Dp-minimal profinite groups and
valuations on the integers
Abstract
We study dp-minimal infinite profinite groups that are equipped with a uniformly definable fundamental system of open subgroups. We show that these groups have an open subgroup such that either is a direct product of countably many copies of for some prime , or is of the form where and is a finite abelian -group for each prime . Moreover, we show that if is of this form, then there is a fundamental system of open subgroups such that the expansion of by this family of subgroups is dp-minimal. Our main ingredient is a quantifier elimination result for a class of valued abelian groups. We also apply it to and we show that if we expand by any chain of subgroups , we obtain a dp-minimal structure. This structure is distal if and only if the size of the quotients is bounded.
1 Introduction
A profinite group together with a fundamental system of open subgroups can be viewed as a two-sorted structure in the two-sorted language . In these structures the fundamental system of open subgroups is definable. Since a fundamental system of open subgroups is a neighborhood basis at the identity, this implies that the topology on is definable.
These structures have been studied by Macpherson and Tent in [19]. They mainly considered full profinite groups, i.e. profinite groups where the family consists of all open subgroups. Their main result states that a full profinite group is NIP if and only if it is NTP2 if and only if it is virtually a finite direct product of analytic pro- groups.
Since analytic pro- groups can be described as products of copies of with a twisted multiplication, profinite NIP groups are composed of “one-dimensional” profinite NIP groups. In the setting of full profinite groups the combinatorial structure of the lattice of open subgroups is visible in the model theoretic structure. This plays an important role in the classification.
Without the fullness assumption, only a portion of this lattice is visible. In general the family could simply consist of a chain of open subgroups. In this more general setting, we will restrict ourselves to the “one-dimensional”, i.e. dp-minimal case. A profinite group is dp-minimal if it has NIP and is dp-minimal in the group sort. We prove the following classification result:
Theorem.
Let be a dp-minimal profinite group. Then has an open abelian subgroup such that either is a direct product of countably many copies of for some prime , or is isomorphic to where and is a finite abelian -group for each prime . Moreover, every abelian profinite group of the above form admits a fundamental system of open subgroups such that the corresponding -structure is dp-minimal.
The main ingredient of this theorem is a quantifier elimination result which is also applicable in other settings. We apply it to this situation: Consider the structure . If we expand it by the full lattice of subgroups, then the expanded structure interprets Peano Arithmetic and hence is not tame in any sense. However, if we only name a chain in this lattice, we obtain a tame structure. A chain of subgroups is the same as a valuation defined by
Theorem.
Let be a strictly descending chain of subgroups of , , and let be the valuation defined by
Then is dp-minimal. Moreover, is distal if and only if the size of the quotients is bounded.
This stays true if we expand the value sort by unary predicates and monotone binary relations. There has been recent interest in dp-minimal expansions of (e.g. [2], [21], [1], and [22]). Alouf and d’Elbée showed in [2] that if is a prime and denotes the -adic valuation, then is a minimal expansion of in the sense that there are no proper intermediate expansions. We show that this does not hold true for all valuations and we conjecture that the -adic valuations are essentially the only examples with this property among valuations such that is distal.
The proof of the classification theorem for dp-minimal profinite groups consists of three parts: We analyze the algebraic structure of dp-minimal profinite groups in Section 3. This will imply the first part of the theorem. It then remains to show that these groups appear as dp-minimal profinite groups. This is done in Section 4. The case where the group is given by an -vector space has already been done by Maalouf in [10]. We explain this result in Section 4.1. The remaining case is handled by a quantifier elimination result (see Section 4.2). This quantifier elimination result allows us to show that a certain class of profinite groups as -structures is dp-minimal (Theorem 4.29) and we are able to characterize distality in this class (Theorem 4.32).
We will also apply the quantifier elimination result to valuations on . This will be done in Section 5 where we discuss the second theorem and its consequences for the study of dp-minimal expansions of . We also show that the -adic valuations have a limit theory (Proposition 5.9) and we consider expansions of given by multiple valuations.
Section 6 contains a few results which are related to dp-minimal profinite groups. We show that our main result implies some structural consequences for uniformly definable families of finite index subgroups in dp-minimal groups (Proposition 6.4). Jarden and Lubotzky [8] showed that two elementarily equivalent profinite groups are isomorphic if one of them is finitely generated. This was generalized to strongly complete profinite groups by Helbig [7]. We will give an alternative proof for these results in Section 6.2. Finally, we prove a result about uniformly definable families of normal subgroups in NTP2 groups (Proposition 6.13)222Thanks to Pierre Simon for bringing this question to my attention.: If such a family is closed under finite intersections, then it must be defined by an NIP formula.
Acknowledgments
I would like to thank my supervisor, Katrin Tent, for her help and support during the last years and for giving me the opportunity to work on these exciting topics. I would also like to thank Pierre Simon for interesting and useful discussions and valuable suggestions while I visited UC Berkeley.
2 Preliminaries
We assume that the reader is familiar with both profinite groups and model theory. We will give a quick overview about the notions and tools that are used to prove the main result.
2.1 Profinite groups
A topological group is profinite if it is the inverse limit of an inverse system of (discrete) finite groups. This condition is equivalent to the group being Hausdorff, compact, and totally disconnected. If is a profinite group, then
where ranges over all open normal subgroups.
The open subgroups generate the topology on , i.e. every open set is a union of cosets of open subgroups. A fundamental system of open subgroups is a family consisting of open subgroups which generate the topology on . Equivalently, every open subgroup of contains a subgroup in . If is a property of groups, we will say that is virtually if has an open subgroup which satisfies .
