0 An and B = i=0 Bn. Then limn→∞ dG(An,Bn) = 0. For each n, An and Bn are compact, so πAn and πBn 0 0 are also compact and has finite measure. Let An and Bn be defined for An and Bn as in the proof of Theorem 6.5. Then for n sufficiently large, we have −1 0 −1 0 µG(π An 4 An) < 3dG(An,Bn) and µG(π Bn 4 Bn) < 3dG(An,Bn) and 0 0 0 0 µG/H (AnBn) < µG/H (An) + µG/H (Bn) + 5dG(An,Bn). 0 0 Moreover, we can arrange that the sequences (An) and (Bn) are increasing. Take 0 S∞ 0 0 S∞ 0 A = n=1 An and B = n=1 Bn. By taking n → ∞, we have −1 0 −1 0 µG(π A 4 A) = 0 and µG(π B 4 B) = 0. 0 0 and dG/H (A ,B ) = 0 as desired.  6.2. Coarse versions of the main theorems. For the given G, there might be no continuous surjective group homomorphism to either T or R (e.g. G = SO3(R)). However, the famous theorem below by Gleason [Gle52] and Yamabe [Yam53] allows us to naturally obtain continuous and surjective group homomorphism to a Lie group. Using together with Corollary 6.6, this allows us to reduce the noncompact case of Theorem 1.1 to that of Lie group. The connectedness of H is not often stated as part of the result, but can be arranged by replacing H with its identity component. Fact 6.7 (Gleason–Yamabe Theorem). For any connected locally compact group G and any neighborhood U of the identity in G, there is a connected compact normal subgroup H ⊆ U of G such that G/H is a Lie group.

With some further effort, we can also arrange that µG/H (πA) + µG/H (πB) < µG/H (G/H) as necessary to apply Theorem 6.5. However, when dG(A, B) > 0, we will need a dimension control on the Lie group we obtained from the Gleason–Yamabe MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 31

Theorem. For that, we need Fact 6.9, which can be thought of as a refinement of the Gleason–Yamabe theorem coming from arithmetic combinatorics and model theory. Recall that an open and precompact set S ⊆ G is a K-approximate group if −1 2 idG ∈ S, S = S, and S ⊆ XS for some finite set X of cardinality K. The next theorem by Tao [Tao08] allows us to extract an approximate group from a nearly minimally expanding pair; as stated in [Tao08], this theorem is only applicable when A, B are open, but the proof also goes through without this assumption. Fact 6.8 (Approximate groups from small expansion). Suppose K is a constant and 1/2 1/2 O(1) µG(AB) < KµG (A)µG (B), then there is an open precompact O(K )-approximate O(1) 1/2 1/2 group S with µG(S) = O(K )µG (A)µG (B) and a finite set X or cardinality O(KO(1)) such that A ⊆ XS and B ⊆ SX. The study of continuous approximate groups by Carolino [Car15] is able to find a Lie model, and to control the dimension of the Lie model. This can be seen as a finer version of the Gleason–Yamabe theorem. Fact 6.9 (Lie model from approximate groups). Suppose K is a constant and S is an open precompact K-approximate group on G. Then there is a connected compact normal subgroup H of G, such that H ⊆ S4 and G/H is a Lie group of dimension OK (1). Fact 6.9 can also be deduced from the main theorem in [MW15] and the strong approximate group theory in [BGT12]. The following sketch was explained to us by Arturo Rodriguez Fanlo. We assume the reader is sufficiently familiar with model theory and the results in [BGT12] Sketch of proof of Fact 6.9. Suppose to the contrary. Take an ultraproduct (G, S) of counterexamples (Gn,Sn). By the main result in [MW15], using the ultraproduct of the Haar measures, G00 is contained in S4, so this ultraproduct has a Lie model L = G/H. Now, using a result in [BGT12], namely, the strong approximate groups theory, 0 4 we can find a definable strong approximate group S ⊆ S such that OK (1) left- translates of S0 cover S. ByLo´s’theorem, for sufficiently large n, this gives us 0 4 strong approximate groups in the factors Sn ⊆ Sn such that OK (1) left-translates of 0 Sn cover Sn. By the theory of strong approximate groups, taking Hn the subgroups of non- 0 escaping elements, we get that Gn = hSni and Gn/Hn has dimension OK (1), a contradiction.  The next lemma is the main result in this subsection. Using this, we can pass the problem to connected Lie groups with bounded dimensions. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 32

1/2 1/2 Lemma 6.10 (Lie model from small expansions). If µG(AB) ≤ KµG (A)µG (B) and then there is a connected compact subgroup H of G such that G/H is a Lie group of dimension OK (1) and, with π : G → G/H the quotient map, πA and πB have O(1) 1/2 1/2 µG-measure O(K )µG (A)µG (B). Proof. By Fact 6.8, there is an open K-approximate group S, with

O(1) 1/2 1/2 µG(S) = O(K )µG (A)µG (B) such that A can be covered by O(KO(1)) right translation of S, and B can be covered by O(KO(1)) left translation of S. By Fact 6.9, there is a closed connected normal 4 subgroup H in S , such that G/H is a Lie group of dimension at most OK (1). Let π be the quotient map. Since H ⊆ S4, we have

5 O(1) 1/2 1/2 µG/H (π(S)) = µG(SH) ≤ µG(S ) = O(K )µG (A)µG (B). Note that π(A) can be covered by O(KO(1)) right translations of π(S), and π(B) can be covered by O(KO(1)) left translations of π(S). Hence, we get the desired conclusion.  The following proposition tells us that small measure expansion phenomenon can always be reduced to the same phenomenon on a Lie group with small dimension. Proposition 6.11 (Coarse versions of the main theorems). There is a constant τ such that if either G is noncompact and µG(A) = µG(B) or G is compact and µG(A) = µG(B) < τ, and

dG(A, B) < min{µG(A), µG(B)}, then there is a connected compact normal subgroup H of G, and σ-compact subsets A0,B0 of G/H satisfying: (i) G/H is a Lie group of dimension O(1); (ii) With π : G → G/H the quotient map,

−1 0 −1 0 µG(A 4 π A ) < 3dG(A, B) and µG(B 4 π B ) < 3dG(A, B);

0 0 (iii) dG/H (A ,B ) < 7dG(A, B).

1/2 1/2 Proof. From the assumption we have µG(AB) < 3µG (A)µG B. Obtain H as in Lemma 6.10, and when µG(A), µG(B) are small enough, we have µG/H (πA) + µG/H (πB) < 1. Then G/H is a Lie group of dimension O(1). By applying Theo- rem 6.5, we get the desired conclusion.  MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 33

6.3. Structure control on the nearly minimally expanding pairs. The fol- lowing useful lemma is a corollary of Theorem 6.5, which will be used at various points in the later proofs. It tells us the character given in the parallel Bohr sets is essentially unique.

Lemma 6.12 (Stability of characters). Suppose G is compact, χ : G → T is a continuous surjective group homomorphism, J ⊆ T is a compact interval, and η is a constant. Suppose we have

(i) dG(A, B) < min{µG(A), µG(B), cG/7} where cG is from Fact 3.11; −1 (ii) µT(J) = µG(B) and µG(B 4 χ (J)) = ηdG(A, B); (iii) µG(A) + µG(B) ≤ 1 − (2η + 20)dG(A, B) and (3η + 30)dG < µG(B).

Then there is a compact interval I in T such that µT(I) = µG(A) and −1 µG(A 4 χ (I)) ≤ 10dG(A, B). Proof. Let H = ker(χ). Note that χ is an open map, and hence by Fact 3.4 we have ∼ 0 0 0 0 G/H = T. By Theorem 6.5, there are sets A and B in T such that dT(A ,B ) ≤ 7dG(A, B), and −1 0 −1 0 µG(A 4 χ (A )) = 3dG(A, B), µG(B 4 χ (B )) = 3dG(A, B). 0 Then µT(B 4 J) ≤ (η + 3)dG(A, B). By Fact 3.11, there is a continuous surjective group homomorphism ρ : T → T and two compact intervals IA,IB ⊆ T, such that −1 0 −1 0 µT(ρ (IA) 4 A ) ≤ 7dG(A, B), µT(ρ (IB) 4 B ) ≤ 7dG(A, B), 0 0 0 and µT(A ) = µT(IA), µT(B ) = µT(IB). On the other hand, as µT(B 4 J) ≤ (η + −1 3)dG(A, B), thus µT(ρ (IB) 4 J) ≤ (η + 10)dG(A, B). Assume ρ is not the identity map, then by an elementary analysis, the fact that 1 − µT(I) > (2η + 20)dG(A, B), and the upper bound on η, we have that ρ = id. Set I = IA. Then, we have −1 −1 0 0 µG(A 4 χ (I)) ≤ µG(A 4 χ (A )) + µT(A 4 I) ≤ 10dG(A, B) which finishes the proof.  The next lemma shows that, if the symmetric difference of a set A and an interval is small, then A is also contained in an interval of bounded length. To handle arbitrary sets, we use inner measure here.

Lemma 6.13. Suppose G is compact, µeG is the inner measure associated to µG, and A, B ⊆ G has edG(A, B) < ε with edG(A, B) = µeG(A, B) − µeG(A) − µeG(B). Assume further that Ae ⊆ A and Be ⊆ B are σ-compact with µG(Ae) = µeG(A) and µG(Be) = µeG(B), χ : G → T is a continuous surjective group homomorphism, and MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 34

I,J compact intervals in T, with µT(I) = µG(A), µT(J) = µG(B), and −1 −1 µG(Ae4 χ (I)) < ε, µG(Be 4 χ (J)) < ε. Then there are intervals I0,J 0 ⊆ T, such that A ⊆ χ−1(I0), B ⊆ χ−1(J 0), and 0 0 µT(I ) − µeG(A) < 10ε, µT(J ) − µeG(B) < 10ε. Proof. We will show that for all g ∈ A \ χ−1(I), the distance between χ(g) and I in T is at most 5ε, and for all g0 ∈ B \ χ−1(J), the distance between g0 and J is at most 5ε. This implies there are intervals I0,J 0 in T such that A ⊆ χ−1(I0) and B ⊆ χ−1(J 0), and −1 0 −1 0 µG(χ (I ) \ A) < 10ε µT(χ (J ) \ B) < 10ε, as desired. Observe that AeBe is a σ-compact subset of AB, and so dG(A,e Be) < ε. By symmetry, it suffices to show the statement for g ∈ A \ χ−1(I). Suppose to the contrary that g is in A \ χ−1(I), and the distance between χ(g) and I in T is strictly greater than 5ε. By replacing Ae with Ae ∪ {g} if necessary, we can assume g ∈ Ae. We then have

µT(χ(g)χ(Be) \ Iχ(Be)) ≥ 5ε − µT(J \ χ(Be)) ≥ 4ε. −1 −1 and this implies that µG(gBe \ χ (I)χ (J)) ≥ 3ε. Therefore, −1 −1  −1 −1 µG(AeBe) ≥ µG (χ (I) ∩ Ae)(χ (J) ∩ Be) + µG(gBe \ χ (I)χ (J))

≥ µG(Ae) + µG(Be) − 2ε + 3ε, and this contradicts the fact that dG(A,e Be) < ε.  The stability lemma, together with Theorem 6.5, will be enough to derive a dif- ferent proof of a theorem by Tao [Tao18], with a sharp exponent bound. As we mentioned in the introduction, the same result with a sharp exponent bound was also obtained by Christ and Iliopoulou [CI21] recently, via a different approach. Theorem 6.14 (Theorem 1.2 for compact abelian groups). Let G be a connected compact abelian group, and A, B be compact subsets of G with positive measure. Set

λ = min{µG(A), µG(B), 1 − µG(A) − µG(B)}. Given 0 < ε < 1, there is a constant K = K(λ) does not depend on G, such that if δ < Kε and

µG(A + B) < µG(A) + µG(B) + δ min{µG(A), µG(B)}. Then there is a surjective continuous group homomorphism χ : G → T together with two compact intervals I,J ∈ T with

µT(I) − µG(A) < εµG(A), µT(J) − µG(B) < εµG(B), MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 35 and A ⊆ χ−1(I), B ⊆ χ−1(J).

Proof. We first assume that dG(A, B) is sufficiently small, and we will compute the bound on dG(A, B) later. As G is abelian, by Proposition 6.11, there is a quotient map π : G → Td, and A0,B0 ⊆ Td, such that −1 0 −1 0 µG(A 4 π A ) < 3dG(A, B) and µG(B 4 π B ) < 3dG(A, B) 0 0 and dG/H (A ,B ) < 7dG(A, B). Let c = c(τ) be as in Fact 3.10, and by Lemma 5.3, 00 00 d 00 there is a constant L depending only on λ and c, and sets A ,B ⊆ T with µTd (A ) = 00 µTd (B ) = c such that 00 0 00 00 0 00 max{dG(A ,B ), dG(A ,B ), dG(A ,B )} < LdG(A, B). 0 0 0 0 By Fact 3.10, there are intervals I ,J ⊆ T with µT(I ) = µT(J ) = c, and a continuous surjective group homomorphism ρ : Td → T, such that 00 −1 0 00 −1 0 µTd (A 4 ρ (I )) < LdG(A, B) and µTd (B 4 ρ (J )) < LdG(A, B). By Lemma 6.12, there are intervals I0,J 0 ⊆ T with 0 −1 0 0 −1 0 µTd (A 4 ρ (I )) < 10LdG(A, B) and µTd (B 4 ρ (J )) < 10LdG(A, B). Let χ = π ◦ ρ. Hence, we have −1 0 −1 0 µG(A 4 χ (I )) < (3 + 10L)dG(A, B) and µG(B 4 χ (J )) < (3 + 10L)dG(A, B). Using Lemma 6.13, there are intervals I,J ⊆ T, such that A ⊆ χ−1(I), B ⊆ χ−1(J), and

µT(I) − µG(A) < (30 + 100L)dG(A, B),

µT(J) − µG(B) < (30 + 100L)dG(A, B). Now, we fix n 1 c o K := min , , 30 + 100L L and δ < Kε, where dG(A, B) = δ min{µG(A), µG(B)}. Clearly, we will have

µT(I) − µG(A) < ε min{µG(A), µG(B)},

µT(J) − µG(B) < ε min{µG(A), µG(B)}. Note that in the above argument, we apply Fact 3.10 on A00,B00, and this would require that LdG(A, B) < c. By the way we choose K, we have

LdG(A, B) = Lδ min{µG(A), µG(B)} < c, as desired.  The following theorem shows that, once we have a certain group homomorphism to tori, we will get a good structural control on the (nearly) minimal expansion sets. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 36

Proposition 6.15 (Toric domination from a given homomorphism). Suppose A, B have dG(A, B) < min{µG(A), µG(B)}, and χ : G → T is a continuous surjective group homomorphism such that µT(χ(A)) + µT(χ(B)) < 1/5. Then there is a contin- uous and surjective group homomorphism ρ : G → T, a constant K0 only depending on min{µG(A), µG(B)}, and compact intervals I,J ⊆ T with µT(I) = µG(A) and µT(J) = µG(B), such that

−1 −1 µG(A 4 ρ (I)) < K0dG(A, B), and µG(B 4 ρ (J)) < K0dG(A, B).

Proof. By Theorem 6.5, there are A0,B0 ⊆ T, such that

−1 0 −1 0 (6) µG(A 4 χ (A )) < 3dG(A, B) and µG(B 4 χ (B )) < 3dG(A, B),

0 0 and dT(A ,B ) < 7dG(A, B). By Theorem 6.14, there are continuous surjective group homomorphism η : T → T, a constant L depending only on min{µG(A), µG(B)}, and 0 0 compact intervals I,J ⊆ T such that µT(I) = µT(A ), µT(J) = µT(B ), and

0 −1 0 −1 (7) µT(A 4 η (I)) < LdG(A, B) and µT(B 4 η (J)) < LdG(A, B).

Set ρ = η ◦ χ. The conclusion follows from (6) and (7) with K0 = L + 3. 

In light of Proposition 6.15, in the rest of the paper, we will be focusing on finding the desired group homomorphism mapping G to tori.

7. Pseudometrics and group homomorphisms onto tori Proposition 6.11 and Proposition 6.15 reduce the proof of Theorem 1.1 and Theo- rem 1.2 to problems of constructing certain group homomorphisms from a Lie group onto T or R. In this section, we show these problems can be reduced further to prob- lems of constructing pseudometrics with certain properties on the ambient group. Section 7.1 shows that a linear pseudometric suffices, and Section 7.2 and Section 7.3 does so when the pseudometric is almost linear and almost monotone. Throughout, G is a connected and unimodular Lie group with Haar measure µG. Recall that a pseudometric on a set X is a function d : X × X → R satisfying the following three properties: (1) (Reflexive) d(a, a) = 0 for all a ∈ X, (2) (Symmetry) d(a, b) = d(b, a) for all a, b ∈ X, (3) (Triangle inequality) d(a, c) ≤ d(a, b) + d(b, c) ∈ X. Hence, a pseudometric on X is a metric if for all a, b ∈ X, we have d(a, b) = 0 implies a = b. If d is a pseudometric on G, for an element g ∈ G, we set kgkd = d(idG, g). MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 37

7.1. Linear pseudometrics. Suppose d is a pseudometric on G. We say that d is left-invariant if for all g, g1, g2 ∈ G, we have d(gg1, gg2) = d(g1, g2). left-invariant pseudometrics arise naturally from measurable sets in a group; the pseudometric we will construct in Section 8 is of this form.

Proposition 7.1. Suppose A is a measurable subset of G. For g1 and g2 in G, define

d(g1, g2) = µG(A) − µG(g1A ∩ g2A). Then d is a continuous left-invariant pseudometric on G.

Proof. We first verify the triangle inequality. Let g1, g2, and g3 be in G, we need to show that

(8) µG(A) − µG(g1A ∩ g3A) ≤ µG(A) − µG(g1A ∩ g2A) + µG(A) − µG(g2A ∩ g3A).

As µG(A) = µG(g2A), we have µG(A)−µG(g1A∩g2A) = µG(g2A\g1A), and µG(A)− µG(g2A ∩ g3A) = µG(g2A \ g3A). Hence, (8) is equivalent to

µG(g2A) − µG(g2A \ g1A) − µG(g2A \ g3A) ≤ µ(g1A ∩ g3A).

Note that the left-hand side is at most µG(g1A ∩ g2A ∩ g3A), which is less than the right-hand side. Hence, we get the desired conclusion. The continuity of d follows from Fact 3.1(vii), and the remaining parts are straightforward.  Another natural source of left-invariant pseudometrics is group homomorphims onto metric groups. Suppose de is a continuous left-invariant metric on a group H and π : G → H is a group homomorphism, then for every g1, g2 in G, one can naturally define a pseudometric d(g1, g2) = de(π(g1), π(g2)). It is easy to see that such d is a continuous left-invariant pseudometric, and {g ∈ G : kgkd = 0} = ker(π) is a normal subgroup of G. The latter part of this statement is no longer true for an arbitrary continuous left-invariant pseudometric, but we still have the following: Lemma 7.2. Suppose d is a continuous left-invariant pseudometric on G. Then the set {g ∈ G : kgkd = 0} is the underlying set of a closed subgroup of G.

Proof. Suppose g1 and g2 are elements in G such that kg1kd = kg2kd = 0. Then

d(idG, g1g2) ≤ d(idG, g1) + d(g1, g1g2) = d(idG, g1) + d(idG, g2) = 0.

Now, suppose (gn) is a sequence of elements in G converging to g with kgnkd = 0. Then kgkd = 0 by continuity, we get the desired conclusions.  In many situations, a left-invariant pseudometric allows us to construct surjective continuous group homomorphism to metric groups. The following lemma tells us precisely when this happens. We omit the proof as the result is motivationally relevant but will not be used later on. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 38

Lemma 7.3. Let d be a continuous left-invariant pseudometric on G. The following are equivalent,

(i) The set {g ∈ G : kgkd = 0} is the underlying set of a closed normal subgroup of G. (ii) There is a continuous surjective group homomorphism π : G → H, and de is a left-invariant metric on H. Then

d(g1, g2) = de(πg1, πg2).

Moreover, when (ii) happens, {g ∈ G : kgkd = 0} = ker π, hence H and de if exist are uniquely determined up to isomorphism.

The group R and T = R/Z are naturally equipped with the metrics dR and dT induced by the Euclidean norms, and these metrics interact in a very special way with the additive structures. Hence one would expect that if there is a group homo- morphism from G to either R or T, then G can be equipped with a pseudometric which interacts nontrivially with addition. Let d be a left-invariant pseudometric on G. The radius ρ of d is defined to be sup{kgkd : g ∈ G}; this is also sup{d(g1, g2): g1, g2 ∈ G} by left invariance. We say that d is locally linear if it satisfies the following properties: (1) d is continuous and left-invariant; (2) for all g1, g2, and g3 with d(g1, g2) + d(g2, g3) < ρ, we have either

(9) d(g1, g3) = d(g1, g2) + d(g2, g3), or d(g1, g3) = |d(g1, g2) − d(g2, g3)|.

A pseudometric d is monotone if for all g ∈ G such that kgkd < ρ/2, we have 2 kg kd = 2kgkd. To investigate the property of this notion further, we need the following fact about the adjoint representations of Lie groups [HN12, Proposition 9.2.21]. Fact 7.4. Let g be the Lie algebra of G, and let Ad : G → Aut(g) be the adjoint representation. Then ker(Ad) is the center of G. The following result is the first time we need G to be a Lie group instead of just a locally compact group. Proposition 7.5. If d is a locally linear pseudometric on G, then d is monotone. Proof. We first prove an auxiliary statement. Claim. Suppose s : G → G, g 7→ g2 is the squaring map. Then there is no open U ⊆ G and proper closed subgroup H of G such that s(U) ⊆ H. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 39

Proof of Claim. Consider the case where G is a connected component of a linear algebraic subgroup of GLn(R). Let Js be the Jacobian of the function s. Then the set {g ∈ G : det Js(g) = 0} has the form G ∩ Z where Z is a solution set of a system of polynomial equations. It is not possible to have G∩Z = G, as s is a local diffeomorphism at idG. Hence, G∩Z must be of strictly lower dimension than G. By the inverse function theorem, s|G\Z is open. Hence s(U) is not contained in a subgroup of G with smaller dimension. We also note a stronger conclusion for abelian Lie group: If V is an open subset of a not necessarily connected abelian Lie group A, then the image of A under a 7→ a2 is not contained in a closed subset of A with smaller dimension. Indeed, A is isomorphic as a topological group to D × Tm × Rm, with D a . If m m U ⊆ D × T × R , then it is easy to see that {a2 : a ∈ V } contains a subset of D × Tm × Rm, and is therefore not a subset of a closed subset of A with smaller dimension. Finally, we consider the general case. Suppose to the contrary that s(U) ⊆ H with H a proper closed subgroup of G. Let Z(G) be the center of G, G0 = G/Z(G), π : G → G0 be the quotient map, U 0 = π(U), and s0 : G0 → G0, g0 7→ (g0)2. Then U 0 is an open subset of G0, which is isomorphic as a topological group to a connected component of an algebraic group by Fact 7.4. By the earlier case, s0(U 0) is not contained in any proper closed subgroup of G0, so we must have π(H) = G0. In particular, this implies dim(H ∩ Z(G)) < dim Z(G), and HZ(G) = G. Choose h ∈ H such that hZ(G) ∩ U is nonempty. Then s(hZ(G) ∩ U) = {h2a2 : a ∈ Z(G) ∩ h−1U}. As s(hZ(G)∩U) ⊆ H, we must have {a2 : a ∈ Z(G)∩h−1U} is a subset of H ∩Z(G). Using the case for abelian Lie groups, this is a contradiction, because H ∩ Z(G) is a closed subset of Z(G) with smaller dimension. We now get back to the problem of showing that d is monotone. As d is invariant,1 2 d(idG, g) = d(g, g ) for all g ∈ G. From local linearity of d, for all g ∈ G with kgkd < ρ/2, we either have 2 2 kg kd = 2kgkd or kg kd = 0. 2 It suffices to rule out the possibility that 0 < kgkd < ρ/4, and kg kd = 0. As d is continuous, there is an open neighborhood W of g such that for all g0 ∈ W , 0 0 2 we have kg kd > 0 and k(g ) kd = 0. From Lemma 7.2, the set {g ∈ G : kgkd = 0} is a closed subgroup of G. As d is nontrivial and G is a connected Lie group, MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 40

{g ∈ G : kgkd = 0} must be a Lie group with smaller dimension. Therefore, we only need to show that if W is an open subset of G, then s(W ) is not contained in a closed subgroup of G with smaller dimension, where s : G → G is the squaring map, and this is guaranteed by the earlier claim.  The next result confirms our earlier intuition: locally linear pseudometric in G will induce a homomorphism mapping to either T or R. Proposition 7.6. Suppose d is a locally linear pseudometric with radius ρ > 0. Then ker d is a normal subgroup of G, G/ ker d is isomorphic to T if G is compact, and G/ ker d is isomorphic to R if G is noncompact.

Proof. We first prove that ker d is a normal subgroup of G. Suppose kgkd = 0 and h ∈ G satisfies khkd < ρ/4. We have −1 −1 d(h, hgh ) = d(idG, gh ) −1 −1 = |d(idG, g) ± d(g, gh )| = d(idG, h ) = d(idG, h).

−1 −1 Hence, d(idG, hgh ) = |d(idG, h) ± d(h, hgh )| is either 0 or 2d(idG, h). Assume −1 first that khgh kd = 0 for every such h when kgkd = 0. Let

U := {h : khkd < ρ/4}. By the continuity of d, U is open. Hence for every h in G, h can be written as a finite product of elements in U. By induction, we conclude that for every h ∈ G, −1 khgh kd = 0 given kgkd = 0, and this implies that ker d is normal in G. −1 Suppose khgh kd = 2khkd. By Proposition 7.5, d is monotone. Hence, we have 2 −1 khg h kd = 4khkd. 2 −1 On the other hand, as kgkd = 0, repeating the argument above, we get khg h kd is −1 either 0 or 2khkd. Hence, khkd = 0, and so khgh kd = 0. We now show that G0 = G/ ker d has dimension 1. Let d0 be the pseudometric 0 0 on G induced by d. Choose g ∈ G in the neighborhood of idG0 such that g is in 0 the image of the exponential map and kgkd0 < ρ/4. If g is another element in the 0 neighborhood of idG0 which is in the image of the exponential map and kg kd0 < ρ/4. 0 0 Without loss of generality, we may assume kg kd0 ≤ kgkd0 . Suppose g = exp(X). 0 k Then, by monotonicity, there is k ≥ 1 such that k(g ) kd0 ≥ kgkd0 . By the continuity of the exponential map, there is t ∈ (0, 1] such that

kgkd0 = kexp(tkX)kd0 . This implies that g and g0 are on the same one parameter subgroup, which is the desired conclusion.  MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 41

7.2. Almost linear pseudometrics: relative sign and total weight functions. In this section, we will introduce a weakening of the notion of a locally linear pseudo- metric and define the relative sign function and total weight function associate to it. When d is a pseudometric arising from a measurable subset A as in Proposition 7.1, these roughly give the “direction” and the “distance” that an element of the group translates A. Throughout this section, d is a pseudometric on G with radius ρ > 0, and γ is a constant with 0 < γ < 10−8ρ. For a constant λ, we write I(λ) for the interval (−λ, λ) in either R or T, and we write N(λ) for {g ∈ G : kgkd ∈ I(λ)}. By Fact 3.1(vii), N(λ) is an open set, and hence measurable. We say that d is γ-linear if it satisfies the following conditions: (1) d is continuous and left-invariant; (2) for all g1, g2, g3 ∈ G with d(g1, g2) + d(g2, g3) < ρ − γ, we have either

d(g1, g3) ∈ d(g1, g2) + d(g2, g3) + I(γ), or d(g1, g3) ∈ |d(g1, g2) − d(g2, g3)| + I(γ).

Given α ≤ ρ, let N(α) = {g ∈ G : kgkd ≤ α}. We say that d is γ-monotone if for all g ∈ N(ρ/2 − γ), we have 2 kg kd ∈ 2kgkd + I(γ). The next lemma says that under the γ-linearity condition, the group G essentially has only one “direction”: if there are three elements have the same distance to idG, then at least two of them are very close to each other.

Lemma 7.7. Suppose d is a γ-linear pseudometric on G. If g, g1, g2 ∈ G such that

kgkd = kg1kd = kg2kd ∈ I(ρ/4 − γ) \ I(2γ), and d(g1, g2) ∈ 2kgkd + I(γ). Then either d(g, g1) ∈ I(γ) or d(g, g2) ∈ I(γ).

Proof. Suppose both d(g, g1) and d(g, g2) are not in I(γ). By γ-linearity of d, we have d(g, g1) ∈ |d(idG, g) ± d(idG, g1)| + I(γ), and so d(g, g1) ∈ 2kgkd + I(γ). Similarly, we have d(g, g2)2kgkd + I(γ). Suppose first that d(g1, g2) ∈ d(g, g1) + d(g, g2) + I(γ), then

d(g1, g2) ∈ 4kgkd + I(3γ).

On the other hand, by γ-linearity we have d(g1, g2) ≤ 2kgkd + γ. Hence, we have kgkd ∈ I(2γ), a contradiction. The other two possibilities are d(g1, g2) + d(g, g2) ∈ d(g, g1) + I(γ) or d(g1, g2) + d(g, g1) ∈ d(g, g2) + I(γ), but similar calculations also lead to contradictions.  MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 42

Proposition 7.8 below is a partial replacement for Proposition 7.5 for linear pseu- dometric. The fact that we do not automatically have monotonicity is a reason that the later Section 8.3 is much harder than Section 8.2. Proposition 7.8 (Path monotonicity implies global monotonicity). Let g be the Lie algebra of G, exp : g → G the exponential map, and d a γ-linear pseudometric on G. Suppose for each X in g, we have one of the following two possibilities:

(i) kexp(tX)kd < γ for all t ∈ R; >0 (ii) there is t0 ∈ R with kexp(t0X)kd ∈ I(ρ/2 − γ) \ I(ρ/4),

(10) kexp(2t0X)kd = 2kexp(t0X)kd + I(γ), and

(11) kexp(tX)kd + kexp((t0 − t)X)kd ∈ kexp(t0X)kd + I(γ)

for all t ∈ [0, t0]. Then d is (9γ)-monotone.

Proof. Fix an element g of G with kgkd ∈ I(ρ/2 − 16γ). Our job is to show that 2 kg kd ∈ 2kgkd + I(9γ). Since G is compact and connected, the exponential map exp is surjective. We get X ∈ g such that g ∈ {exp(tX): t ∈ R}. If we are in scenario 2 (i), then kgkd < γ, hence kg kd ∈ 2kgkd + I(3γ). Therefore, it remains to deal with the case where we have an t0 as in (ii). Set g0 = exp(t0X). We consider first the special case where kgkd < kg0kd − 2γ. As d is continuous, there is t1 ∈ [0, t0] such that with g1 = exp(t1X), we have −1 kg1kd = kgkd. Let t2 = −t1, and g2 = exp(t2X) = g1 . Since d is invariant, −1 kg2kd = d(g1 , idG) = d(idG, g1) = kg1kd. 2 Hence, kg1kd = kg2kd = kgkd. If kgkd < 2γ, then kg kd ∈ 2kgkd + I(5γ) and we are done. Thus we suppose kgkd ≥ 2γ. Then, by Lemma 7.7, either d(g, g1) < γ, or d(g, g2) < γ. 2 Since these two cases are similar, we assume that d(g, g1) < γ. By γ-linearity, kg1kd 2 is in either 2kg1kd + I(γ) or I(γ). Using kg0kd ∈ 2kg0kd + I(γ) and the assumption that kgkd < kg0kd − 2γ, in either case, we have 2 2 (12) kg1kd < kg0kd − 2γ. −1 −1 Since g0 g1 = g1g0 , and by γ-linearity of d, we get 2 2 −2 2 −1 2 (13) d(g1, g0) = d(idG, g1 g0) = d(idG, (g1 g0) ) ∈ {0, 2d(g1, g0)} + I(γ).

By (11), we have kg1kd + d(g1, g0) ∈ kg0kd + I(γ). Recalling that kg1kd = kgkd > 2γ, and from (10) and (13), we have 2 2 2 (14) d(g1, g0) < 2kg0kd − 3γ = kg0k − 2γ. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 43

By (12), (14), and the γ-linearity of d, we have 2 2 2 2 kg1kd ∈ kg0kd − d(g1, g0) + I(γ). Therefore by (11) and (13), we have either 2 2 kg1kd ∈ 2kg1kd + I(5γ) or kg1kd ∈ 2kg0kd + I(3γ). 2 2 As kg1kd ≤ 2kg1k + γ < 2kg0k − 5γ, we must have kg1k ∈ 2kg1k + I(5γ). Now, since −1 kg1 gkd = d(g1, g) < γ, again by the γ-linearity we conclude that 2 2 −1 2 d(g1, g ) = k(g1 g) kd < 3γ. 2 Thus, kg kd ∈ 2kgkd + I(9γ). Finally, we consider the other special case where kg0kd + 2γ < kgkd < ρ/2 − 16γ. 2 For g1 = exp(t1X) with t1 ∈ [0, t0], we have kg1kd ∈ 2kg1k + I(8γ) by a similar 2 argument as above. Using continuity, we can choose t1 such that kg1kd = kgkd, and −1 let g2 = g1 . The argument goes in exactly the same way with the role of g1 replaced 2 2 by g1 and the role of g2 replaced by g2. 

Suppose d is γ-linear. We define s(g1, g2) to be the relative sign for g1, g2 ∈ G satisfying kg1kd + kg2kd < ρ − γ by  0 if min{kg k , kg k } ≤ 4γ,  1 d 2 d s(g1, g2) = 1 if min{kg1kd, kg2kd} > 4γ and kg1g2kd ∈ kg1kd + kg2kd + I(γ).  −1 if min{kg1kd, kg2kd} > 4γ and |g1g2|d ∈ kg1kd − kg2kd + I(γ). Note that this is well-defined because when min{kg k , kg k } ≥ 4γ in the above 1 d 2 d definition, the differences between kg1kd − kg2kd and kg1kd + kg2kd is at least 6γ. The following lemma gives us tools to relate signs between different elements.

Proposition 7.9. Suppose d is γ-linear and γ-monotone. Then for g1, g2, and g3 in N(ρ/4 − γ) \ N(4γ), we have the following −1 (i) s(g1, g1 ) = −1 and s(g1, g1) = 1. (ii) s(g1, g2) = s(g2, g1). −1 −1 −1 −1 (iii) s(g1, g2) = s(g1 , g2 ) = −s(g1 , g2) = −s(g1, g2 ). (iv) s(g1, g2)s(g2, g3)s(g3, g1) = 1. (v) If kg1kd ≤ kg2kd, and g1g2 is in N(ρ/4 − γ) \ N(4γ), then

s(g0, g1g2) = s(g0, g2g1) = s(g0, g2).