We will use a number of results about the structure of abelian profinite groups. Recall that a profinite group is pro- if it is the inverse limit of finite -groups. A free abelian pro- group is a direct product of copies of .
Proposition 2.1 (Theorem 4.3.4 of [13]).
Let be a prime.
-
(a)
If is a torsion free pro- abelian group, then is a free abelian pro- group.
-
(b)
Let be a finitely generated pro- abelian group. Then the torsion subgroup is finite and
where is a free pro- abelian group of finite rank.
Proposition 2.2 (Corollary 4.3.9 of [13]).
Let be a torsion profinite abelian group. Then there is a finite set of primes and a natural number such that
where each is a cardinal and each is the cyclic group of order . In particular, is of finite exponent.
Proposition 2.3 (Proposition 1.13 and Proposition 1.14 of [6]).
Let be a pro- group. Then is (topologically) finitely generated if and only if the Frattini subgroup is open in .
Proposition 2.4.
Let be an abelian profinite group. Then is an open subgroup for all if and only if
where and is a finite abelian -group for each prime .
Proof.
An abelian profinite group is the direct product of its -Sylow subgroups. Let be a -Sylow subgroup of . If has finite index, then is finitely generated by Proposition 2.3. Then by Proposition 2.1 the -Sylow subgroup has the desired form. ∎
We will also need the following result by Zelmanov:
Theorem 2.5 (Theorem 2 of [24]).
Every infinite compact group has an infinite abelian subgroup.
We will view profinite groups as two-sorted structures in the following language which was introduced in [19]:
Definition 2.6.
is a two-sorted language containing the group sort and the index sort . The language then consists of:
-
•
the usual language of groups on ,
-
•
a binary relation on , and
-
•
a binary relation .
Remark 2.7.
A profinite group together with a fundamental system of open subgroups can be viewed as an structure as follows:
-
•
we set if and only if , and
-
•
the relation is defined by for all .
2.2 Model theoretic notions of complexity
We will mostly work in the context of an NIP theory. We use [16] as our main reference for this section.
2.2.1 The independence property
An important class of model theoretic theories is the class of NIP (or dependent) theories, i.e. the class of theories which cannot code the -relation on an infinite set. This notion was introduced by Shelah.
Definition 2.8.
A formula has the independence property (IP) if there are sequences and such that
We say that has NIP if it does not have IP. This notion is symmetric in the sense that a formula has NIP if and only if the formula has NIP (see Lemma 2.5 of [16]).
We will make use of the following characterization of IP:
Lemma 2.9 (Lemma 2.7 of [16]).
A formula has IP if and only if there exists an indiscernible sequence and a tuple such that
We call a theory NIP if all formulas have NIP.
Definition 2.10.
A subset is externally definable if there is a formula , an elementary expansion of , and an element such that .
By a result of Shelah, naming all externally definable sets in an NIP structure preserves NIP:
Theorem 2.11 (Proposition 3.23 and Corollary 3.24 of [16]).
Let be a model of an NIP theory and let be the Shelah expansion, i.e. the expansion of by all externally definable sets. Then has quantifier elimination and is NIP.
Theorem 2.12 (Baldwin-Saxl, Theorem 2.13 of [16]).
Let be an NIP group and let be a family of uniformly definable subgroups of . Then there is a constant such that for any finite subset there is of size such that
As an easy consequence we obtain:
Corollary 2.13.
If is an NIP profinite group, then can only contain finitely many subgroups of any given finite index.
By a result of Shelah, abelian subgroups of NIP groups have definable envelopes given by centralizers of definable sets:
Theorem 2.14 (Proposition 2.27 of [16]).
Let be an NIP group and let be a set of commuting elements. Then there is a formula and a parameter (in some elementary extension ) such that is an abelian (definable) subgroup of and contains .
2.2.2 Dp-minimality
NIP theories admit a notion of dimension given by dp-rank:
Definition 2.15 (Definition 4.2 of [16]).
Let be a partial type over a set , and let be a cardinal. We define
if and only if for every family of mutually indiscernible sequences over and , one of these sequences is indiscernible over .
A theory is called dp-minimal if where is a singleton. We call a multi-sorted theory with distinguished home-sort dp-minimal if it is NIP and it is dp-minimal in the home-sort, i.e. where is a singleton in the home-sort.
Remark 2.16.
As a consequence of the quantifier elimination in Theorem 2.11 the Shelah expansion of a dp-minimal structure is dp-minimal.
We will use the fact that definable subgroups in a dp-minimal group are always comparable in the following sense:
Lemma 2.17 (Claim in Lemma 4.31 of [16]).
Suppose is dp-minimal and and are definable subgroups. Then or is finite.
2.2.3 Distality
Distality is a notion introduced by Simon to describe the unstable part of an NIP theory. The general definition of distality is slightly more complicated than the definitions of NIP and dp-minimality (see Definition 2.1 in [15] or Chapter 9 in [16]). In case of a dp-minimal theory distality can be described as follows:
Proposition 2.18.
A dp-minimal theory is distal if and only if there is no infinite non-constant totally indiscernible set of singletons.
Proof.
This characterization follows from Example 2.4 and Lemma 2.10 in [15]. ∎
By Exercise 9.12 of [16] distality is preserved under going to :
Proposition 2.19.
If is distal, then so is .
2.3 Quantifier elimination
Recall that a theory has quantifier elimination if every formula is equivalent to a quantifier free formula modulo . The proof of Theorem 3.2.5 in [20] gives the following useful criterion for quantifier elimination:
Proposition 2.20.
Let be a theory and let be a formula. Then is equivalent to a quantifier free formula modulo if and only if for all with common substructure and all we have
If is a two-sorted theory and the only symbols that connect the two sorts are functions from one sort to the other, then it suffices to check quantifier elimination for very specific formulas:
Lemma 2.21.
Let be a theory in a two-sorted language with sorts and where is purely in the sort , is purely in the sort , and each is a function from sort to sort . Suppose
-
(a)
every -formula is equivalent to a quantifier free formula modulo and
-
(b)
every formula of the form
is equivalent to a quantifier free formula modulo where is a singleton, , , and each is either a basic -formula or is of the form where is an term and is one of the variables in the tuple .
Then eliminates quantifiers.
Proof.
To show quantifier elimination it suffices to consider simple existential formulas. Consider a formula of the form
where is a singleton, , , and each is a basic formula. We may assume that appears non-trivially in each formula . Then each is a basic -formula where the variables only appear as terms of the form
where is a function symbol and is an -term. Now the -quantifier can be eliminated by (a).
Now consider a formula of the form
where is a singleton, , , and each is a basic formula.
Let be the set of all such that is a basic -formula. If , then we may write as
where is a basic -formula such that all variables of are in . Then is equivalent to
Now we may rewrite
as a formula of the form
We can now eliminate the -quantifier by (b) and then eliminate the -quantifiers as in the first step. ∎
3 Algebraic properties of dp-minimal profinite groups
We view a profinite group together with a fundamental system of open subgroups as an -structure (as in Remark 2.7). The aim of this chapter is to prove the first part of the main theorem: If is a dp-minimal profinite group, then has an open abelian subgroup such that either is a vector space over for some prime , or where and is a finite abelian -group for each prime .
Simon showed in [14] that all dp-minimal groups are abelian-by-finite-exponent. An example of a dp-minimal group that is not abelian-by-finite was given by Simonetta in [17].
We will show that all dp-minimal profinite groups have an open abelian subgroup. We will then analyze the structure of this abelian profinite group. For dp-minimal profinite groups the fundamental system of open subgroups can always be replaced by a chain of open subgroups:
Lemma 3.1.
Let be a dp-minimal profinite group. Then the subgroups
are uniformly definable open subgroups and hence the topology on is generated by a definable chain of open subgroups.
Proof.
The are open subgroups by Corollary 2.13. By Lemma 2.17 and compactness we can find a constant such that for all or . Given we have
Moreover, we have or . Therefore this is a definable condition and hence the subgroups are uniformly definable. ∎
In a dp-minimal profinite group we cannot find infinite definable subgroups of infinite index:
Lemma 3.2.
Let be a dp-minimal profinite group. Let be an elementary extension and let be a definable subgroup. If is infinite, then is finite.
Proof.
If is finite for some , then clearly .
Now assume is infinite for all . We aim to show that must be unbounded: Since is infinite and , must be unbounded. Therefore must be unbounded. This contradicts Lemma 2.17. ∎
As a consequence of Zelmanov’s theorem (Theorem 2.5) and the existence of definable envelopes for abelian subgroups (Theorem 2.14) we get that a dp-minimal profinite group must be virtually abelian:
Proposition 3.3.
Let be a dp-minimal profinite group. Then is virtually abelian.
Proof.
By Theorem 2.5 has an infinite abelian subgroup . By Theorem 2.14, we can find an elementary extension , a formula , and a parameter such that is an abelian subgroup of and contains . Therefore has finite index in by Lemma 3.2. By elementarity there is some such that is an abelian group and has finite index in . Moreover, is closed since it is a centralizer. Closed subgroups of finite index are open and therefore is an open abelian subgroup of . ∎
We are now able to prove the first part of the main theorem:
Theorem 3.4.
Let be an abelian dp-minimal profinite group. Then either is virtually a direct product of countably many copies of for some prime , or where and is a finite abelian -group for each prime .
Proof.
Consider the closed subgroup . Suppose there is a minimal such that is infinite. Then has finite index in (by Lemma 3.2) and hence is an open subgroup of . Therefore we may assume . Now the minimality of and Proposition 2.2 imply that must be prime and therefore is a direct product of copies of (again by Proposition 2.2). Since admits a countable fundamental system of open subgroups, this direct product must be a direct product of countably many copies of .
Now assume is finite for all . Then the closed subgroup must be open in for all (by Lemma 3.2). Now Proposition 2.4 implies the theorem. ∎
4 Valued abelian profinite groups
If is an abelian group and is a strictly descending chain of subgroups such that and , then we can define a valuation map by setting
We have if and only if , and this valuation satisfies the inequality
where we have equality in case .
The valued group can be seen as a two-sorted structure consisting of the group , the linear order , and the valuation .
Our goal is to classify dp-minimal profinite groups up to finite index. We know by Lemma 3.1 that the fundamental system of open subgroups can be assumed to be a chain. Moreover, by Theorem 3.4 we only need to consider groups of the form
where and is a finite abelian -group for each prime .
If is such a group and is a fundamental system of open subgroups which is given by a strictly descending chain, then the above construction yields a definable valuation . Conversely, given such a valuation , we can recover the fundamental system of open subgroups by setting
Hence the valuation and the fundamental system are interdefinable.