Proof. As g1, g2, and g3 are in N(ρ/4 − γ) \ N(4γ), one has s(gi, gj) 6= 0 for all i, j ∈ {1, 2, 3}. The first part of (i) is immediate from the fact that kidGkd = 0, and the second part of (i) follows from the γ-monotonicity and the definition of the relative sign. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 44

We now prove (ii). Suppose to the contrary that s(g1, g2) = −s(g2, g1). Without loss of generality, assume s(g1, g2) = 1. Then kg1g2g1g2kd is in 2kg1g2kd +I(γ), which is a subset of 2kg1kd + 2kg2kd + I(3γ). On the other hand, as s(g2, g1) = −1, we have

kg1g2g1g2kd ∈ kg1kd ± (kg2kd − kg1kd) ± kg2kd + I(3γ).

This contradicts the assumption that g1 and g2 are not in N(4γ). −1 Next, we prove the first and third equality in (iii). Note that kgkd = kg kd for all −1 −1 g ∈ G as d is symmetric and invariant. Hence, kg1g2kd = kg2 g1 kd. This implies −1 −1 that s(g1, g2) = s(g2 , g1 ). Combining with (ii), we get the first equality in (iii). The third equality in (iii) is a consequence of the first equality in (iii). −1 −1 −1 Now, consider the second equality in (iii). Suppose s(g1 , g2 ) = s(g1 , g2). Then, −1 from (ii) and the first equality of (iii), we get s(g2, g1) = s(g1 , g2). Hence, either −1 kg2g1g1 g2kd ∈ 2 (kg1kd + kg2kd) + I(3γ) or −1 kg2g1g1 g2kd ∈ 2 kg1kd − kg2kd + I(3γ). −1 2 On the other hand, kg2g1g1 g2kd = kg2kd, which is in 2kg2kd + I(γ). We get a contradiction with the fact that g1 and g2 are not in N(4γ). We now prove (iv). Without loss of generality, assume kg1kd ≤ kg2kd ≤ kg3kd. −1 Using (iii) to replace g3 with g3 if necessary, we can further assume that s(g2, g3) = 1. We need to show that s(g1, g2) = s(g1, g3). Suppose to the contrary. Then, from (iii), −1 −1 we get s(g1, g2) = s(g1 , g3). Using (iii) to replacing g1 with g1 if necessary, we can −1 assume that s(g1, g2) = s(g1 , g3) = 1. Using (ii), we get s(g2, g1) = 1. Hence, either −1 kg2g1g1 g3kd ∈ 2kg1kd + kg2kd + kg3kd + I(3γ) or −1 kg2g1g1 g3kd ∈ kg3kd − kg2kd + I(3γ). −1 On the other hand, kg2g1g1 g3kd = kg2g3kd is in kg2kd + kg3kd + I(γ). Hence, we get a contradiction to the fact that g1, g2, and g3 are not in N(4γ). Finally, we prove (v). Using (iv), it suffices to show s(g1g2, g2) = s(g2g1, g2) = 1. We will only show the former, as the proof for the latter is similar. Suppose to the 2 contrary that s(g1g2, g2) = −1. Then kg1g2kd is in kg1g2kd − kg2kd + I(γ), which is 2 2 a subset of kg1kd +I(2γ). On the other hand, kg1g2kd is also in kg1kd −kg2kd +I(γ) which is a subset of 2kg2kd − kg1kd + I(2γ). Hence, we get a contradiction with the assumption that g1 and g2 are not in N(4γ).  The notion of relative sign corrects the ambiguity in calculating distance, as can be seen in the next result.

Lemma 7.10. Suppose d is γ-monotone γ-linear, and g1 and g2 are in N(ρ/16 − γ) with kg1kd ≤ kg2kd. Then we have the following MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 45

(i) Both kg1g2kd and kg2g1kd are in s(g1, g2)kg1kd + kg2kd + I(5γ). (ii) If g0 is in N(ρ/4) \ N(4γ), then both s(g0, g1g2)kg1g2kd and s(g0, g2g1)kg2g1kd are in

s(g0, g1)kg1kd + s(g0, g2)kg2kd + I(25γ).

Proof. We first prove (i). When g1, g2 ∈/ N(4γ), the statement for kg1g2kd is im- mediate from the definition of the relative sign, and the statement for kg2g1kd is a consequence of Proposition 7.9(ii). Now suppose kg1kd < 4γ. From the γ-linearity, we have

kg2kd − kg1kd − γ < kg1g2kd < kg1kd + kg2kd + γ.

We deal with the case where kg2kd < 4γ similarly. We now prove (ii). Fix g0 in N(ρ/4 − γ) \ N(4γ). We will consider two cases, when g1 is not in N(4γ) and when g1 is in N(4γ). Suppose we are in the first case, that is g1 ∈/ N(4γ). As kg1kd ≤ kg2kd, we also have g2 ∈/ N(4γ). If both g1g2 and g2g1 are not in N(4γ), then the desired conclusion is a consequence of (i) and Proposition 7.9(iv, v). Within the first case, it remains to deal with the situations where g1g2 is in N(4γ) or g2g1 is in N(4γ). Since these two situations are similar, we may assume g1g2 is in N(4γ). From (i), we have s(g1, g2) = −1 and kg2kd − kg1kd is at most 5γ. Therefore, kg2g1kd is in I(6γ). By Proposition 7.9(iv), we have s(g0, g1) = −s(g0, g2), and so

s(g0, g1)kg1kd + s(g0, g2)kg2kd ∈ I(6γ).

Since both s(g0, g1g2)kg1g2kd and s(g0, g2g1)kg1g2kd are in I(6γ), they are both in s(g0, g1)kg1kd + s(g0, g2)kg2kd + I(12γ) giving us the desired conclusion. Continuing from the previous paragraph, we consider the second case when g1 is in N(4γ). If g2 is in N(16γ), then both kg1g2kd and kg2g1kd are in I(25γ) by (i), and the desired conclusion follows. Now suppose g2 is not in N(16γ). Then from (i) and the fact that g1 ∈ N(4γ), we get g1g2 and g2g1 are both not in N(4γ). Note that −1 s(g1g2, g2 ) = −1, because otherwise we get −1 kg1kd ≥ kg1g2kd + kg2 kd − 5γ > 4γ. −1 A similar argument gives s(g2 , g2g1) = −1. Hence, s(g1g2, g2) = s(g2g1, g2) = 1. By Proposition 7.9(v), we get

s(g0, g2) = s(g0, g1g2) = s(g0, g2g1).

From (i), kg1g2kd and kg2g1kd are both in kg2kd + I(9γ). On the other hand, as s(g0, g1) = 0, we have s(g0, g1)kg1kd + s(g0, g2)kg2kd = s(g0, g2)kg2kd. The desired conclusion follows.  The next corollary will be important in the subsequent development. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 46

0 Corollary 7.11. Suppose d is γ-linear and γ-monotone, g0 and g0 are elements in N(ρ/4 − γ) \ N(4γ), and (g1, . . . , gn) is a sequence with gi ∈ N(ρ/4 − γ) \ N(4γ) for i ∈ {1, . . . , n}. Then

n n X X 0 s(g0, gi)kgikd = s(g0, gi)kgikd . i=1 i=1 0 Proof. As s(g0, gi) = s(g0, gi) = 0 whenever kgikd < 4γ, we can reduce to the case −1 where min1≤i≤n kgikd ≥ 4γ. Using Proposition 7.9(iii) to replace g0 with g0 if 0 necessary, we can assume that s(g0, g1) = s(g0, g1). Then by Proposition 7.9(iii), 0 s(g0, gi) = s(g0, gi) for all i ∈ {1, . . . , n}. This gives us the desired conclusion. 

The following auxiliary lemma allows us to choose g0 as in Corollary 7.11. Lemma 7.12. The set N(ρ/4 − γ) \ N(4γ) is not empty.

Proof. It suffices to show that µG(N(4γ)) < µG(N(ρ/4 − γ). Since idG is in N(4γ), 2 N(4γ) is a nonempty open set and has µG(N(4γ)) > 0. Therefore, N (4γ) and N 4(4γ) are also open. By γ-linearity, we have 2 4 N (4γ) ⊆ N9γ and N (4γ) ⊆ N19γ. As 19γ < ρ, we have N 4(4γ) 6= G. Using Corollary 4.3, we get 2 µG(N (4γ)) ≤ 2/3 and µG(N(4γ)) < 1/3. 2 Hence, by Kemperman’s inequality µG(N(4γ)) < µG(N (4γ)) ≤ µG(N(ρ/4 − γ)), which is the desired conclusion. 

Suppose (g1, . . . , gn) is a sequence of elements in N(ρ/4 − γ) \ N(4γ). We set

n X t(g1, . . . , gn) = s(g0, gi)kgikd i=1 with g0 is an arbitrary element in N(ρ/4−γ)\N(4γ), and call this the total weight associated to (g1, . . . , gn). This is well-defined by Corollary 7.11 and Lemma 7.12. 7.3. Almost linear pseudometrics: group homomorphisms onto tori. In this section, we will use the relative sign function and the total weight function defined in Section 7.2 to define a universally measurable multivalued group homomorphism onto T. We will then use a number or results in descriptive set theory and geometry to refine this into a continuous group homomorphism. We keep the setting of Section 7.2, and assume further that G is compact. Let s and t be the relative sign function and the total weight function defined earlier. Set λ = ρ/36, and N[λ] = {g ∈ G : kgkd ≤ λ}. The set N[λ] is compact, and hence MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 47 measurable. Moreover, Lemma 7.10 is applicable when g0 is an arbitrary element in N(ρ/4 − γ) \ N(4γ), and g1 are g2 are in N[λ]. A sequence (g1, . . . , gn) of elements in G is a λ-sequence if gi is in N[λ] for all i ∈ {1, . . . , n}. We are interested in expressing an arbitrary g of G as a product of a λ-sequence where all components are “in the same direction”. The following notion captures that idea. A λ-sequence (g1, . . . , gn) is irreducible if for all 2 ≤ j ≤ 4, we have gi+1 ··· gi+j ∈/ N(λ).

A concatenation of a λ-sequence (g1, . . . , gn) is a λ-sequence (h1, . . . , hm) such that there are 0 = k0 < k1 < ··· < km = n with

hi = gki−1+1 ··· gki for i ∈ {1, . . . , m}. The next lemma allows us to reduce an arbitrary sequence to irreducible λ-sequences via concatenation.

Lemma 7.13. Suppose d is γ-linear and γ-monotone, and (g1, . . . , gn) is a λ-sequence. 0 0 Then (g1, . . . , gn) has an irreducible concatenation (g1, . . . , gm) with 0 0 t(g1, . . . , gm) ∈ t(g1, . . . , gn) + I(25(n − m)γ). Proof. The statement is immediate when n = 1. Using induction, suppose we have proven the statement for all smaller values of n. If (g1, . . . , gn) is irreducible, we are done. Consider the case where gi+1gi+2 is in N(λ) for some 0 ≤ i ≤ n − 2. Fix g0 in N(λ/4 − γ) \ N(4γ). Using Lemma 7.10(ii)

s(g0, gi+1gi+2)kgi+1gi+2kd ∈ s(g0, gi+1)kgi+1kd + s(g0, gi+2)kgi+2kd + I(25γ).

From here, we get the desired conclusion. The cases where either gi+1gi+2gi+3 for some 0 ≤ i ≤ n − 3 or gi+1gi+2gi+3gi+4 is in N(λ) for some 0 ≤ i ≤ n − 4 can be dealt with similarly.  The following lemma makes the earlier intuition of “in the same direction” precise:

Lemma 7.14. Suppose d is γ-linear and γ-monotone, g0 is in N(ρ/4 − γ) \ N(4γ), 0 0 and (g1, . . . , gn) is an irreducible λ-sequence. Then for all i, i , j, and j such that 2 ≤ j, j0 ≤ 4, 0 ≤ i ≤ n − j, and 0 ≤ i0 ≤ n − j0, we have

s(g0, gi+1 ··· gi+j) = s(g0, gi0+1 ··· gi0+j0 ). Proof. It suffices to show for fixed i, j with 0 ≤ i ≤ n − j − 1 and 2 ≤ j ≤ 3 that

s(g0, gi+1 ··· gi+j) = s(g0, gi+1 ··· gi+j+1).

Note that both gi+1 ··· gi+j and gi+1 ··· gi+j+1 are in N(ρ/4 − γ) \ N(4γ). Hence, applying Proposition 7.9(iv), we reduce the problem to showing −1 −1 s(gi+j ··· gi+1, gi+1 ··· gi+j+1)) = −1. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 48

This is the case because otherwise, kgi+j+1kd ≥ 2λ − γ > λ, a contradiction.  We now get a lower bound for the total distance of an irreducible λ-sequence:

Corollary 7.15. Suppose d is γ-linear and γ-monotone, and (g1, . . . , gn) is an irre- ducible λ-sequence. Then

t(g1, . . . , gn) > nλ/4.

Proof. If n = 2k, let hi = g2i−1g2i for i ∈ {1, . . . , k}. If n = 2k + 1, let hi = g2i−1g2i for i ∈ {1, . . . , k − 1}, and hk = g2n−1g2ng2n+1. From Lemma 7.10, we have

(15) t(h1, . . . , hk) ∈ t(g1, . . . , gn) + I(25(n − k)γ).

As (g1, . . . , gn) is irreducible, hi is in N(3λ)\N(λ) for i ∈ {1, . . . , k}. By Lemma 7.14, we get s(g0, hi) = s(g0, hj) for all i and j in i ∈ {1, . . . , k}. Thus, by the definition of the total weight again, t(h1, . . . , hk) > nλ/3. Combining with the assumption on λ and (15), we get t(g1, . . . , gn) > nλ/3 − 11nγ > nλ/4. 

When (g1. . . . , gn) is an irreducible λ-sequence, g1 ··· gm is intuitively closer to g0 than g1 ··· gm+k for some positive k. However, as G is compact, the sequence may “return back” to idG when n is large. The next proposition provides a lower bound estimate on such n.

Lemma 7.16 (Monitor lemma). Suppose d is γ-linear and γ-monotone, and (g1, . . . , gn) is an irreducible λ-sequence with g1 ··· gn = idG. Then n ≥ 1/µG(N(4λ)).

Proof. Let m > 0. For convenience, when m > n we write gm to denote the element gi with i ≤ n and i ≡ m (mod n). Define (m) N (4λ) = {g ∈ G | d(g, g1 ··· gm) < 4λ}. Note that we have N (m)(4λ) = N (m0)(4λ) when m ≡ m0 (mod n). By invariance of (m) (0) d and µG, clearly µG(N (4λ)) = µG(N(4λ)) for all m. We also write N (4λ) = N(4λ). We will show that

n−1 [ [ G = N (m)(4λ) = N (m)(4λ), m∈Z m=0 which yields the desired conclusion. As g ··· g = id , we have id is in N (0)(2λ), and hence in S N (m)(4λ). As 1 n G G m∈Z every element in G can be written as a product of finitely many elements in N(λ), it suffices to show for every g ∈ S N (m)(4λ) and g0 = gh with h ∈ N(λ) that g0 m∈Z is in S N (m)(4λ). The desired conclusion then follows from the induction on the m∈Z number of translations in N(λ). MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 49

Fix m which minimizes d(g, g1 . . . gm). We claim that d(g, g1 . . . gm) < 2λ + γ. 0 This claim gives us the desired conclusion because we then have d(g , g1 . . . gm) < 3λ + 2γ < 4λ by the γ-linearity of d. We now prove the claim that d(g, g1 . . . gm) < 2λ + γ. Suppose to the contrary −1 that d(g, g1 . . . gm) ≥ 2λ + γ. Let u = (g1 ··· gm) g. Now by Lemma 7.14 we have −1 −1 either s(u, gm+1gm+2) = 1, or s(u, gm gm−1) = 1. Suppose it is the former, since −1 −1 the latter case can be proved similarly. Then s(u, gm+2gm+1) = −1. Note that −1 −1 g = g1 ··· gmu = (g1 ··· gm+2)gm+2gm+1u. By the definition of u, and the linearity of d, we have kukd ≥ 2λ + γ > kgm+1gm+2kd, therefore by the irreducibility we have −1 −1 d(g, g1 ··· gm+2) = kgm+2gm+1ukd −1 −1 < kukd − kgm+2gm+1kd + γ < kukd − λ + γ < kukd.

This contradicts our choice of m having d(g, g1, . . . , gm) minimized. 

In the later proofs of this section, we will fix an irreducible λ sequence g1 ··· gn = idG to serve as “monitors”. As each element of G will be captured by one of the monitors, this will help us to bound the error terms in the final almost homomorphism we obtained from the pseudometric. Suppose d is γ-linear and γ-monotone, and N(ρ/4 − γ) \ N(4γ) 6= ∅. Define the returning weight of d to be

ω = inf{t(g1, . . . , gn):(g1, . . . , gn) is an irreducible λ-sequence with g1 ··· gn = idG}. The following corollary translate Lemma 7.16 to a bound on such ω: Corollary 7.17. Suppose d is γ-linear and γ-monotone, and ω is the returning weight of d. Then we have the following:

(i) λ/4µG(N(4λ)) ≤ ω ≤ 4λ/µG(N(λ)). (ii) There is an irreducible λ-sequence (g1, . . . , gn) such that ω = t(g1, . . . , gn) and 1/µG(N(4λ)) ≤ n ≤ 4/µG(N(λ)).