We will show that if is of the above form, then admits a fundamental system given by a chain of open subgroups such that the expansion of by the corresponding valuation (and hence the corresponding -structure) is dp-minimal. If , this follows from results by Maalouf in [10] and will be explained in Section 4.1.
Definition 4.1.
-
(a)
The subgroups are called the -balls of radius . We will also denote them by to emphasize that they correspond to the valuation .
-
(b)
A valuation is good if for all the subgroup is of the form for some positive integer .
In case , we will prove a quantifier elimination result for good valuations. Note that by Proposition 2.4 each such group can be equipped with a good valuation such that is a fundamental system of open subgroups. We will show the following theorem:
Theorem 4.2.
Let as above and let be a good valuation. Then the structure is dp-minimal. Moreover, it is distal if and only if the size of the quotients is bounded.
This theorem will be proven in this chapter. If is a set of primes, a natural number is called a -number if the prime decomposition of only contains primes in . An immediate consequence of the above theorem is the following:
Corollary 4.3.
Let be a sequence of finite non-empty disjoint sets of primes. For each fix a finite non-trivial abelian group such that is a -number. Set
and let be the valuation defined by
Then is dp-minimal but not distal.
Proof.
We have . Hence is a good valuation and the theorem applies. ∎
4.1 Valued vector spaces
Valued vector spaces have been studies by S. Kuhlmann and F.-V. Kuhlmann in [9] and by Maalouf in [10]. Set and let be the valuation given by
It follows from results by Maalouf in [10] that this valued abelian profinite group is dp-minimal:
Proposition 4.4.
The valued abelian profinite group is dp-minimal.
Proof.
Set and let be the valuation given by
By Proposition 4 of [10] the valued vector space is C-minimal and hence dp-minimal (by Theorem A.7 of [16]).
Théorème 1 of [10] implies that and are elementarily equivalent. Hence is dp-minimal. ∎
Remark 4.5.
The last step of the previous proof also follows from results in Section 6.1. Let be as in the proof of Proposition 4.4 and set
Then and hence is dp-minimal by Lemma 6.2.
4.2 A quantifier elimination result
We denote the set of primes by . For each prime we fix an integer and a finite -group . Let
be an abelian group that is (as a pure group) an elementary substructure. We will always assume to be infinite. We fix a set of constants containing such that the set is dense with respect to the profinite topology on and contains every torsion element. It follows from Proposition 2.4 that the set of constants is also dense with respect to the profinite topology on .
Definition 4.6.
If is a set of primes, we set
Note that we have for any set . The group is the -torsion part of and the group has no -torsion.
Let be a good valuation and set . For each we define a function by
Note that if , then . Now together with the valuation may be viewed as a two-sorted structure with group sort and value sort in the language where
-
•
is the obvious language on ,
-
•
consists of symbols for the functions , and
-
•
is the obvious language on .
Since we consider the group and the constants to be fixed, this structure only depends on the valuation and we denote it by .
We define the following binary relations on :
-
•
and ,
-
•
and divides ,
where is a finite set of primes, is a prime, and . We set and to be always true.
Observation 4.7.
-
(a)
If then divides if and only if has an element of order . In particular, the predicate is definable.
-
(b)
In the standard model is equivalent to the statement that divides . In that sense the expression makes sense even in non-standard models.
-
(c)
We have if and only if . Hence the subgroups are quantifier free -definable. Since the subgroups generate the profinite topology on , this implies that the open subgroups are quantifier free 0-definable for finite subsets . Moreover, in that case is also quantifier free 0-definable since it is a finite set of constants.
Let be the set of good valuations on . We set to be the common -Theory of structures , . The following quantifier elimination result will be shown in the next sections:
Theorem 4.8.
Let be an expansion on the sort and let be an expansion of to the language . Suppose that:
-
1.
The relations and are quantifier free 0-definable modulo .
-
2.
The successor function on is contained in .
-
3.
Every -formula is equivalent to a quantifier free -formula modulo .
Then eliminates quantifiers.
To prove the quantifier elimination result we will need to understand formulas that describe systems of linear congruences. Therefore we will need to understand linear congruences in models of the theory .
4.2.1 Linear congruences in
We will need generalizations of the following well-known fact:
Fact 4.9.
A linear congruence in has a solution if and only if divides . In that case it has exactly many solutions modulo . If is a solution, then a complete system of solutions modulo is given by
Observation 4.10.
4.9 has two important consequences:
-
(a)
If has a solution and , then divides and hence is a solution.
-
(b)
If has a solution, then all solutions agree modulo .
Part (a) will be important since in that case a solution will be determined by the constant . Part (b) tells us that solutions of linear congruences can “collapse”. We will need to understand this collapsing of solutions.
We now fix a group of the form , . If is a positive integer, let be the unique integer such that . Note that 4.9 can be applied to because .
We consider linear congruences
in where and are positive integers and . Note that solving the above linear congruence is equivalent to solving it in each copy of in the product :
Lemma 4.11.
-
(a)
Let be a linear congruence in . Write
The solutions for in are exactly the tuples where each is a solution for in .
-
(b)
Set . Then the linear congruence has a solution if and only if divides in (i.e. ). In that case it has exactly many solutions modulo in .
We call a finite family of linear congruences (and negations of linear congruences) a system of linear congruences. Recall Bézout’s identity:
Fact 4.12 (Bézout’s identity).
If are integers, then is a -linear combination of .
We will look at systems of linear congruences where the modulus is fixed:
Proposition 4.13.
Let be a system of linear congruences in (where is a finite set). Set and . By Bézout’s identity we can find integers such that . Put .
-
(a)
If the system has a common solution, the solutions of are exactly the solutions of .
-
(b)
Set and . Then the system has a solution if and only if the system :
has a solution. Moreover, the systems and have the same number of solutions modulo .
Proof.
(a) It suffices to show this for each factor in the product . Hence we may assume and . Clearly any common solution of the system solves .
Now suppose is a solution of (and hence a solution of ). Then by 4.9 all solutions of are of the form where . Fix . Now divides , say . Therefore
solves for all (by 4.9). Hence every solution of solves .
(b) We have . If we divide by , we get that . We aim to show that has a solution if and only if has a solution. If is a solution for , then solves . Now assume that has a solution. Then by (a) the system has the same solutions as the linear congruence
Since we assume that has a solution, this implies that divides (by part (b) of Lemma 4.11). Then the linear congruence
also has solutions by part (b) of Lemma 4.11 since divides . If solves , then solves and hence is a solution for . This implies that solves . Hence has a solution if and only if has a solution. Moreover, if and have solutions, then by (a) the solutions of are exactly the solutions of and the solutions of are exactly the solutions of . Hence they have the same number of solutions modulo by part (b) of Lemma 4.11. ∎
We will now consider systems of linear congruences where we vary the modulus:
Lemma 4.14.
Let be a linear congruence in . Set and suppose divides such that divides . Then
have exactly solutions modulo respectively and all these solutions agree modulo .
Proof.
Proposition 4.15.
Let be a linear congruence in . Set . Suppose divides and is such that for all we have or does not divide . Set
Then the linear congruences
have the same number of solutions modulo respectively . Moreover, if is the set of solutions modulo of , is the set of solutions modulo of , and and are the images of and in , then and each element in (resp. ) has exactly many preimages in (resp. ).
Proof.
By an application of Lemma 4.11 it suffices to show this in case . Hence we will assume .
If does not divide , then is a unit in and hence each of the congruences has a unique solution in .
Hence we may assume divides . If does not divide , then and . Then and therefore the linear congruences
have the same solutions (in ). Since solutions modulo (resp. modulo ) are the same as solutions modulo , each element of (resp. ) has a unique preimage in (resp. ).
Now assume divides . Then by assumption divides . In that case the result follows by Lemma 4.14. Note that each element in (resp. ) has exactly preimages in (resp. ). ∎
4.2.2 Linear congruences in
Fix as in Theorem 4.8 (for a group as in the beginning of Section 4.2). We have if and only if . Therefore we will consider certain formulas as linear congruences:
Here will be a variable and will be a constant. The integer will be part of the formula. In particular, it will always be a standard integer. Recall that for a subset a natural number is called a -number if the prime decomposition of only contains primes in .
We will often work in the -torsion free group defined in Definition 4.6. If we assume that is finite, then by part (c) of 4.7 the subgroup is quantifier free 0-definable. If is any model, then we set to be the subgroup defined by the formula which defines in . The subgroup is defined analogously.
If we use the notation in part (b) of 4.7, then
is well-defined even if has infinite index because is always a standard integer. Therefore the results in Section 4.2.1 can be formulated using the divisibility predicates and they will hold true for models of .
Proposition 4.16.
Let be a model of and let be a linear congruence in . Let be a finite set of primes such that is a -number and . Then the linear congruence has a solution in if and only if divides (i.e. ). In that case there are exactly many solutions modulo in .
Proof.
This is essentially part (b) of Lemma 4.11. Since this is a first-order statement, it suffices to consider good valuations on . Since the statement only affects the quotients , we may assume that is of the form
Put and . Then
and has no -torsion. Write for and . Then we can apply Lemma 4.11 to the linear congruence
in . Note that the linear congruence
has a unique solution modulo in (namely ). This shows the proposition. ∎
Proposition 4.17.
Let be a model of and let be a system of linear congruences. Let be a finite set of primes such that all are -numbers and all are contained in . Set and . By Bézout’s identity we can find integers such that . Put .
-
(a)
If the system has a common solution in , the solutions of the system in are exactly the solutions of in .
-
(b)
Set and . Then the system has a solution in if and only if the system
has a solution in . Moreover, these systems have the same number of solutions modulo in .
Proof.
This follows from Proposition 4.13 by the same arguments that are used in Proposition 4.16. ∎
Proposition 4.18.
Let be a model of and let be a linear congruence. Let be a finite set of primes such that is a -number and . Set . Fix and such that is a subgroup and is such that for all we have that divides or does not divide . Set
Then the linear congruences
have the same number of solutions modulo respectively in . Moreover, if is the set of solutions modulo of , is the set of solutions modulo of , and and are the images of and modulo , then and each element in (resp. ) has exactly many preimages in (resp. ).
Proof.
This follows from Proposition 4.15 by the same arguments that are used in Proposition 4.16. ∎
The following lemma will often be useful:
Lemma 4.19.
Fix a model , let be a finite set of primes, let , and let be a -number. Then the linear congruences
have the same solutions in .
Proof.
Multiplying by is injective since does not have -torsion. ∎
4.2.3 Systems of linear congruences in
Fix as in Theorem 4.8. Note that we assume that the successor function (on ) is contained in .
Lemma 4.20.
Let and be models of and let be a common substructure. Let be a finite set of primes and let
be a system of linear congruences where each is a -number, , and . Suppose there is a -number and a constant such that divides and solves in . Then solves in .
Proof.