Proof. Note that each irreducible λ-sequence (g1, . . . , gn) has n ≥ 1/µG(N(4λ)) by using Lemma 7.16. Hence, by Corollary 7.15, we get ω ≥ λ/4µG(N(4λ)). On the k other hand, by Corollary 4.3, G = (N(λ)) for all k > 1/µG(N(λ)). Hence, with Lemma 7.13, there is an irreducible λ-sequence (g1, . . . , gn) with g1 ··· gn = idG and n ≤ 4/µG(N(λ)). From the definition of t, we get ω ≤ 4λ/µG(N(λ)). Now if an irreducible λ-sequence (g1, . . . , gn) has n > 4/µG(N(λ)), then by (i) and Corollary 7.15, 4λ t(g1, . . . , gn) > ≥ ω, µG(N(λ)) a contradiction. Therefore, we have

ω = inf{t(g1, . . . , gn):(g1, . . . gn) is an irreducible λ-sequence with MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 50

n ≤ 4/µG(N(λ)) and g1 ··· gn = idG}. For fixed n the set of irreducible λ-sequence of length n is closed under taking limit. Hence, we obtain desired (g1, . . . , gn) using the Bozalno–Wierstrass Theorem. 

The next lemma allows us to convert between µG(N(λ)) and µG(N(4λ)): Lemma 7.18. Suppose d is γ-linear and γ-monotone. Then

µG(N(4λ)) ≤ 16µG(N(λ)). Proof. Fix h ∈ N(λ) \ N(λ/2 − γ). Such h exists since by γ-monotonicity we have 2 N (λ/2−γ) ⊆ N(λ), and by Kemperman’s inequality, µG(N(λ) > 2µG(N(λ/2−γ)). Let g be an arbitrary element in N(4λ), and assume first s(g, h) = 1. Let k ≥ 0 be −1 k an integer, and define gk = g(h ) . Then by Proposition 7.9 and Lemma 7.10,

kgkkd ∈ kgkd − kkhkd + I(5kγ) for k < kgkd/khkd.

Hence, there is k < 8 such that gk ∈ N(λ). When s(g, h) = −1, one can similarly 0 k 0 construct gk as gh , and find k < 8 such that gk ∈ N(λ). Therefore 7 7  [   [  N(4λ) ⊆ N(λ)hi ∪ N(λ)h−j . i=0 j=0

Thus, µG(N(4λ)) ≤ 16µG(N(λ)).  The following proposition implicitly establish that t defines an approximate mul- tivalued group homomorphism from G to R/ωZ. Proposition 7.19. Suppose d is γ-linear and γ-monotone, ω is the returning weight of d, and (g1, . . . , gn) is a λ-sequence with g1 . . . gn = idG and n ≤ 4/µG(N(λ)). Then

t(g1, . . . , gn) ∈ ωZ + I(ω/400).

Proof. Let g0 be in N(ρ/4 − γ) \ N(4γ). Using Proposition 7.9(iii) to replace g0 with −1 g0 if necessary, we can assume that n X t(g1, . . . , gn) = s(g0, gi)kgikd. i=1

As n ≤ 4/µG(N(λ)), we have t(g1, . . . , gn) ≤ 4λ/µG(N(λ)). From Corollary 7.17(i), we have λ ≤ 4ωµG(N(4λ)). Hence,

16ωµG(N(4λ)) (16) t(g1, . . . , gn) < . µG(N(λ)) MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 51

Using Corollary 7.17 again, we obtain an irreducible λ-sequence (h1, . . . , hm) such that t(h1, . . . , hm) = ω and 1/µG(N(4λ)) ≤ m ≤ 4/µG(N(λ)). Using Proposi- −1 −1 tion 7.9(iii) to replace (h1, . . . , hm) with (hm , . . . , h1 ) if necessary, we can assume that n X t(h1, . . . , hm) = − s(g0, hi)khikd. i=1 0 0 We now define a sequence (g1, . . . , gn0 ) such that (1) n0 = n + km for some integer k ≥ 0. 0 (2) gi = gi for 1 ≤ i ≤ n. 0 (3) For i ≥ n + 1, gi = hj with j ≡ i − n (mod m). From the definition of the total weight, for k < t(g1, . . . , gn)/ω, we have 0 0 t(g1, . . . , gn0 ) = t(g1, . . . , gn) − kω. 0 0 We choose an integer k < t(g1, . . . , gn)/ω + 1 such that |t(g1, . . . , gn0 )| ≤ ω/2. Then by (16), and the trivial bound µG(N(λ)) < µG(N(4λ)), we have   0 4 16µG(N(4λ)) 4 72µG(N(4λ)) n < n + km < + + 1 ≤ 2 . µG(N(λ)) µG(N(λ)) µG(N(λ)) µG(N(λ)) 0 0 0 0 Note that (g1, . . . , gn0 ) is a λ-sequence with g1 . . . gn0 = idG. We assume further 0 0 that 0 ≤ t(g1, . . . , gn0 ) < ω/2 as the other case can be dealt with similarly. Obtain 0 0 0 0 an irreducible concatenation (h1, . . . , hm0 ) of (g1, . . . , gn0 ). From Lemma 7.13, we get

0 0 0 0 0 0 ω 1800µG(N(4λ))γ t(h1, . . . , hm0 ) < t(g1, . . . , gn0 ) + 25(n − m )γ ≤ + 2 . 2 µG(N(λ)) Using Corollary 7.17(i) and Lemma 7.18, we have 5 1800µG(N(4λ))γ 1800µG(N(4λ))γ 5 · 10 γ 5 4ω 2 ≤ 2 2 < ≤ 5 · 10 γ . µG(N(λ)) µG(N(4λ))/16 N(4λ) λ As γ < 10−8ρ, and λ = ρ/16 − γ, one can check that the lass expression is at 0 0 most ω/400. Hence, t(h1, . . . , hm0 ) < ω. From the definition of ω, we must have 0 0 t(h1, . . . , hm0 ) = 0. Thus by Lemma 7.13 again, 0 0 0 t(g1, . . . , gn) ∈ I(25n γ) ⊆ I(ω/400), which completes the proof.  Recall that a is a topological space which is separable and completely metrizable. In particular, the underlying topological space of any connected compact Lie group is a Polish space. Let X be a Polish space. A subset B of X is Borel if B can be formed from open subsets of X (equivalently, closed subsets of X) through taking countable unions, taking countable intersections, and taking complement. A function f : X → Y between Polish space is Borel, if the inverse image of any Borel MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 52 subset of Y is Borel. A subset A of X is analytic if it is the continuous image of another Polish space Y . Below are some standard facts about these notions; see [Kec95] for details. Fact 7.20. Suppose X,Y are Polish spaces, and f : X → Y is continuous. We have the following: (i) Every Borel subset of X is analytic. (ii) Equipping X × Y with the product topology, the graph of a Borel function from X to Y is analytic. (iii) The collection of analytic subsets of X is closed under taking countable unions, taking intersections and cartesian products. (iv) Images of analytic subsets in X under f is analytic.

Given x ∈ R, let kxkT be the distance of x to the nearest element in Z. We now obtain a consequence of Lemma 7.19. Corollary 7.21 (Analytic multivalued almost homomorphism). There is an analytic subset Γ of G × T satisfying the following properties: (i) The projection of Γ on G is surjective.

(ii) (idG, idR/ωZ) is in Γ. (iii) If g1, g2 ∈ G and t1, t2, t3 ∈ R are such that (g1, t1/ω + Z), (g2, t2/ω + Z), and (g1g2, t3/ω + Z) in Γ, then

k(t1 + t2 − t3)/ωkT < 1/400.

(iv) There are g1, g2 ∈ G and t1, t2 ∈ R with that (g1, t1/ω + Z), (g2, t2/ω + Z) ∈ Γ and k(t1 − t2)/ωkT > 1/3. Proof. Let Γ consist of (g, t/ω + Z) ∈ G × T with g ∈ G and t ∈ R such that there is n ≤ 1/µG(N(λ)) + 1 and an irreducible λ-sequence (g1, . . . , gn) satisfying

g = g1 ··· gn and t = t(g1, . . . , gn). Note that the relative sign function s : G × G → R is Borel, the set N[γ] is com- pact, and the function x → kxkd is continuous. Hence, by Fact 7.20(i,ii), the func- tion (g1, . . . , gn) 7→ t(g1, . . . , gn) is Borel, and its graph is analytic. For each n, by Fact 7.20(iii)

n Γen := {(g, t, g1, . . . , gn) ∈ G × R × G :

kgikd < λ for 1 ≤ i ≤ n, g = g1 ··· gn, t = t(g1, . . . , gn)} is analytic. Let Γn be the image of Γen under the continuous map

(g, t, g1, . . . , gn) 7→ (g, t/ω + Z). MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 53

Then by Fact 7.20(iv), Γ is analytic. Finally, Γ = S Γ is analytic by n n<1/µG(N(λ))+1 n Fact 7.20(iii). We now verify that Γ satisfies the desired properties. It is easy to see that (i) and (ii) are immediately from the construction, and (iii) is a consequence of Lemma 7.19. We now prove (iv). Using Corollary 7.17, we obtain an irreducible λ-sequence (g1, . . . , gn) with t(g1, . . . , gn) = ω and n < 4/µG(N(λ)). Note that

|t(g1, . . . , gk+1) − t(g1, . . . , gk)| ≤ λ.

Hence, there must be k ∈ {1, . . . n} such that ω/3 < t(g1, . . . , gk) < 2ω/3. Set t1 = 0 and t2 = t(g1, . . . , gk) for such k. It is then easy to see that k(t1 −t2)/ωkT > 1/3. 

To construct a group homomorphism from G to T, we will need three more facts. Recall the following measurable selection theorem from descriptive set the- ory; see [Bog07, Theorem 6.9.3]. Fact 7.22 (Kuratowski and Ryll–Nardzewski measurable selection theorem). Let (X, A) be a measurable space, Y a complete separable metric space equipped with the usual Borel σ-algebra, and F a function on X with values in the set of nonempty closed subsets of Y . Suppose that for every open U ⊆ Y , we have

{a ∈ X : F (a) ∩ U 6= ∅} ∈ A. Then F has a selection f : X → Y which is measurable with respect to A. A Polish group is a topological group whose underlying space is a Polish space. In particular, Lie groups are Polish groups. A subset A of a Polish space X is uni- versally measurable if A is measurable with respect to every complete probability measure on X for which every Borel set is measurable. In particular, every analytic set is universally measurable; see [Ros19] for details. A map f : X → Y between Polish spaces is universally measurable if inverse images of open sets are univer- sally measurable. We have the following recent result from descriptive set theory by [Ros19]; in fact, we will only apply it to Lie groups so a special case which follows from an earlier result by Weil [Wei40, page 50] suffices. Fact 7.23 (Rosendal). Suppose G and H are Polish groups, f : G → H is a univer- sally measurable group homomorphism. Then f is continuous. Finally, we need the following theorem from geometry by Grove, Karcher, and Ruh [GKR74] and independently by Kazhdan [Kaz82], that in a compact Lie groups an almost homomorphism is always close to a homomorphism uniformly. We remark that the result is not true for general compact topological groups, as a counterexam- ple is given in [vZ04]. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 54

Fact 7.24 (Grove–Karcher–Ruh; Kazhdan). Let G, H be compact Lie groups. There is a constant c only depending on H, such that for every real number q in [0, c], if π : G → H is a q-almost homomorphism, then there is a homomorphism χ : G → H which is 1.36q-close to π. Moreover, if π is universally measurable, then χ is universally measurable. When H = T, we can take c = π/6. The next theorem is the main result in this subsection. It tells us from an almost linear pseudometric, one can construct a group homomorphism to T or to R; this can also be seen as a stability theorem of Proposition 7.6. Theorem 7.25. Let λ = ρ/36, and γ < 10−6ρ. Suppose d is γ-linear and γ- monotone. Then there is a continuous surjective group homomorphism χ : G → T such that for all g ∈ ker(χ) ∩ N(λ), we have kgkd ∈ N(λ/2). Proof. Let ω be the returning weight of d, and let Γ be as in the proof of Corol- lary 7.21. Equip G with the σ-algebra A of universally measurable sets. Then A in particular consists of analytic subsets of G. Define F to be the function from G to the set of closed subsets of T given by F (g) = {t/ω + Z : t ∈ R, (g, t/ω + Z) ∈ Γ}. If U is an open subset of T, then {g ∈ G : F (g)∩U 6= ∅} is in A being the projection on G of the analytic set {(g, t/ω+Z) ∈ G×T :(g, t/ω+Z) ∈ Γ and t ∈ U}. Applying Fact 7.22, we get a universally measurable 1/400-almost homomorphism π : G → T. Using Fact 7.24, we get a universal measurable group homomorphism χ : G → T satisfying

kχ(g) − π(g)kT < 1.36/400 = 0.0068. The group homomorphism χ is automatically continuous by Fact 7.23. Combining with Corollary 7.21(iv), we see that χ cannot be the trivial group homomorphism, so χ is surjective. Finally, for g in ker(χ) ∩ (N(λ)), we need to verify that g is in N(λ/2). Suppose to the contrary that g∈ / N(λ/2). Choose n = b1/µG(N(4λ))c, and (g1, . . . , gn) the λ-sequence such that gi = g for i ∈ {1, . . . , n}. By Proposition 7.9, (g1, . . . , gn) is irreducible. Hence, by Lemma 7.16, t(g1, . . . , gn) < ω. As n ≤ 1/µG(N(λ) + 1), by construction and Corollary 7.21(iii), we have n π(g ) ∈ t(g1, . . . , gn)/ω + I(1/400) + Z = nkgkd/ω + I(1/400) + Z. n n Since g ∈ ker χ, we have kπ(g )kT < 0.0068, so nkgkd/ω < (0.0068 + 1/400). By Corollary 7.17(i) and Lemma 7.18, this implies (0.0068 + 1/400) · 4λ λ kgkd ≤ µG(N(4λ)) < , µG(N(λ)) 2 which is a contradiction. This completes our proof.  MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 55

8. Geometry of minimal and nearly minimal expansions II In this section, we study the shape of a nearly minimally expanding pair relative to a connected closed proper subgroup of the ambient Lie group such that the cosets of the subgroup intersect the nearly minimal expanding pair “transversally in measure”. Section 8.1 shows that in a compact connected Lie group, such a subgroup exists and can in fact be chosen to be a one-dimensional torus. In Section 8.2 and 8.3, we obtain shape description for the minimally expanding pair and nearly minimally expanding pair, respectively. Using that we will construct linear and almost linear pseudometric as described in Section 7. Throughout this section, G is a connected unimodular Lie group, and H is a connected unimodular closed subgroup of G. In particular, the left Haar measures µG and µH are Haar measures. We let A, and B be σ-compact subsets of G. We will assume familiarity with the preliminary Section 3.1 on locally compact group and Haar measure. We set µG/H and µH\G to be the Radon measures on G/H and H\G such that we have the quotient integral formulas (Fact 3.5 and Lemma 3.6(vi)). We also remind the reader that we normalize the measure whenever a group under consideration is compact, and π : G → G/H, πe : G → H\G are quotient maps. Hence, if H is compact, then we have

µG(AH) = µG/H (πA) and µG(HB) = µH\G(πBe ), and if χ : H → R is a continuous and surjective group homomorphism with compact kernel, then the pushforward of µH is the Lesbegue measure µR. 8.1. Toric transversal intersection in measure. In this section, we assume that G is compact. We will consider a more general situation than what we need assuming

µG(A) = µG(B) = κ.

We will prove that if κ is sufficiently small measure, and when µG(AB) < Mκ for some constant M, then there is a torus H ⊆ G such that we are in the short fiber scenario (i.e., for every x, y ∈ G,

(17) min{µH (A ∩ xH), µH (Hy ∩ B)} < λ for some given constant λ). Assume (17) fails for every maximal tori H, which means we have a long fiber “in every direction”. Then both A and B can be seen as a variant of Kakeya sets in Lie groups; see [MP15] for some properties of Kakeya sets in this setting. In general, it is well-known that Kakeya sets can have arbitrarily small measure; but when A, B has nearly minimally expansion, we will show in this section that both A and B must not be too small. Our result in particular applies to approximate groups, as described in Section 6.2. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 56

We use the following lemma, which can be seen as a corollary of the quotient integral formula (Fact 3.5). Lemma 8.1. For every b ∈ G, the following identity holds Z   µG A(B ∩ Hb) = µH (A ∩ aH)(B ∩ Hb) dµG(a). G Proof. Let C = A(B ∩ Hb). From Fact 3.5, one has Z Z −1 µG(C) = µH (C ∩ ab Hb) dµG(a) = µH (C ∩ aHb) dµG(a). G G Hence, it suffices to check that C ∩ aHb = (A ∩ aH)(B ∩ Hb) for all a, b ∈ G. The backward inclusion is clear. Note that aHHb = aHb = ab(b−1Hb) for all a and b in G. For all a, a0, and b in G, we have we have a0Hb = aHb when aH = a0H and aHb ∩ a0Hb = ∅ otherwise. An arbitrary element c ∈ C is in (A ∩ a0H)(B ∩ Hb) for some a0 ∈ A. Hence, if c is also in aHb, we must have a0H = aH. So we also get the forward inclusion. 

Suppose r and s are in R, the sets A(r,s] and πA(r,s] are given by

A(r,s] := {a ∈ A : µH (A ∩ aH) ∈ (r, s]} and πA(r,s] := {aH ∈ G/H : µH (A ∩ aH) ∈ (r, s]}.

In particular, πA(r,s] is the image of A(r,s] under the map π. By Lemma 3.6, πA(r,s] −1 is µG/H -measurable, and A(r,s]H = π (πA(r,s]) and A(r,s] are µG-measurable.