We have divides if and only if . This does not depend on the model. Moreover, solves if and only if . By Lemma 4.19 we have
Therefore this value does not depend on the model. ∎
Lemma 4.21.
Let and be models of and let be a common substructure. Let be a finite set of primes and let
be a system of linear congruences where each is a -number, , and . Then the system has the same number of solutions modulo in and .
Proof.
Set , say (by Bézout’s identity), and put . Set , , and . Then by Proposition 4.17 (b) the system
has the same number of solutions modulo in (resp. ) as the system
We have By Proposition 4.17 (a) any solution of the system
is a solution of . Now by Proposition 4.16 the linear congruence has a solution if and only if divides . In that case must be a solution and we can apply Lemma 4.20 to see that this must hold true in both models.
Hence contains a solution if and only if contains a solution. In that case the solutions are exactly the solutions of and by Proposition 4.16 the number of solutions modulo does not depend on the model. ∎
Lemma 4.22.
Let and be models of and let be a common substructure. Let be a finite set of primes and let be a system
of linear congruences where and each is a -number, , . Suppose moreover, that the index is a -number whenever it is finite. Fix such that is maximal. Then has the same number of solutions modulo in and .
Proof.
If is finite, then there is a -number such that . Lemma 4.19 allows us to replace the linear congruence
by the linear congruence
Hence we may assume that the index is infinite whenever .
For set and consider the system :
By Lemma 4.21 the system has the same number of solutions modulo in and . If has no solution, then has no solution and we are done. Hence assume that has a solution (and hence has the same number of solutions in both models by Lemma 4.21).
Then by Proposition 4.17 we can replace the system by a single linear congruence without changing the solutions.
Hence we may assume
for all . Now we may write such that . We prove the lemma by induction on . The case is done by Lemma 4.21. Hence we assume .
The system has the form
Now set and put
Set p and consider the system
By Proposition 4.18 the linear congruences and have the same number of solutions modulo respectively and the sets of solutions agree modulo . The statement about the number of preimages in Proposition 4.18 implies that and have the same number of solutions modulo respectively . By Lemma 4.19 we can rewrite as follows:
By induction the system has the same number of solutions modulo in and . Hence has the same number of solutions modulo in and . ∎
To deal with the general case we will make use of the following:
Fact 4.23 (Inclusion-exclusion priciple).
Let be finite sets. Then
Proposition 4.24.
Let and be models of and let be a common substructure. Let be a finite set of primes and let be a system
of linear congruences where each is a -number, , for all . Assume there is such that is maximal in . Suppose moreover, that the index is a -number whenever it is finite. Then has the same number of solutions modulo in and .
Proof.
For pairwise distinct let be the set of solutions modulo of the system :
By Lemma 4.22 the system has the same number of solutions in and . In particular, this holds true for the system :
Moreover, is exactly the set of solutions modulo for that do not solve .
Note that and hence by an application of the inclusion-exclusion principle the number (which is finite since we only count solutions modulo ) does not depend on the model.
Now the system is solved by exactly
many solutions modulo and this number does not depend on the model. ∎
4.2.4 Proof of quantifier elimination
Proof of Theorem 4.8.
By Lemma 2.21 it suffices to show that every formula of the form
is equivalent to a quantifier free formula modulo where , and each is either a basic -formula or is of the form
where is an -term and is one variable in the tuple .
Write such that
Now let be a finite set of primes such that , , and the cardinalities of all finite quotients are -numbers. Fix two models of and let be a common substructure, . Set . We have (since is a finite set of constants) and hence each can be written as with and . Now suppose the formula has a solution in . By Proposition 2.20 it suffices to show that it has a solution in .
If , then must be satisfied in and . If , then it suffices if is satisfied in or . If , then we have
This is satisfied if we have “=” in or and “” in the other subgroup. Hence there are subsets and such that the formulas
and
have a solution in respectively . Since is a finite set of constants, this implies that has a solution in . It remains to show that has a solution in .
If , then we are done since the formulas are quantifier free -definable and hence the result follows from the usual quantifier elimination for abelian groups. Therefore we assume .
If , say , then is the solution of in . Lemma 4.20 implies that also solves in . Hence we may assume .
If for some , then we have
Hence we may assume for all .
Given there is a finite set of constants in the language such that the formula is equivalent to
Thus we may also assume for all .
Note that each formula of the form excludes only a single solution. Since we assume and all formulas of the form
are solved by cosets of , we may moreover assume .
By Lemma 4.19 we have for all -numbers . Thus we may use Lemma 4.19 to replace each by .
We consider formulas as linear congruences:
Hence it suffices to show that the system of linear congruences
has a solution in . After slightly adjusting the system (by using Lemma 4.19) and renaming, we get a system
where and every index is infinite or trivial. If there is an element such that is maximal in , then we are done by Proposition 4.24. Hence suppose there is such that for all . Then is infinite for all . In particular, the congruence
can be ignored, since each -class consists of infinitely many classes. Hence we removed one linear congruence from the system. After iterating this, we can find such that is maximal. ∎
4.3 The monotone hull
Theorem 4.8 gives quantifier elimination up to a suitable language on . The following gives a tame expansion of which allows us to analyze the definable sets.
A binary relation on a linear ordering is called monotone if and only if it satisfies
The following result by Simon states that expanding a linear ordering by monotone binary relations is tame:
Proposition 4.25 (Proposition 4.1 and Proposition 4.2 of [14]).
Let be a linear order equipped with monotone binary relations and unary predicates such that every -definable monotone binary relation is given by one of the and every -definable unary predicate is given by one of the . Then has quantifier elimination and is dp-minimal.