Lemma 8.2. Suppose µG(AB) < Mκ, and λ < 1 is a constant, and either there is a ∈ A such that µH (A ∩ aH) > λ, or there is b ∈ B such that µH (B ∩ Hb) > λ. Then,   µG(A) µG(B) λ min , ≥ 2 . µG/H (πA) µH\G(πBe ) (M + 2) Proof. Without loss of generality, suppose µH (B ∩ Hb) > λ for a fixed b ∈ B. By the quotient integral formula, κ = µG(A) is at least Z 1 1 µ (A ∩ xH) dµ xH > µ (πA ) = µ (A H), H G/H 2 G/H (1/2,1] 2 G (1/2,1] πA(1/2,1]

Hence, µG(A(1/2,1]H) < 2κ. We now prove that µG(A(0,1/2]H) < Mκ/λ. Suppose that it is not the case. By Lemma 8.1 we get Z µG(A(B ∩ Hb)) ≥ µH ((A ∩ aH)(B ∩ Hb)) dµG(a) A(0,1/2]H MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 57

Observe that µH ((A ∩ aH)(B ∩ Hb)) > λ since µH (B ∩ Hb) > λ, we have

µG(AB) ≥ µG(A(B ∩ Hb)) ≥ λµG(A(0,1/2]H) > Mκ, contradicting the assumption that dG(AB) < (M − 1)κ. Hence µG(AH) < (M + 2)κ/λ. This implies that there is a ∈ A such that µH (A ∩ aH) > λ/(M + 2). Now we apply the same argument switching the role of A and B, we get µH\G(πBe ) < 2 (M + 2) µG(B)/λ which completes the proof.  Lemma 8.2 leads us to consider the problem of obtaining lower bound for the measure of toric nonexpanders, which is of independent interest. Definition 8.3. We say A is called a toric K-expander, if there is a one-dimensional torus subgroup H of G such that µG(AH) ≥ KµG(A). The next lemma shows that the notion of toric expanders is stable under right translations and taking unions.

0 Lemma 8.4. Suppose A is not a toric K-expander, g1, . . . , gn are in G, and A = Sd 0 j=1 Agj. Then A is not a toric (nK)-expander. 0 0 Proof. We need to verify for each T that µG(A T ) < nKµG(A ). Note that n n n n 0  [  [ [ −1 [ A T = Agi T = A(giT ) = A(giT gi )gi = ATigi, i=1 i=1 i=1 i=1 −1 where Ti is the torus subgroup giT gi of G. Hence, 0 0 µG(A T ) < nKµG(A) ≤ nKµG(A ) as desired.  We will need the following inequality: Fact 8.5 (Bhatia–Davis inequality). Suppose (X, A, µ) is a measure space, α and β are constants, and f : X → R>0 is a measurable function with α ≤ inf f(x) < sup f(x) ≤ β x∈X x∈X 2 2 Then (Exf (x)) − (Exf(x)) ≤ (β − Exf(x))(Exf(x) − α). The following lemma will help us to translate the set along some direction. Lemma 8.6. Suppose K, α, and β are constant with K > 1, 0 ≤ α ≤ β ≤ 1, µG(AH) = KµG(A), and

α ≤ inf µH (A ∩ gH) ≤ sup µH (A ∩ gH) ≤ β. g∈A g∈A

Then for every number γ ≥ (α+β−Kαβ)µG(A), there is h ∈ H with µG(A∩Ah) = γ. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 58

Proof. Let µG, µH be normalized Haar measures of G and H. Choose h from H uniformly at random. Note that Z Eh∈H µG(A ∩ Ah) = µG(A ∩ Ah) dµH (h) H Z Z = 1A(g)1A(gh) dµG(g) dµH (h). H G Using the quotient integral formula (Fact 3.5), the above equality is Z Z 1 2 A(g)µH (A ∩ gH) dµG(g) = µH (A ∩ gH) dµG/H (gH) G G/H 2  = EgH∈G/H µH (A ∩ gH) 2  = µG/H (πA)EgH∈πA µH (A ∩ gH)

Note that µG/H (πA) = µG(AH) = KµG(A), and  EgH∈πA µH (A ∩ gH) = 1/K. Hence, applying the Bhatia–Davis inequality (Fact 8.5), we get

Eh∈H µG(A ∩ Ah) ≤ (α + β − Kαβ)µG(A). The desired conclusion follows from the continuity of H → R, h 7→ µG(A ∩ Ah).  The following lemma says that for a toric nonexpander A and a torus subgroup H of G, one can slightly modify A to get A0 such that most of A0H can be covered by finitely many right translations of A0. Lemma 8.7. Suppose K > 1 is a constant, A is not a toric K-expander, and H is a one-dimensional torus subgroup of G. Then for every 0 < ε < 1/2K, there is a 0 σ-compact A ⊆ A, integer m = m(K, ε) > 0, and h1, . . . , hm ∈ H, satisfying (i) A0 is not a toric 2K-expander 0 (ii) µG(A ) > (1 − εK)µG(A) 0 Sm 0 (iii) µG (A H \ i=1 A hi) < εµG(AH).

Proof. Let µG, µH be normalized Haar measures on G and H, and let µG/H be the invariant Radon measure induced by µG and µH on the homogeneous space G/H. Let

πA(ε,1] = {g ∈ G/H | ε < µH (A ∩ gH) ≤ 1},

πA(0,1] = {g ∈ G/H | A ∩ gH 6= ∅}.

Let x = µG/H (πA(0,ε))/µG/H (πA(0,1]), then µG/H (πA(ε,1]) = (1 − x)(µG/H (πA(0,1]). One has µ (πA ) G/H (0,1] < µ (A) ≤ (εx + 1 − x)µ (πA )). K G G/H (0,1] MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 59

It follows that µ (πA ) K − 1 x = G/H (0,ε] < . µG/H (πA(0,1]) K(1 − ε) 0 0 Choose σ-compact A ⊆ A(ε,1] such that µG(A(ε,1] \ A ) = 0. One has K − 1 µ (A0) ≥ µ (A) − ε µ (πA ) G G K(1 − ε) G/H (0,1]  K − 1 1 − εK ≥ 1 − ε µ (A) = µ (A) ≥ (1 − εK)µ (A). 1 − ε G 1 − ε G G Hence we have 1 − εK 1 µ (A0) ≥ (1 − εK)µ (A) ≥ µ (AT ) > µ (A0T ), G G K G 2K G for every torus T when ε < 1/2K. It remains to obtain m = m(ε, K) and h1, . . . , hm such that (iii) is satisfied. Con- 0 0 0 struct a sequence (An) of σ-compact subsets of G with AnH = A H as follows. Let 0 0 0 A0 = A . Suppose An has been constructed. Set 0 0 0 εn = inf µH (An ∩ gH) and Kn = µG(AnH)/µG(An). gH∈πA 0 Using Lemma 8.6, obtain hn ∈ H such that 0 0 0 0 0 µG(A hn \ A ) = εn(Kn − 1)µG(An). 0 0 0 0 0 Finally, let An+1 = An ∪ Anhn. Then, An+1H = AnH = A H. It is also easy to see 0 that εn ≥ ε for all n. Now, if µG(An) < (1 − ε)µG(A H) for some n, then 0 µG(A H) 1 Kn = ≥ , µG(An) 1 − ε and hence, 1 − ε + ε2 µ (A0 ) ≥ µ (A0 ). G n+1 1 − ε G n As (1 − ε + ε2)/(1 − ε) > 1, this cannot be the case for all n. Let N be the first n 0 0 such that µG(AN ) > (1 − ε)µG(A H). Then 1 − ε + ε2 N 2Kµ (A0) > µ (A0H) ≥ µ (A0 ) ≥ µ (A0). G G G N 1 − ε G This implies that log 2K N ≤ . log(1 − ε + ε2) − log(1 − ε) N 0 Sm 0 Finally set m = 2 , and choose h1, . . . , hm such that AN = i=1 A hi, we get the desired conclusion.  MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 60

The next simple fact shows that we can find finitely many tori such that the product of them is G. Fact 8.8. Let G be compact. Then there is a constant n depending only on the dimension of G such that there are n tori H1,...,Hn in G with H1 ··· Hn = G. Suppose A is not a toric expander. The next important “cage” lemma provides an inductive construction to construct a set C from A, such that the size of C is bounded from above, and any right translations of A cannot “escape” C. Lemma 8.9 (Cage lemma). Suppose A ⊆ G is not a toric K-expander, then there is a σ-compact C ⊆ A such that µG(C) = OK (1)µG(A) and for all g ∈ G µ (C ∩ Ag) 1 G > . µG(A) 2

Proof. Using Fact 8.8, we obtain one-dimensional torus subgroups H1,...,Hn of G such that G = H1 ··· Hn. For every constant ε0, . . . , εn−1, we construct a sequence n n (Ai)i=0 of σ-compact subsets of G and a sequence (Ki)i=0 of constants satisfying the following conditions

(1) A0 = A and K0 = K. (2) Ai is not a toric Ki-expander for 0 ≤ i ≤ n. (3) Ai ⊆ Ai+1 with µG(Ai+1) ≤ KiµG(Ai) for 0 ≤ i ≤ n − 1. (4) µG(Ai \ Ai+1hi+1) ≤ εiµG(Ai) for any hi+1 ∈ Hi+1 and 0 ≤ i ≤ n − 1. (5) Ki+1 = Ki(ε0, . . . , εi) for 0 ≤ i ≤ n − 1.

Suppose we have A0,...,Ai and K0,...,Ki satisfying all the conditions restricted down to i. We are going to construct Ai+1. By (2), Ai is not a toric Ki-expander. Let δ > 0 be a parameter which will be determined later. Using Lemma 8.7, we 0 0 0 obtain a σ-compact Ai ⊆ Ai, m = m(Ki, δ), and h1, . . . , hm ∈ Hi+1 such that 0 µG(Ai) > (1 − δKi)µG(Ai), 0 Ai is not a toric 2Ki-expander, and  m  0 [ 0 0 0 µG AiHi+1 \ Aihj < δµG(AiHi+1). j=1 0 0 By adding one element of Hi+1 if necessary, we can arrange that idG is in {h1, . . . , hm}. Sm 0 0 Set Ai+1 = j=1 Aihj and set Ki+1 = mKi. By Lemma 8.4, Ai+1 is not a toric Ki+1- expander, so (2) is satisfied. By construction Ai ⊆ Ai+1, and 0 µG(Ai+1) ≤ µG(AiHi+1) ≤ µG(AiHi+1) < KiµG(Ai), 0 0 0 so we have (3). For every h ∈ Hi+1, since Ai+1h ⊆ AiHi+1, we have 0 0 0 µG(Ai \ Ai+1h ) < δµG(AiHi+1) ≤ δµG(AiHi+1) < δKiµG(Ai). MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 61

Therefore, 0 µG(Ai \ Ai+1h ) < 2δKiµG(Ai).

Note that the construction so far depends on δ. Now, by choosing δ = δ(Ki, εi) sufficiently small, we can make 0 µG(Ai \ Ai+1h ) < εiµG(Ai), so we get (4). Finally, note that Ki+1 = mKi, m = m(Ki, δ), δ = δ(Ki, εi), and Ki = Ki(ε0, . . . , εi−1), so Ki+1 = Ki+1(ε0, . . . , εi), which gives us (5). We now proceed with the proof of the lemma. Let ε0, . . . , εn−1 be parameters n n which we will determine later, and obtain (Ai)i=0 and (Ki)i=0 as in the earlier step. Set C = An. Note that −1 −1 G = G = (H1 ...Hn) = Hn ··· H1.

Hence, an arbitrary g ∈ G can be written as a product hn ··· h1 with hi ∈ Hi for i ∈ {1, . . . , n}. Now consider Cg = Anhn ··· h1. By (4), µG(An−1 \ Anhn) < εn−1µG(An−1). Next, for Anhnhn−1, again by (4),

µG(An−2 \ Anhnhn−1) ≤ µG(An−2 \ An−1hn−1) + µG(An−1hn−1 \ Anhnhn−1)

< εn−2µG(An−2) + εn−1µG(An−1). Hence, by induction and (3), we conclude that n−1 n−1 i−1 ! X X Y µG(A \ Cg) < εiµG(Ai) ≤ εi Kj µG(A). i=0 i=0 j=0 Pn−1 Qi−1  Using (5), we can choose εi sufficiently small such that i=0 εi j=0 Kj < 1/2. Then, for all g ∈ G. µ (C ∩ Ag) µ (Cg−1 ∩ A) 1 G = G > . µG(A) µG(A) 2

Finally, note that we can choose ε0, . . . , εn−1 depending only on K. Hence,

µG(C) = OK (1)µG(A), which completes the proof.  Suppose µ and ν are measures on G. Their convolution µ ∗ ν is the unique measure satisfying the property Z Z f(x) dµ ∗ ν(x) = f(xy) dµ(x) dν(y). G G×G MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 62

The convolution exists for all the case we care about. If µ(A) > 0, the uniform measure on A is defined by

µG(A ∩ X) µA(X) := for measurable X ⊆ G. µG(A) The following lemma is an immediate consequence of the definition and Fubini’s theorem.

Lemma 8.10. Let µA be the uniform measure on A. Then µA ∗ µG = µG. Proof. Let X be a measurable set in G, then Z Z  Z µA ∗ µG(X) = 1X (xy) dµG(y) dµA(x) = µG(X) dµA(x) = µG(X) G G G as desired.  The next proposition gives us a lower bound control on the measure of sets which are not toric expanders. Together with Lemma 8.2, this quantitative result shows that a sufficiently small set which inside some nearly minimal expansion pair cannot contain a long fiber in every direction. Proposition 8.11 (Bounding size of toric nonexpanders). For each K, there is S = OK (1) such that if A is not a toric K-expander, then µG(A) > S.

Proof. Let C be as in Lemma 8.9. Recall µA is the uniform measure on A, and µA ∗ µG is the convolution measure. Then Z Z Z 1C (x) dµA ∗ µG(x) = 1C (xy) dµA(x) dµG(y) G G G Z −1 µG(Cy ∩ A) 1 = dµG(y) ≥ . G µG(A) 2

This means µA ∗ µG(C) ≥ 1/2. By Lemma 8.10, this implies µG(C) ≥ 1/2. Since µG(C) = OK (1)µG(A) we get the desired conclusion.  We now deduce the main theorem of this section.

Theorem 8.12. There is S = OM,λ(1) such that if µG(A) = µG(B) = κ < S and µG(AB) < Mκ, then there is a one-dimensional torus subgroup H of G such that for all x, y ∈ G µH (A ∩ xH) < λ and µH (B ∩ Hy) < λ.

Proof. Suppose for every one-dimensional torus subgroup H of G, either µH (A ∩ xH) > λ some x ∈ G or µH (B ∩ By) > λ for some y ∈ G. Then by Lemma 8.2, A is not a toric K-expander with K = (M + 2)2/λ. Hence, by Proposition 8.11, we have µG(A) > S with S = OM,λ(1). Thus, we get the desired conclusion.  MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 63

By computing the constant coefficients carefully in Lemma 8.9, the constant S in d+2 Theorem 8.12 is of order M −M 2 . We did not try to optimise this number in the proof, and we suspect a single exponential bound should be enough. We get the following immediate corollary for approximate groups. Since our proof is quantitative, we can make the constant below quantitative if we wish.

Corollary 8.13. There is S = OK (1) such that if A is a K-approximate group with µG(A) < S, then there is a one-dimensional torus subgroup H of G such that for all x, y ∈ G µH (A ∩ xH) < λ and µH (B ∩ Hy) < λ. 8.2. Linear pseudometric from minimal expansions. Throughout the subsec- tion G is a connected noncompact unimodular group, and H is a closed subgroup of G which is either isomorphic to R, or some smaller dimensional connected unimodu- lar group, so that by induction on dimension we may assume Theorem 1.1 holds on H. Suppose (A, B) is minimally expanding, that is,

µG(AB) = µG(A) + µG(B), and both A, B have positive measure. This section can also be seen as a preview of Section 8.3. The strategy of this section also works for compact G replacing H with T and results of Section 8.1. For convenience of notation, we will treat the compact case in Section 8.3 together with the situation where (A, B) is nearly minimally expanding. Lemma 8.14. For all a ∈ A and b ∈ B, we have  (1) µH (A ∩ aH)(B ∩ Hb) ≥ µH (A ∩ aH) + µH (B ∩ Hb).  (2) µG A(B ∩ Hb) ≥ µG(A) + µG/H (πA)µH (B ∩ Hb).  (3) µG (A ∩ aH)B ≥ µH (A ∩ aH)µH\G(πBe ) + µG(B). The equality in (2) holds if and only if the equality in (1) holds for almost all a ∈ AH. A similar conclusion holds for (3). Proof. The first inequality comes from a direct application of the Kemperman in- equality. For the second inequality, by right translating B and using the unimodular- ity of G, we can arrange that Hb = H. The desired conclusion follows from applying (1) and Lemma 8.1.  The next theorem gives us the important geometric properties of A and B. Theorem 8.15 (Rigidity fiberwise). There is a continuous surjective group homo- morphism χ : H → R, two compact intervals I,J ⊆ R with µG(A) µG(B) µR(I) = and µR(J) = , µG/H (πA) µH\G(πBe ) MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 64

σ-compact sets A0 ⊆ A and B0 ⊆ B with 0 0 µG/H (πA ) = µG/H (πA) and µH\G(πBe ) = µH\G(πBe ), such that the following hold: (i) µG/H (πA) = µH\G(πBe ); (ii) for each a ∈ A0H, we have

µG(A) µH (A ∩ aH) = , µG/H (πA)

and there is ζa ∈ R such that −1  µH (A ∩ aH) 4 aχ (ζa + I) = 0. (iii) for each b ∈ HB0, we have

µG(B) µH (B ∩ Hb) = , µH\G(πBe )

and there is ζeb ∈ R such that −1  µH (B ∩ Hb) 4 χ (ζeb + J)b = 0.