Fix a theory as in the quantifier elimination statement and let be a model. Note that the definable relations , , and are monotone.
Definition 4.26.
Let be a set of unary predicates and monotone binary relations on the value set of .
-
(a)
We define to be the monotone hull of
i.e. the expansion of by all 0-definable (in ) unary relations and all 0-definable monotone binary relations on the value sort.
-
(b)
Set and define .
Note that is an expansion by definitions.
Proposition 4.27.
Let be as in Definition 4.26. Then admits quantifier elimination in the language .
Proof.
The successor function and its inverse are -definable. If is a monotone binary relation, then so is for all . The same holds true for -definable unary predicates. Therefore adding the successor function to the language does not add any new definable sets in . Hence Theorem 4.8 and Proposition 4.25 imply quantifier elimination in . ∎
4.4 Dp-minimality and distality
Let be a complete -theory as in Proposition 4.27.
Lemma 4.28.
Let be a sufficiently saturated model and let and be mutually indiscernible sequences in the group sort. Let be a singleton. Then one of the sequences is indiscernible over .
Proof.
We may assume that both sequences are indexed by a dense linear order. Suppose is not indiscernible over . By the quantifier elimination result this must be witnessed by a formula of the form
where is an -term, is a monotone binary relation on , and . Hence we can find tuples of the same order type such that
where is the tuple corresponding to .
After replacing or if necessary, we may assume that and have disjoint convex hulls in . We can extend to a sequence such that is an indiscernable sequence. Then
is a non-constant indiscernible sequence in the value sort that is not indiscernible over .
By Proposition 4.25 the value sort is dp-minimal. Therefore must be indiscernible over : Otherwise we could apply the above argument to the sequence to get a second non-constant indiscernible sequence in the value sort which is not indiscernible over . Since these two sequences would be mutually indiscernible, this would contradict dp-minimality of the value sort. ∎
Theorem 4.29.
is dp-minimal.
Proof.
Let be a sufficiently saturated model and let and be mutually indiscernible sequences. We will assume that both of them are indexed by a dense linear order. Let be a singleton. We aim to show that one of the sequences is indiscernible over .
Since is essentially an imaginary sort, we may assume that the sequences and live in the sort . Note that equality on the value sort can be expressed using the monotone binary relation . By the quantifier elimination result, the failure of indiscernibility must be witnessed by formulas of the following form:
-
1.
,
-
2.
,
-
3.
,
where is an -term, is a coloring on , is a monotone binary relation on , , and . One of the terms in the third case could also be a quantifier free 0-definable constant in the value sort. This case is analogous to case (b) below and therefore we will not consider it separately.
Note that a formula of the first type would imply that is algebraic over the parameters plugged in for . Hence it suffices to consider the other two types of formulas. If an indiscernible sequence is not indiscernible over , then this must be witnessed by of the same order type such that we are in one of the following cases:
-
(a)
We have
and
for some choice of and a relation .
-
(b)
We have
and
for some choice of and a relation .
-
(c)
We have
and
or
for some choice of and a monotone binary relation .
-
(d)
We have
and
or
for some choice of and a monotone binary relation .
The case corresponding to a coloring is essentially the same as (b) so we will not do it explicitly.
We will use Lemma 4.19 to assume that all the coincide: Let be a finite set of primes. We want to be able to work in . Fix a term
and write , for , . Since is a finite set of constants, the value of only depends on the order type of . Therefore
also only depends on the order type of . We have
because . If for all of the same order type as , then this value is a constant. If for all of the same order type as , then this value can always be calculated in . If we are not in one of these two cases, then the quantifier free 0-definable coloring
witnesses (in ) that is not indiscernible over . Hence we can work in and therefore we can assume that all the coincide (by Lemma 4.19). Moreover, to simplify the notation we will assume that all the are equal to .
We say that an indiscernible sequence has an approximation for over if there is a set such that is indiscernible over , is definable over , and the residue class of modulo is algebraic (in ) over parameters in .
We now assume that the mutually indiscernible sequences and both fail to be indiscernible over . Then this must be witnessed as in (a) to (d). Such a witness for (resp. ) is good if (resp. ) has an approximation for for a suitable defined as follows:
-
•
If the witness is given as in (a), then we set
If (for example) (), then . Therefore the residue class of modulo is algebraic over .
-
•
If the witness is given as in (b), then we set
If , then and therefore the residue class of modulo is algebraic over .
-
•
If the witness is given as in (c), we set
-
•
Now assume the witness is given as in (d). We set
Now put .
In particular, every witness of type (a) or (b) is good because and are mutually indiscernible. Recall that if , then . We aim to show that we can always find a good witness:
Suppose the witness is given as in (c). Choose of the same order type as and such that all indices involved in are smaller than the indices in and (from now on, we will write in that case). If , then either the pair or the pair gives a good witness as in case (b).
Hence we will assume . Let be the sequence consisting of all elements of with index larger than all indices in and set to be the sequence where each tuple is expanded by . Then and are mutually indiscernible. Moreover, is not indiscernible over (as witnessed by and ). Hence is indiscernible over by Lemma 4.28. Now is indiscernible over the set and we have . Therefore the residue class of modulo is algebraic over and hence the witness is good.
Now suppose the witness is given as in (d). We set
If , then we have a good witness by the same arguments as in (a) and (b). Hence assume . Suppose for all we have . Fix
Consider the mutually indiscernible sequences and .