Proof. Without loss of generality we assume µG/H (πA) ≥ µH\G(πBe ). Below, we let Hb range over πBe , and choose Hb uniformly at random in the expectation. By Lemma 8.1 and the quotient integration formula, we have

sup µG(A(B ∩ Hb)) ≥ µG(A) + µG/H (πA) sup µH (B ∩ Hb)(18) Hb Hb ≥ µG(A) + µG/H (πA)EHbµH (B ∩ Hb)

µG/H (πA) = µG(A) + µG(B) ≥ µG(A) + µG(B). µH\G(πBe )

Note that µG(AB) ≥ supHb µG(A(B ∩ Hb)). Since µG(AB) = µG(A) + µG(B), the equality must hold at each step. In particular, we have

(19) µG/H (πA) = µH\G(πBe ), and for µH\B-almost all Hb ∈ πe(B), we have

0 µG(B) µH (B ∩ Hb) = EHb0 (B ∩ Hb ) = . µH\G(πBe ) Now, using (19) and applying the same argument again switching the role of A and B, we conclude that for µG/H -almost all aH in πA, we have

0 µG(A) (20) µH (A ∩ aH) = Ea0H µH (A ∩ a H) = . µG/H (πA) MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 65

Moreover, the fact that equality holds in (18) shows that for µG/H -almost all aH ∈ πA and µH\G-almost all Hb ∈ πBe we have

µH ((A ∩ aH)(B ∩ Hb)) = µH (A ∩ aH) + µH (B ∩ Hb).

By the relationship between µG and µG/H , in the preceding statement, we can replace µG/H -almost all aH ∈ πA with µG-almost all a ∈ AH. We can do the same for µG and µH\G. Now, as H satisfies Theorem 1.1, for a an b such that (20) holds, we can choose continuous surjective group homomorphisms χa, χb : H → R and compact intervals Ia, and Jb in R with

µR(Ia) = µH (A ∩ aH), µR(Jb) = µH (B ∩ Hb), and −1  −1  µH (A ∩ aH) 4 χa (Ia) = 0, µH (B ∩ Hb) 4 χb (Jb) = 0.

Applying Fact 3.12, we deduce that χa, χb are the same for µG-almost all a ∈ AH and µG-almost all b ∈ HB. We also have Ia and Jb have constant lengths for µG- almost all a ∈ AH and µG-almost all b ∈ HB. Hence there are ζa, ζeb ∈ T for almost all a ∈ AH and almost all b ∈ HB, such that Ia = ζa + I, and Jb = ζeb + J, where  µ (A)   µ (A)  I = 0, G and J = 0, G . µG/H (πA) µH\G(πBe ) It follows that we can choose A0 ⊆ A and B0 ⊆ B as described in the statement of the theorem.  Corollary 8.16 (Global structure of AH). For all g ∈ G, we have

µG/H (πA 4 π(gA)) = 0.

<2ρ Proof. Set ρ = µG(A). Recall that StabG (A) is open in G, and every g ∈ G can <2ρ be expressed as a finite products of elements in StabG (A) since G is connected. <2ρ Therefore, it suffices to consider the case where g is in StabG (A). Clearly, (gA, B) is a minimally expanding pair. In the current case, we also have µG(A \ gA) < ρ and µG(A ∩ gA) > 0. By Lemma 5.2, (A ∪ gA, B) is also a minimally expanding pair. Theorem 8.15(i) then gives us

µG/H (πA) = µG/H (π(gA)) = µG/H (πA ∪ π(gA)) = µH\G(πB).

This gives us µG/H (πA 4 π(gA)) = 0 as desired.  Theorem 8.15 and Corollary 8.16 essentially allows us to define a “directed linear pseudometric” on G by “looking at the generic fiber” as discussed in the following remark: MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 66

Remark 8.17. Fix a ∈ AH and let the notation be as in Theorem 8.15. For g1, g2 −1 −1 0 in G such that g1 a, g2 a ∈ A H, set

δa,A(g1, g2) = ζ −1 − ζ −1 . g1 a g2 a

We have the following linearity property of δa,A when the relevant terms are defined, which is essentially the linearity property of the metric from R. (1) δa,A(g1, g1)=0. (2) δa,A(g1, g2) = δa,A(g2, g1). (3) δa,A(g1, g3) = δa,A(g1, g2) + δa,A(g2, g3). Properties (1) and (2) are immediate, and property (3) follows from the easy calcu- lation below:

δa,A(g1, g2) = ζ −1 − ζ −1 g1 a g2 a = ζ −1 − ζ −1 + ζ −1 − ζ −1 g1 a g3 a g3 a g2 a = δa,A(g1, g3) ± δa,A(g3, g2). Properties (3) also implies that

|δa,A(g1, g3)| = ± |δa,A(g1, g2)| ± |δa,A(g2, g3)| . which tells us that |δa,A| is a linear pseudometric. The problem with the above definitions is that they are not defined everywhere. There are two ways to overcome this difficulty. The new approach, using difference in measure, will be presented later on. The old approach, presented in an earlier version of this paper, is to define a pseudometric on G directly by setting

d(g1, g2) = ξ if for µG-almost all a ∈ AH, |δa,A(g1, g2)| = ξ. This is indeed possible. In fact, one can bypass the pseudometric machinery alto- gether and define the group homomorphism χ : G → T directly by setting

χ(g) = ζ if for µG-almost all a ∈ AH, |δa,A(idG, g)| = ζ. However, this does not come for free, and one need to work equally hard to verify that χ is indeed a group homomorphism. The old approach can moreover be extended to the case of nearly minimal expan- sion. However, we can only handle a quadratic error with this old approach because we only have Corollary 8.21, which lacks the global property of Corollary 8.16. The real problem solved by introducing the pseudometric is to get the linear error bound for the nearly minimal expansion problem.

Recall that for every g1, g2 ∈ G, dA(g1, g2) = µG(g1A \ g2A). By Proposition 7.1,1 dA is a pseudometric on G. The next lemma builds an important bridge between fiberwise information δa(g1, g2) to the pseudometric dA(g1, g2). MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 67

Lemma 8.18 (From local to global). Let χ : H → R be as in Theorem 8.15. For 00 all g1, g2 ∈ G with µG(g1A ∩ g2A) > 0, there is a σ-compact A ⊆ A with 00 µG/H (πA ) = µG/H (πA) such that for all a ∈ A00H, the following holds

(i) there are ζ −1 , ζ −1 ∈ such that for i ∈ {1, 2}; g1 a g2 a T −1  µH (A ∩ aH) 4 aχ (ζ −1 + I) = 0. gi a

(ii) with any ζ −1 , ζ −1 satisfying (i) and δa,A(g1, g2) = ζ −1 − ζ −1 , if we have g1 a g2 a g2 a g1 a µG(g1A ∩ g2A) > 0, then

dA(g1, g2) = µG/H (πA)|δa,A(g1, g2)|. 0 00 Proof. Obtain A , I, J, ζa as in Theorem 8.15. Let A ⊆ G be the σ-compact set −1 −1 0 {a ∈ A : g1 a, g2 a ∈ A H}. 00 00 By Corollary 8.16, µG/H (πA ) = µG/H (πA). Fix a ∈ A . We then have −1 −1 −1 −1 −1 −1 A ∩ g aH = g aχ (ζ −1 + I) and A ∩ g aH = g aχ (ζ −1 + I). 1 1 g1 a 2 2 g2 a

Multiplying by g1 and g2 respectively, we get (i). As µG(g1A ∩ g2A) > 0, by Lemma 5.2, (g1A ∩ g2A, B) is minimally expanding. From Theorem 8.15(ii), for µG/H -almost all aH ∈ π(g1A ∩ g2A), we have

µG(g1A) µG(g1A ∩ g2A) µH (g1A ∩ aH) = and µH (g1A ∩ g2A ∩ aH) = . µG/H (π(g1A)) µG/H (π(g1A ∩ g2A))

Note that π(g1A ∩ g2A) ⊆ π(g1A) ∩ π(g2A). However, by Theorem 8.15,

µG/H (π(g1A ∩ g2A)) = µH\G(πBe ) = µG/H (π(g1A)) = µG/H (π(g2A)). Combining with Corollary 8.16, we get

µG/H (π(g1A) 4 πA) = 0 and µG/H (π(g1A ∩ g2A) 4 πA) = 0. Hence, removing a set of measure 0 from the above A00 if necessary, we can arrange that for all a ∈ A00H,

µG(g1A) µG(g1A ∩ g2A) µH (g1A ∩ aH) = and µH (g1A ∩ g2A ∩ aH) = . µG/H (πA) µG/H (π(A)) Finally, from (i), for all a ∈ A00H, we have

µH (g1A) − µH (g1A ∩ g2A ∩ aH) = |ζ −1 − ζ −1 |. g1 a g1 a Recalling that dA(g1, g2) = µG(g1A)−µG(g1A∩g2A), we learn that (ii) is satisfied.  We now prove the pseudometric is linear as promised. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 68

<ρ Proposition 8.19 (Linearity of the pseudometric). For all, g1, g2, g3 in StabG (A), we have

dA(g1, g2) ∈ {dA(g1, g3) + dA(g3, g2), |dA(g1, g3) − dA(g3, g2)|}. <ρ Proof. For i ∈ {1, 2, 3}, gi is in StabG (A) by assumption, so dA(idG, gi) < ρ/2. Hence, for i, j ∈ {1, 2, 3} we have

dA(gi, gj) < ρ and µG(giA ∩ gjA) > 0. 000 000 Applying Lemma 8.18(ii), we get σ-compact A ⊆ A with µG(A H) 6= ∅ such that for each a ∈ A000H, we have

µG(giA ∩ gjA) = µG/H (πA)|ζ −1 − ζ −1 | for i, j ∈ {1, 2, 3}. gi a gj a The desired conclusion immediately follows.  8.3. Almost linear pseudometric from near minimal expansions. In this sub- section, G is always a connected compact Lie group. Let H be a closed subgroup of G, and H is isomorphic to the one dimension torus T. Throughout the subsection, A, B ⊆ G are σ-compact subsets such that

κ/2 < µG(A) < 2κ and µG(B) = κ, and dG(A, B) ≤ ηκ for some constant η > 0, that is

µG(AB) ≤ µG(A) + µG(B) + ηκ. In this section, we assume η ≤ 10−12. We did not try to optimise η, so it is very likely that by a more careful computation, one can make η much larger. But we believe this method does not allow η to be very close to 1. By Fact 3.10, let τ = 2, and cB = c(τ) be the constant obtained from the theorem. In this subsection, we consider the case when

max{µH (A ∩ aH), µH (B ∩ Hb)} < cB for all a, b ∈ G. The proofs in this section is more involved compared to the equality case, and the main difficulty is to control the error term coming from the nearly minimally expanding pairs. For the readers who do not care the exact quantitative bound on the error terms, one can always view η as an infinitesimal element, then one can use equalities to replace all the inequalities in the proofs by pretending to take the standard part, and apply the methods given in the previous section. Towards showing that sets A and B behave rigidly, our next theorem shows that most of the nonempty fibers in A and B have the similar lengths, and the majority of them behaves rigidly fiberwise. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 69

Theorem 8.20 (Near rigidity fiberwise). There is a continuous surjective group homomorphism χ : H → T, two compact intervals I,J ⊆ T with

µG(A) µH (B) µT(I) = and µT(J) = . µG/H (πA) µH\G(πBe ) σ-compact A0 ⊆ A and B0 ⊆ B with 0 0 µG/H (πA ) > 99µG/H (πA)/100 and µH\G(πBe ) > 99µH\G(πBe )/100, and a constant ν ≤ 10−10 such that the following statements hold: (i) we have 1 µ (πB) ≤ µ (πA) ≤ (1 + η)µ (πB). 1 + η H\G e G/H H\G e (ii) For every a in A0H,

µG(A) µG(A) (1 − ν) ≤ µH (A ∩ aH) ≤ (1 + ν) , µG/H (πA) µG/H (πA)

and there is ζa ∈ T with

−1  n µG(A) µG(B) o µH (A ∩ aH) 4 aχ (ζa + I) < ν min , . µG/H (πA) µH\G(πBe ) (iii) For every b in HB0,

µG(B) µG(B) (1 − ν) ≤ µH (B ∩ Hb) ≤ (1 + ν) , µH\G(πBe ) µH\G(πBe )

and there is ζeb ∈ T with

−1  n µG(A) µG(B) o µH (B ∩ Hb) 4 χ (ζeb(B) + J)b < ν min , . µG/H (πA) µH\G(πe)

Proof. Without loss of generality, we assume that µG/H (πA) ≥ µH\G(πB). Let β be ∗ e a constant such that β < (κ/800µH\G(πBe )). Obtain b ∈ G such that ∗ µH (B ∩ Hb ) ≥ sup µH (B ∩ Hb) − β for all b ∈ G, b and the fiber B ∩ Hb∗ has at least the average length, that is

∗ µG(B) (21) µH (B ∩ Hb ) ≥ Eb∈BH µH (B ∩ Hb) = . µH\G(πBe ) −12 −10 Set δ = 100ηκ/µG/H (πA). As η ≤ 10 , we get ν ≤ 10 such that

(22) δ < νµG(A)/µG/H (πA). MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 70

Set ∗ N = {a ∈ AH : dH (A ∩ aH, B ∩ Hb ) > δ}. Note that N is measurable by Lemma 3.6. By Lemma 8.1 we have ∗  µG A(B ∩ Hb ) Z Z ∗  ∗  = µH (A ∩ aH)(B ∩ Hb ) dµG(a) + µH (A ∩ aH)(B ∩ Hb ) dµG(a). N G\N Since A ∩ aH is nonempty for every a ∈ AH, using Kemperman’s inequality on H ∗ we have that µG(A(B ∩ Hb )) is at least Z Z ∗  ∗  µH (A∩aH)+µH (B∩Hb )+δ dµG(a)+ µH (A∩aH)+µH (B∩Hb ) dµG(a). N G\N ∗ Suppose we have µG(N) > µG/H (πA)/100. Therefore, by the choice of b we get δµ (πA) µ A(B ∩ Hb∗) > µ (A) + G/H + µ (B ∩ Hb∗)µ (πA)(23) G G 100 H G/H µG/H (πA) ≥ µG(A) + µG(B) + ηκ. µH\G(πBe ) ∗ Since µG/H (πA) ≥ µH\G(πBe ), and A(B ∩ Hb ) ⊆ AB, we have

µG(AB) > µG(A) + µG(B) + ηκ.

This contradicts the assumption that dG(A, B) is at most ηκ. Using the argument in equation (23) with trivial lower bound on µG(N) we also get  (24) µH\G(πBe ) ≤ µG/H (πA) ≤ 1 + η µH\G(πBe ), which proves (i). From now on, we assume that µG(N) ≤ µG/H (πA)/100. Since dG(A, B) is at most ηκ, by (23) again (using trivial lower bound on µG(N)), we have ∗ µH (B ∩ Hb )µH\G(πBe ) ≤ µG(B) + ηκ, and this in particular implies that for every b ∈ G, we have

µG(B) ηµG(B)  µG(B) µH (B ∩ Hb) ≤ + + η < 1 + ν . µH\G(πBe ) µH\G(πBe ) µH\G(πBe ) Thus there is Y ⊆ B with µG(HY ) < µH\G(πBe )/100 such that for every b ∈ HY , µG(B) ηµG(B)  µG(B) µH (B ∩ Hb) ≥ − 100 − 100η > 1 − ν . µH\G(πBe ) µH\G(πBe ) µH\G(πBe ) MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 71

Next, we apply the similar argument to A. Let α < (µG(A) − 2ηκ)/200µG/H (πA), ∗ ∗ and choose a such that µH (A ∩ a H) > µH (A ∩ aH) − α for all a ∈ AH, and

∗ µG(A) µH (A ∩ a H) ≥ Ea∈AH µH (A ∩ aH) = . µG/H (πA) 0 0 ∗ Let N ⊆ HB such that for every b in N , dH (A ∩ a H,B ∩ Hb) ≥ δ. Hence we have ∗  µG (A ∩ a H)B Z Z ∗  ∗  = µH (A ∩ a H)(B ∩ Hb) dµG(b) + µH (A ∩ a H)(B ∩ Hb) dµG(b) N 0 G\N 0 (25) ∗ 0 ≥ µG(B) + µH (A ∩ a H)µH\G(πBe ) + δµG(N ) µG(A)ηκ 0 ≥ µG(A) + µG(B) − + δµG(N ). µG(A) + ηκ ∗ By the fact that µG(AB) ≥ µG((A ∩ a H)B) and dG(A, B) ≤ ηκ, we have that 1 1 µ (N 0) ≤ µ (πA) ≤ µ (πB). G 200 G/H 150 H\G e Now, by equation (25), and the choice of a∗, we have that for all a ∈ AH,

µG(A) ηµG(A)  µG(A) µH (A ∩ aH) ≤ + + α < 1 + ν . µG/H (πA) µH\G(πBe ) µG/H (πA)

Again by equation (25), there is X ⊆ A with µG(XH) ≤ µG/H (πA)/200, such that for every a ∈ X,

µG(A) ηµG(A)  µG(A) µH (A ∩ aH) ≥ − 200 − 200α ≥ 1 − ν . µG/H (πA) µH\G(πBe ) µG/H (πA) Let A0 = A ∩ (AH \ (XH ∪ N 0)), and let B0 = B ∩ (HB \ (HY ∪ N)). Then 99 99 µ (A0) ≥ µ (πA), µ (B0) ≥ µ (πB), G 100 G/H G 100 H\G e Let a be in A0H and b be in B0H. By our construction, the first parts of (ii) and ∗ ∗ (iii) are satisfied. Moreover, both dH (A ∩ aH, B ∩ Hb ) and dH (A ∩ a H,B ∩ Hb) are at most δ. By the way we construct A0 and B0, we have that a∗ ∈ A0 and b∗ ∈ B0. Recall that µH (A ∩ aH), µH (B ∩ Hb) < λ for every a, b ∈ G. Therefore, by the inverse theorem on T (Fact 3.10), and Lemma 6.12, there is a group homomorphism χ : H → T, and two compact intervals IA, IB in T, with

µG(A) µG(B) µT(IA) = , µT(IB) = , µG/H (πA) µH\G(πBe ) MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 72

0 0 such that for every a ∈ A and b ∈ B , there are elements ζa, ζeb(B) in T, and −1 −1 µH (A ∩ aH 4 aχ (ζa + IA)) < δ, µH (B ∩ Hb 4 χ (ζeb + IB)b) < δ, and the theorem follows from (22).  The next corollary gives us an important fact of the structure of the projection of A on G/H.

Corollary 8.21 (Global structure of AH). Suppose µG(A) = κ. Then for all g ∈ κ/2 StabG (A), we have µG(AH 4 gAH) ≤ 10ηµG/H (πA).

Proof. By Lemma 5.2, dG(A ∪ gA, B) ≤ 2ηκ, and by Theorem 8.20, we have 1 µ (AH ∪ gAH) < G < 1 + 2η. 1 + 2η µH\G(πBe ) On the other hand, since dG(A, B) and dG(gA, B) are at most ηκ, we have 1 µ (AH) µ (gAH) < G , G < 1 + η 1 + η µH\G(πBe ) µH\G(πBe ) −12 Since η < 10 , by inclusion-exclusion principle, we get the desired conclusion.  Given τ < 1/4, we use I(τ) to denote the open interval (−τ, τ) in T. As in Remark 8.17, one can similarly consider

δa,A(g1, g2) = ζ −1 − ζ −1 g1 a g2 a for a fix a ∈ A. Recall that dA = µG(g1A \ g2A) is a pseudometric (Proposition 7.1). The next lemma can be seen as an approximate version of Lemma 8.18, which gives the connection between δa,A(g1, g2) and dA(g1, g2). κ/4 Lemma 8.22 (From local to global). Suppose µG(A) = κ, g1, g2 ∈ StabG (A), and χ : H → T, I ⊆ T, ν are as in Theorem 8.20. Then there is a σ-compact A00 ⊆ A with 00 µG/H (πA ) = 96µG/H (πA)/100 such that for all a ∈ A00H, the following holds

(i) there are ζ −1 , ζ −1 ∈ such that for i ∈ {1, 2}; g1 a g2 a T −1  νκ µH (A ∩ aH) 4 aχ (ζg−1a + I) < . i µG/H (πA)

(ii) with any ζ −1 , ζ −1 satisfying (i) and δa,A(g1, g2) = ζ −1 − ζ −1 , we have g1 a g2 a g2 a g1 a  dA(g1, g2) ∈ µG/H (πA)|δa,A(g1, g2)| + I 20νκ . MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 73

0 0 Proof. Obtain A ,I,J as in Theorem 8.20. Let A1 ⊆ G be the σ-compact set −1 −1 0 {a ∈ A : g1 a, g2 a ∈ A H}. 0 0 It is easy to see that µG/H (πA1) > 98/100µG/H (πA). Fix a ∈ A1, by Theorem 8.20 again we then have

−1 −1 −1  κ µH (A ∩ g1 aH) 4 g1 aχ (ζg−1a + I) < ν , 1 µG/H (πA) and −1 −1 −1  κ µH (A ∩ g2 aH) 4 g2 aχ (ζg−1a + I) < ν . 2 µG/H (πA) 00 0 Multiplying by g1 and g2 respectively, we get (i) when A ⊆ A1. κ/4 As g1, g2 ∈ StabG (A), we have µG(g1A ∩ g2A) > 0. Note that dG(g1A ∩ g2A, B) ≤ 2ηκ by Lemma 5.2. By Theorem 8.20(i), we have 1 1 (26) µ ((g A ∩ g A)H) ≥ µ (HB) ≥ µ (AH). G 1 2 1 + 2η G (1 + 2η)(1 + η) G 0 0 Also Theorem 8.20(ii) implies that there is A2 ⊆ g1A ∩ g2A with µG/H (πA2) > 0 99µG/H (π(g1A ∩ g2A))/100, such that for all a ∈ A2H, we have

 µG(g1A) µG(g1A) (27) 1 − ν ≤ µH (g1A ∩ aH) ≤ (1 + ν) µG/H (π(g1A)) µG/H (π(g1A)) and

µG(g1A ∩ g2A) µG(g1A ∩ g2A) (1 − 2ν) ≤ µH (g1A ∩ g2A ∩ aH) ≤ (1 + 2ν) . µG/H (π(g1A ∩ g2A)) µG/H (π(g1A ∩ g2A))

Note that π(g1A ∩ g2A) ⊆ π(g1A) ∩ π(g2A). However, (26) together with Corol- lary 8.21 give us

µG/H (π(g1A) 4 πA) ≤ 10ηµG/H (πA),

µG/H (π(g1A ∩ g2A) 4 πA) ≤ 14ηµG/H (πA). 00 0 0 00 Hence, Let A = A1H ∩ A2H ∩ A. Note that µG/H (πA ) ≥ 96µG/H (πA)/100, and for all a ∈ A00H, by (26)

µG(g1A ∩ g2A) µG(g1A ∩ g2A) (28) (1 − 5ν) ≤ µH (g1A ∩ g2A ∩ aH) ≤ (1 + 5ν) . µG/H (π(A)) µG/H (π(A))

Finally, recall that dA(g1, g2) = µG(g1A) − µG(g1A ∩ g2A). Note that

µG/H (πA)|δa,A(g1, g2)| = µG/H (πA)|ζ −1 − ζ −1 | g1 a g2 a ∈ µH (g1A ∩ aH) − µH (g1A ∩ g2A ∩ aH) + I(2νκ). MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 74

Hence, by (27) and (28), µ (g A ∩ aH) µ (g A ∩ g A ∩ aH) d (g , g ) ≥ µ (πA) H 1 − H 1 2 A 1 2 G/H 1 + ν 1 − 5ν  ≥ µG/H (πA) µH (g1A ∩ aH) − µH (g1A ∩ g2A ∩ aH) − 18νκ

≥ µG/H (πA)|δa,A(g1, g2)| − 20νκ.

The upper bound on dA(g1, g2) can be computed using a similar method, and this finishes the proof. 

We now deduce properties of the pseudometric dA. Besides the almost linearity, we also need the path monotonicity of the pseudometric to control the errors in the almost homomorphism obtained in Section 7. Proposition 8.23 (Almost linearity and path monotonicity of the pseudometric). Assume that µG(A) = κ, and let ν be as in Theorem 8.20. Then we have the follow- ing: κ/2 (i) For all g1, g2, g3 in StabG (A), we have  dA(g1, g2) ∈ | ± dA(g1, g3) ± dA(g2, g3)| + I 60νκ , (ii) Let g be the Lie algebra of G, and let exp : g → G be the exponential map. For every X ∈ g, either

dA(exp(Xt), idG) < κ/4 for all t ∈ R

or there is t0 > 0 with dA(exp(Xt0), idG) ≥ κ/4 such that for every t ∈ [0, t0],

dA(exp(X(t + t0)), idG)  ∈ dA(exp(X(t + t0)), exp(Xt0)) + dA(exp(Xt0), idG) + I 180νκ . Proof. We first prove (i). Let χ and I be as in Theorem 8.20. Applying Lemma 8.22, we get a ∈ AH and ζ −1 , ζ −1 , ζ −1 ∈ such that for i ∈ {1, 2, 3}, we have g1 a g2 a g1 a T −1  νκ µH (A ∩ aH) 4 aχ (ζg−1a + I) < , i µG/H (πA) and for i, j ∈ {1, 2, 3}, we have  dA(gi, gj) ∈ µG/H (πA)|δa,A(gi, gj)| + I 20νκ . with δa,A(gi, gj) = ζ −1 − ζ −1 . As δa,A(g1, g2) = δa,A(g1, g3) + δa,A(g3, g2), we get gj a gi a the desired conclusion. Next, we prove (ii). Let X ∈ g, and suppose there is t > 0 such that

dA(exp(Xt), idG) ≥ κ. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 75

Using the continuity of g 7→ µG(A \ gA) (Fact 3.1(vii)), we obtain t0 > 0 such that t0 the smallest positive real number with dA(idG, exp(Xt0)) ≥ κ/10. Fix t ∈ [0, t0], and set

g0 = exp(Xt0) and g = exp(Xt).

Note that gg0 = g0g as g0 and g are on the same one parameter subgroup of G. κ/2 One can easily check that g0, g, g0g are in StabG (A). Again, let χ and I be as in Theorem 8.20 and apply Lemma 8.22 to get a ∈ AH and ζ −1 ∈ for gi ∈ gi a T {idG, g, g0, gg0} such that

−1  νκ (29) µH (A ∩ aH) 4 (aχ (ζg−1a + I) < , i µG/H (πA) and for gi, gj ∈ {idG, g, g0, gg0}, we have  (30) dA(gi, gj) ∈ µG/H (πA)|δa,A(gi, gj)| + I 20νκ

As gg0 = g0g, we have

(31) δa,A(idG, g) + δa,A(g, gg0) = δa,A(idG, g0g) = δa,A(idG, g0) + δa,A(g0, g0g)

Using (29), (30), and the fact that dA(idG, g0) = dA(g, gg0), we get  δa,A(g, gg0) ∈ ±δa,A(idG, g0) + I 60νκ .  By a similar argument, δa,A(g0, g0g) ∈ ±δa,A(idG, g)+I 60νκ . Combining with (31), we get that δa,A(idG, ggi) is in both  δa,A(idG, g0) ± δa,A(idG, g) + I 60νκ and  δa,A(idG, g) ± δa,A(idG, g0) + I 60νκ . Using the fact that ν < 10−6, and considering all the four possibilities, we deduce  δa,A(idG, gg0) = δa,A(idG, g0) + δa,A(idG, g) + I 120νκ .

Applying (29) and (30) again, we get the desired conclusion. 

9. Proof of the main theorems In this section, we put everything together to prove some slight generalizations of Theorem 1.1 and Theorem 1.2 as well as Theorem 1.3. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 76

9.1. Minimal expansions in noncompact groups. The next theorem is a re- statement of Theorem 1.1(vi), which is the main result in this subsection. Theorem 9.1 (Main theorem for noncompact groups). Suppose G is a connected unimodular noncompact group, µeG is the inner measure corresponding to a Haar measure of G, the sets A, B ⊆ G have 0 < µeG(A), µeG(B) < ∞, and µeG(AB) = µeG(A) + µeG(B). Then there is a continuous surjective group homomorphism χ : G → R with compact kernel, and compact intervals I,J ⊆ R with µR(I) = µeG(A) and µR(J) = µeG(B), such that A ⊆ χ−1(I), and B ⊆ χ−1(J). If A and B are, moreover, compact, then we also have A = χ−1(I) and B = χ−1(J). Proof. We first treat the case where both A and B are σ-compact. In this case, we can simply use the Haar measure µG instead of the inner measure µe(G). It is easy to see that the last assertion of this theorem is an immediate consequence of the conclusion for this special case. By the Gleason–Yamabe Theorem (Fact 6.7), there is a connected compact normal subgroup H of G such that L = G/H is a connected Lie group. By Fact 3.7, L is unimodular. Let π : G → L be the quotient map. Using Corollary 6.6, there are σ-compact subsets A0,B0 of L such that −1 0 −1 0 µG(A 4 π (A )) = 0 and µG(B 4 π (B )) = 0, 0 0 0 0 and we still have µL(A B ) = µL(A ) + µL(B ). When L is a simple Lie group, by the Iwasawa decomposition, L = KAN where AN is a simply connected closed nilpotent group. Thus AN contains R as a closed subgroup, and so does L. When L is not simple, then L contains a connected closed normal subgroup H, and by Fact 3.7, H is unimodular, and of smaller dimension. Applying induction on dimension, we may assume H satisfies the statement of the theorem. For g1, g2 ∈ L, let

dA(g1, g2) = µL(A) − µL(g1A ∩ g2A).

Then by Proposition 7.1, dA is a pseudometric on L, with radius µG(A). By Propo- sition 8.19, dA is locally linear. Using Proposition 7.6, we have ker dA is a compact normal subgroup of L, and L/ ker dA is isomorphic to R as topological groups. Note that ker dA is the same as its identity component (ker dA)0, because L/(ker dA)0 is a cover group of R, which must be R because it is already simply connected. By the third isomorphism theorem (Fact 3.4), R is a quotient group of G, and the corre- sponding quotient map χ has a connected compact kernel. Applying Corollary 6.6 MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 77 again, as well as the inverse theorem on R (Fact 3.12), there are I,J compact intervals of R such that −1 −1 µG(A 4 χ (I)) = 0 and µG(B 4 χ (J)) = 0.

This also implies that µG(A) = µR(I) and µG(B) = µR(J). Suppose g ∈ A and g∈ / χ−1(I). Since I is compact, χ(g) ∈/ I, and there is α > 0 such that the distance between χ(g) and the nearest element in I is at least α. Thus

µR(χ(g)χ(B) \ χ(A)χ(B)) ≥ α. −1 −1 and this implies that µG(gB \ χ (I)χ (J)) ≥ α. Therefore, −1 −1 −1 −1 µG(AB) ≥ µG(χ (I)χ (J)) + µG(gB \ χ (I)χ (J)) ≥ µG(A) + µG(B) + α, and this contradicts the fact that (A, B) is minimally expanding. Hence, we have A ⊆ χ−1(I) and B ⊆ χ−1(J) as desired. Now we treat the general case without the assumption that A and B are σ-compact. By inner regularity, we obtain σ-compact Ae ⊆ A and Be ⊆ B such that

µG(Ae) = µeG(A) and µG(Be) = µeG(B). Then AeBe is a σ-compact subset of AB. From µeG(AB) = µeG(A) + µeG(B), we obtain

µG(AeBe) = µG(Ae) + µG(Be). Obtain χ, I, and J for Ae and Be as in our earlier proven special case. Argue as in −1 −1 the preceding paragraph, we get A ⊆ χ (I) and B ⊆ χ (I) as desired.  We remark that the same argument almost works for compact groups when (A, B) is a minimal expansion pair, except that when we choose the closed subgroup H, we need to choose one such that we are in the toric transversal scenario. This can be done by using Theorem 8.12 (See Section 9.2). 9.2. Nearly minimal expansions in compact groups. In this subsection, we prove the main theorems for compact groups. We first prove Theorem 1.3.

Proof of Theorem 1.3. Let d > 0 be an integer, let cB be the real number fixed at the beginning of Section 8.3, let ν be in Theorem 8.20. Let G be a connected compact simple Lie group of dimension at most d, and A is a compact subset of G 2 of measure at most S, and µG(A ) < (2 + η)µG(A), where S = S(d) is the constant in Theorem 8.12 (with M = 3), and η is the constant fixed in Section 8.3. By Theorem 8.12, when A satisfies µG(A) < S, there is a closed subgroup H which is isomorphic to T, such that for all g ∈ G, µH (gA ∩ H) < cB. We fix such a closed subgroup H of G. Let

dA(g1, g2) = µG(A) − µG(g1A ∩ g2A). MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 78

2 By Proposition 7.1, dA is a pseudometric. Since µG(A ) < (2 + η)µG(A), Proposi- tion 8.23 shows that dA is a 60νµG(A)-linear pseudometric, and it is 180νµG(A)- path-monotone. By Proposition 7.8, dA is globally 1620νµG(A)-monotone. Let γ = 1620νµG(A). Then dA is γ-monotone γ-linear, and of radius ρ = µG(A)/2. As ν ≤ 10−10, 106γ < ρ, and Theorem 7.25 thus implies that there is a continuous surjective group homomorphism mapping G to T, and this contradicts the fact that G is simple.  Now we are going to prove the inverse theorem. In Proposition 6.15, given a continuous surjective group homomorphism from G to T, we obtain a structural characterization with further assumption that the images of both A and B are small. The next lemma says that, with the homomorphism obtained from almost linear pseudometric in Sections 7 and 8, both A and B should have small image.

Lemma 9.2. Suppose G is a connected compact groups, χ : G → T is a continuous and surjective group homomorphism with connected kernel, A, B ⊆ G are nonempty and σ-compact with

dG(A, B) < µG(A) + µG(B) < max{µG(A), µG(B)} < 1/250.

Suppose for every g ∈ ker(χ) with µG(A \ gA) < µG(A)/36, we further have

µG(A \ gA) < µG(A)/72.

Then µT(χ(A)) + µT(χ(B)) < 1/5.

Proof. Set H = ker(χ). We first show that supg µH (A ∩ gH) > 1/2. Suppose to the contrary that supg µH (A ∩ gH) ≤ 1/2. Note that µ (AH) 1 G > . µG(A) supg µH (A ∩ gH)

Hence by Lemma 8.6, for every ` > 1/2, there is h ∈ H such that µG(A ∩ hA) = `µG(A), and in particular, there is h ∈ H with 35µ (A) 71µ (A) G < µ (A ∩ hA) < G , 36 G 72 which contradicts the assumption.

Now apply Lemma 8.2, we get µT(χA) + µT(χB) ≤ 50(µG(A) + µG(B)) = 1/5 as desired.  With all tools in hand, we are going to prove the following theorem, which is a restatement of Theorem 1.2 and Theorem 1.1(v) for compact groups. MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 79

Theorem 9.3 (Main theorem for compact groups). Let G be a connected compact group, and A, B be compact subsets of G with

0 < λ = min{µG(A), µG(B), 1 − µG(A) − µG(B)}. There is a constant K = K(λ), not depending on G, such that for any 0 ≤ ε < 1, whenever we have δ ≤ Kε and

µeG(AB) < µeG(A) + µeG(B) + δ min{µeG(A), µeG(B)}, there is a surjective continuous group homomorphism χ : G → T together with two compact intervals I,J ∈ T with

µT(I) < (1 + ε)µG(A), µT(J) < (1 + ε)µG(B), and A ⊆ χ−1(I), B ⊆ χ−1(J). If A, B are, moreover, compact with µG(AB) = µG(A) + µG(B), then we also have A = χ−1(I) and B = χ−1(J). Proof. The latter part of the theorem follows immediately from the former part by the fact that open subsets of G has positive µG-measure. We now prove the former part of the theorem. Consider first the case ε > 0. Let δ > 0 to be determined later. Suppose µG(A) ≤ µG(B), and

edG(A, B) = µeG(AB) − µeG(A) − µeG(A) = δµeG(A). By the inner regularity of µeG, we obtain σ-compact A0 ⊆ A and B0 ⊆ B such that µG(A0) = µeG(A) and µG(B0) = µeG(B). As A0B0 is a σ-compact subset of AB, we have µG(A0B0) ≤ µeG(AB) and

dG(A0,B0) ≤ edG(A, B) = δµG(A0). Let τ be the constant from Proposition 6.11. Using Lemma 5.3, we obtain a constant K1 = Oλ(1) and σ-compact A1,B1 ⊆ G such that µG(A1) = µG(B1) < τ,

dG(A1,B1) < K1δµG(A0), and both dG(A1,B) and dG(A, B1) are at most K1δµG(A). Applying Proposition 6.11, we get a connected compact subgroup H of G with G0 = G/H a Lie group of dimen- 0 0 0 0 0 sion O(1) and σ-compact A1,B1 ⊆ G , such that µL(A1) ≤ µL(B1), 0 0 dG0 (A1,B1) < 7K1δµG(A0), and with π : G → G0 the quotient map, we have −1 0 −1 0 (32) max{µG0 (A1 4 π (A1)), µG(B1 4 π (B1))} < 3K1δµG(A0). MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 80

Let S be the constant from Theorem 8.12 with M = 3. We apply Lemma 5.3 0 0 0 again to get a constant K2 and σ-compact A2,B2 ⊆ G , such that 0 0 µG0 (A2) = µG0 (B2) = S, 0 0 0 0 0 0 and all of dG0 (A2,B2), dG0 (A1,B2), and dG0 (A2,B1) are at most K1K2δµG(A0). Let cB be the constant fixed in the beginning of Section 8.3. By Theorem 8.12, we obtain a one-dimensional torus subgroup H0 of G0, such that for every g0 ∈ G0 0 0 0 0 0 0 max{µH0 (A2 ∩ g H ), µT (B2 ∩ H g )} < cB. Similarly as what we did in the proof of Theorem 1.3, we define the pseudometric d : G0 × G0 → R by 0 0 0 d(g1, g2) = min{S/2, µG0 (A2) − µG0 (g1A2 ∩ g2A2)}, this is indeed a pseudometric by Proposition 7.1. Suppose K1K2δµG(A) < ηS where η is the fix constant from Section 8.3. By Proposition 8.23 and Proposition 7.8, we obtain a γ-linear γ-monotone pseudometric, where γ = 1620νS, and ν is from Theorem 8.20. Therefore, Theorem 7.25 gives us a surjective continuous group homomorphism φ : G0 → T, such that for every 0 0 0 g ∈ ker φ ∩ N(λ) with λ = ρ/36, we have µG0 (A2 \ gA2) < S/72. By replacing φ 0 0 with the quotient map G → G /(ker(φ)0) if necessary, where (ker(φ)0) is the identity component of ker(φ), we can arrange that φ has connected kernel. Let χ = π ◦ φ. We now determine the structure of A and B. By Lemma 9.2, we 0 0 have µT(φ(A2)) + µT(φ(B2)) < 1/5. Then by Proposition 6.15, there are compact 0 0 intervals I2,J2 in T, such that 0 0 0 µT(I2) = µL(A2) and µT(J2) = µL(B2), and 0 −1 0 0 −1 0 max{µG0 (A2 4 φ (I2)), µL(B2 4 φ (J2))} < K0K1K2δµG(A), where K0 is the constant in Proposition 6.15. Let cG be the constant in Lemma 6.12. Choose δ such that K0K1K2δµG(A) < cG/700. By Lemma 6.12, there are compact 0 0 intervals I1,J1 in T with 0 0 0 0 µT(I1) = µL(A1) and µT(J1) = µL(B1), and 0 −1 0 0 −1 0 max{µL(A1 4 φ (I1)), µL(B1 4 φ (J1))} < 10K0K1K2δµG(A). 0 0 By (32), and slighltly modifying I1,J1, we get compact intervals I1,J1 ⊆ T such that

µT(I1) = µG(A1) and µT(J1) = µG(B1), and −1 −1 max{µG(A1 4 χ (I1)), µG(B1 4 χ (J1))} < 101K0K1K2δµG(A). MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 81

By Lemma 6.12 again, there are intervals I0 and J0 in T, such that

µT(I0) = µG(A0) and µT(J0) = µG(B0), and −1 −1 max{µG(A0 4 χ (I0)), µG(B0 4 χ (J0))} < 1010K0K1K2δµG(A0). Next, by Lemma 6.13, there are compact intervals I,J ⊆ T, with

µT(I) − µG(A) < 10100K0K1K2δµG(A0),

µT(J) − µG(B) < 10100K0K1K2δµG(A0) ≤ 10100K0K1K2δµG(B0), −1 −1 and A ⊆ χ (I), B ⊆ χ (J). Note that all K0, K1, and K2 only depend on λ, then one can take n 1 ηS c o δ ≤ min , , G ε, 10100K0K1K2 K1K2 700K0K1K2 this finishes the proof. Finally, we consider the case when ε = 0, that is, µG(AB) = µG(A) + µG(B). The proof follows the same argument, by replacing Proposition 8.23 with Proposition 8.19 to construct the locally linear pseudometric, and by replacing Proposition 8.23 and Theorem 7.25 by Proposition 7.6. 

Acknowledgements The authors would like to thank Lou van den Dries, Arturo Rodriguez Fanlo, Kyle Gannon, John Griesmer, Daniel Hoffmann, Ehud Hrushovski, Anand Pillay, Pierre Perruchaud, Daniel Studenmund, Jun Su, Jinhe Ye, and Ruixiang Zhang for valuable discussions. Part of the work was carried out while the first author was visiting the second author at the Department of Mathematics of University of Notre Dame, and he would like to thank the department for the hospitality.

References [BdS16] Yves Benoist and Nicolas de Saxc´e, A spectral gap theorem in simple Lie groups, Invent. Math. 205 (2016), no. 2, 337–361. MR 3529116 [BF19] Michael Bj¨orklund and Alexander Fish, Approximate invariance for ergodic actions of amenable groups, Discrete Anal. (2019), Paper No. 6, 56. MR 3964142 [BG08] Jean Bourgain and Alex Gamburd, On the spectral gap for finitely-generated subgroups of SU(2), Invent. Math. 171 (2008), no. 1, 83–121. MR 2358056 [BG12] , A spectral gap theorem in SU(d), J. Eur. Math. Soc. (JEMS) 14 (2012), no. 5, 1455–1511. MR 2966656 [BGS10] Jean Bourgain, Alex Gamburd, and Peter Sarnak, Affine linear sieve, expanders, and sum-product, Invent. Math. 179 (2010), no. 3, 559–644. MR 2587341 [BGT12] Emmanuel Breuillard, Ben Green, and Terence Tao, The structure of approximate groups, Publ. Math. Inst. Hautes Etudes´ Sci. 116 (2012), 115–221. MR 3090256 MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 82

[BH18] Michael Bj¨orklundand Tobias Hartnick, Approximate lattices, Duke Math. J. 167 (2018), no. 15, 2903–2964. MR 3865655 [BIG17] R´emi Boutonnet, Adrian Ioana, and Alireza Salehi Golsefidy, Local spectral gap in simple Lie groups and applications, Invent. Math. 208 (2017), no. 3, 715–802. MR 3648974 [Bil98] Yuri Bilu, The (α + 2β)-inequality on a torus, J. London Math. Soc. (2) 57 (1998), no. 3, 513–528. MR 1659821 [Bj¨o17] Michael Bj¨orklund, Small product sets in compact groups, Fund. Math. 238 (2017), no. 1, 1–27. MR 3661726 [BL18] Emmanuel Breuillard and Alexander Lubotzky, Expansion in simple groups, arXiv:1807.03879 (2018). [Bog07] Vladimir I. Bogachev, Measure theory. Vol. II, Springer-Verlag, Berlin, 2007. MR 2267655 [Bre14] Emmanuel Breuillard, Geometry of locally compact groups of polynomial growth and shape of large balls, Groups Geom. Dyn. 8 (2014), no. 3, 669–732. MR 3267520 [BV12] Jean Bourgain and P´eterP. Varj´u, Expansion in SLd(Z/qZ), q arbitrary, Invent. Math. 188 (2012), no. 1, 151–173. MR 2897695 [Car15] Pietro Kreitlon Carolino, The Structure of Locally Compact Approximate Groups, Pro- Quest LLC, Ann Arbor, MI, 2015, Thesis (Ph.D.)–University of California, Los Angeles. MR 3438951 [CDR19] Pablo Candela and Anne De Roton, On sets with small sumset in the circle, Q. J. Math. 70 (2019), no. 1, 49–69. MR 3927843 [Chr12] Michael Christ, Near equality in the Brunn-Minkowski inequality, arXiv:1207.5062 (2012). [CI21] Michael Christ and Marina Iliopoulou, Inequalities of Riesz-Sobolev type for compact connected abelian groups, to appear at Amer. J. Math. (2021). [CP20] Gabriel Conant and Anand Pillay, Approximate subgroups with bounded VC-dimension, arXiv:2004.05666 (2020). [CPT21] Gabriel Conant, Anand Pillay, and Caroline Terry, Structure and regularity for subsets of groups with finite VC-dimension, to appear at J. Eur. Math. Soc. (JEMS) (2021). [DE09] Anton Deitmar and Siegfried Echterhoff, Principles of , Universitext, Springer, New York, 2009. MR 2457798 [DeV13] Matt DeVos, The structure of critical product sets, arXiv:1301.0096 (2013). [DF03] Jean-Marc Deshouillers and Gregory A. Freiman, A step beyond Kneser’s theorem for abelian finite groups, Proc. London Math. Soc. (3) 86 (2003), no. 1, 1–28. MR 1971462 [dS15] Nicolas de Saxc´e, A product theorem in simple Lie groups, Geom. Funct. Anal. 25 (2015), no. 3, 915–941. MR 3361775 [FJ17] Alessio Figalli and David Jerison, Quantitative stability for the Brunn-Minkowski inequal- ity, Adv. Math. 314 (2017), 1–47. MR 3658711 [Fre73] Gregory A. Fre˘ıman, Foundations of a structural theory of set addition, American Math- ematical Society, Providence, R. I., 1973, Translated from the Russian, Translations of Mathematical Monographs, Vol 37. MR 0360496 [GKR74] Karsten Grove, Hermann Karcher, and Ernst A. Ruh, Group actions and curvature, Invent. Math. 23 (1974), 31–48. MR 385750 [Gle52] Andrew M. Gleason, Groups without small subgroups, Ann. of Math. (2) 56 (1952), 193– 212. MR 49203 [GR06] Ben Green and Imre Z. Ruzsa, Sets with small sumset and rectification, Bull. London Math. Soc. 38 (2006), no. 1, 43–52. MR 2201602 MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 83

[Gri14] John T. Griesmer, An inverse theorem: when the measure of the sumset is the sum of the measures in a locally compact abelian group, Trans. Amer. Math. Soc. 366 (2014), no. 4, 1797–1827. MR 3152713 [Gri19] , Semicontinuity of structure for small sumsets in compact abelian groups, Discrete Anal. (2019), Paper No. 18, 46. MR 4042161 [Gro81] Mikhael Gromov, Groups of polynomial growth and expanding maps, Publ. Math. Inst. Hautes Etudes´ Sci. (1981), no. 53, 53–73. MR 623534 [Gro03] , Isoperimetry of waists and concentration of maps, Geom. Funct. Anal. 13 (2003), no. 1, 178–215. MR 1978494 [Gry13] David J. Grynkiewicz, Structural additive theory, Developments in Mathematics, vol. 30, Springer, Cham, 2013. MR 3097619 [Hel08] Harald Helfgott, Growth and generation in SL2(Z/pZ), Ann. of Math. (2) 167 (2008), no. 2, 601–623. MR 2415382 [HM53] Ralph Henstock and Murray Macbeath, On the measure of sum-sets. I. The theorems of Brunn, Minkowski, and Lusternik, Proc. London Math. Soc. (3) 3 (1953), 182–194. MR 56669 [HN12] Joachim Hilgert and Karl-Hermann Neeb, Structure and geometry of Lie groups, Springer Monographs in Mathematics, Springer, New York, 2012. MR 3025417 [Hru12] Ehud Hrushovski, Stable group theory and approximate subgroups, J. Amer. Math. Soc. 25 (2012), no. 1, 189–243. MR 2833482 [Hru20] , Beyond the lascar group, arXiv:2011.12009 (2020). [Kaz82] David Kazhdan, On ε-representations, Israel J. Math. 43 (1982), no. 4, 315–323. MR 693352 [Kec95] Alexander S. Kechris, Classical descriptive set theory, Graduate Texts in Mathematics, vol. 156, Springer-Verlag, New York, 1995. MR 1321597 [Kem60] Johannes Kemperman, On small sumsets in an abelian group, Acta Math. 103 (1960), 63–88. MR 110747 [Kem64] , On products of sets in a locally compact group, Fund. Math. 56 (1964), 51–68. MR 202913 [Kne56] Martin Kneser, Summenmengen in lokalkompakten abelschen Gruppen, Math. Z. 66 (1956), 88–110. MR 81438 [Lev20] Vsevolod Lev, Small doubling in cyclic groups, arXiv:2010.03410 (2020). [Lyu35] Lazar’ A. Lyusternik, Die Brunn–Minkowskische ungleichnung f¨urbeliebige messbare mengen, Comptes Rendus de l’Acad´emiedes Sciences de l’URSS. Nouvelle S´erie.III (1935), 55–58. [Mac53] Alexander M. Macbeath, On measure of sum sets. II. The sum-theorem for the torus, Proc. Cambridge Philos. Soc. 49 (1953), 40–43. MR 56670 [McC69] Michael McCrudden, On the Brunn-Minkowski coefficient of a locally compact unimodular group, Proc. Cambridge Philos. Soc. 65 (1969), 33–45. MR 233921 [MP15] Brendan Murphy and Jonathan Pakianathan, Kakeya configurations in Lie groups and homogeneous spaces, Topology Appl. 180 (2015), 1–15. MR 3293263 [MW15] Jean-Cyrille Massicot and Frank O. Wagner, Approximate subgroups, J. Ec.´ polytech. Math. 2 (2015), 55–64. MR 3345797 [MZ52] Deane Montgomery and Leo Zippin, Small subgroups of finite-dimensional groups, Ann. of Math. (2) 56 (1952), 213–241. MR 49204 MINIMAL AND NEARLY MINIMAL MEASURE EXPANSIONS 84

[PS16] L´aszl´oPyber and Endre Szab´o, Growth in finite simple groups of Lie type, J. Amer. Math. Soc. 29 (2016), no. 1, 95–146. MR 3402696 [Rai39] Dmitrii Raikov, On the addition of point-sets in the sense of Schnirelmann, Rec. Math. [Mat. Sbornik] N.S. 5(47) (1939), 425–440. MR 0001776 [Ros19] Christian Rosendal, Continuity of universally measurable homomorphisms, Forum Math. Pi 7 (2019), e5, 20. MR 3996719 [Shi55] Allen Shields, Sur la mesure d’une somme vectorielle, Fund. Math. 42 (1955), 57–60. MR 72201 [Tao08] Terence Tao, Product set estimates for non-commutative groups, Combinatorica 28 (2008), no. 5, 547–594. MR 2501249 [Tao11] , The Brunn-Minkowski inequality for nilpotent groups, available at https://terrytao.wordpress.com/2011/09/16/ (2011). [Tao15] , Expansion in finite simple groups of Lie type, Graduate Studies in Mathematics, vol. 164, American Mathematical Society, Providence, RI, 2015. MR 3309986 [Tao18] , An inverse theorem for an inequality of Kneser, Proc. Steklov Inst. Math. 303 (2018), no. 1, 193–219, Published in Russian in Tr. Mat. Inst. Steklova 303 (2018), 209– 238. MR 3920221 [Vos56] A. G. Vosper, The critical pairs of subsets of a group of prime order, J. London Math. Soc. 31 (1956), 200–205. MR 77555 [vZ04] J´an Spakulaˇ and Pavol Zlatoˇs, Almost homomorphisms of compact groups, Illinois J. Math. 48 (2004), no. 4, 1183–1189. MR 2113671 [Wei40] Andr´eWeil, L’int´egration dans les groupes topologiques et ses applications, Actual. Sci. Ind., no. 869, Hermann et Cie., Paris, 1940, [This book has been republished by the author at Princeton, N. J., 1941.]. MR 0005741 [Yam53] Hidehiko Yamabe, A generalization of a theorem of Gleason, Ann. of Math. (2) 58 (1953), 351–365. MR 58607

Department of Mathematics, University of Illinois Urbana-Champaign, Urbana IL, USA Email address: [email protected]

Department of Mathematics, University of Notre Dame, Notre Dame IN, USA Email address: [email protected]