Assume that is indiscernible over . Then the residue class of modulo is algebraic over . Since , we get by indiscernibility (applied to and ). Therefore and only depend on the residue class of modulo (and can be calculated in ) and hence cannot witness the failure of indiscernibility over .
Hence is not indiscernible over . Then is indiscernible over by Lemma 4.28. Therefore is indiscernible over and the residue class of modulo is algebraic over . Hence we have a good witness.
Hence we assume that there is such that
If , then and we have a good witness as in cases (a) and (b). Hence we assume .
If , then or gives a good witness as in case (a). If , then either gives a witness as in case (a) or the new witness is given by and we have
Hence we are again in case (d) but is indiscernible over
and hence this witness given by must be good.
Now only the case is left. We then have
Assume the witnessing formula was of the form
for a monotone binary relation (the other case is done analogously).
We then have the following implications by monotonicity:
Hence must be false and must be true (since this was a witness for the failure of indiscernibility over ). Then must be true. But then and give a witness as in (a). Hence we can always find a good witness.
Since we assume that both and fail to be indiscernible over , we can find a good witness for each of them. Let be the constant for the witness in and let be the constant for the witness in . We assume . Then is indiscernible over and over the residue class of in .
Suppose the witness for is given as in (a) or (b). If we have
then and indiscernibility (and algebraicity of modulo over a suitable parameter) imply that . Hence those values only depend on the residue class of modulo (and can be calculated in with the restricted valuation). Therefore they cannot witness the failure of indiscernibility over .
Now suppose the witness for is given as in case (c). If , then this cannot be a witness for the failure for indiscernibility. Hence we must have . But then
only depends on the residue class of in and we can argue as before. The same arguments work if the witness for is given as in case (d).
Hence or must be indiscernible over . ∎
To characterize distality we will show that the quotients are stable. We will make use of the following lemma:
Lemma 4.30 (Lemma 5.13 of [2]).
Let be any language and let be an unstable -theory. Let be such that is stable. Then there exists an -formula , , over and a parameter such that is not -definable.
Proposition 4.31.
Suppose is infinite. Then the induced structure on is stable.
Proof.
Suppose is infinite. By Lemma 4.30 it suffices to show that for every formula (in ) and every constant the formula is definable in the pure group .
Given such a formula there is an -formula such that is the preimage of under the natural projection
Now fix a preimage of . Note that is a union of cosets of .
By the quantifier elimination result is equivalent to a boolean combination of atomic -formulas. We aim to show the following:
Claim.
There is a formula which is defined in the pure abelian group such that and coincide on all but finitely many cosets of .
It suffices to prove the claim for atomic formulas. Therefore we may assume that is atomic. Then we are in one of the following cases:
-
(a)
,
-
(b)
,
-
(c)
.
In case (a) there is nothing to show. Therefore we consider the cases (b) and (c) which include valuations. We show that sets of the form
are definable in the pure abelian group up to a unique coset of :
If , then and has finite index in . Hence has finite index in . Moreover, is of the form
for a positive integer because this holds true for the standard models. The cosets of are definable in the pure group language. If , then is definable in the pure group language. Now assume . Then and therefore is trivial outside of a single coset of .
This also shows that there are only finitely many intersections of the form
where everything except is fixed. Therefore the restriction of to is given by a finite chain of definable subgroups (in the pure abelian group ).
Since has only finitely many solutions modulo the same holds true for the valuation restricted to : Outside of finitely many cosets of it is given by a finite chain of -definable subgroups. In that sense is -definable outside of finitely many cosets of .
Therefore the formula in (a) or (b) is definable in outside of finitely many cosets of . This shows the claim.
Hence we can find such a formula defined in the pure abelian group such that and coincide on all but finitely many cosets of . The usual quantifier elimination result for abelian groups shows that is a boolean combination of cosets of the trivial subgroup and groups of the form for . Each subgroup has finite index in and the family is closed under finite intersections. Hence a boolean combination of such groups is a union of finitely many cosets of for a suitable .
Since and agree on all but finitely many cosets of and is a union of cosets of , the same must be true for , i.e. is a boolean combination of cosets of -definable subgroups. Therefore is definable in . ∎
Theorem 4.32.
is distal if and only if there is a constant such that
holds for all .
Proof.
Suppose is unbounded. Then there is some such that is infinite. By Proposition 4.31 the induced structure on is stable. Hence it follows from Proposition 2.19 that is not distal.
Now let be a non-constant totally indiscernible set of singletons and fix . Put . If , then and hence
It follows easily from total indiscernibility that does not depend on the choice of . Hence . ∎
5 Valuations on the integers
The most well-known example of a dp-minimal expansion of is . Based on work by Palacín and Sklinos [12], Conant and Pillay [5] proved the remarkable result that has no proper stable expansions of finite dp-rank. Hence any proper dp-minimal expansion must be unstable. The other known examples of dp-minimal expansions are:
-
•
where is the -adic valuation on . This was shown by Alouf and d’Elbée in [2].
-
•
where is cyclic order. These were found by Tran and Walsberg in [21].
-
•
Proper dp-minimal expansions of , where is a dense cyclic order, and were very recently found by Walsberg in [22].
An overview about the current research on dp-minimal expansions of is given by Walsberg in Section 6 of [22].
5.1 A single valuation
We add the following family of examples which generalize the -adic examples by Alouf and d’Elbée:
Theorem 5.1.
Let be a strictly descending chain of subgroups of , , let be the valuation defined by
and let be a set of unary predicates and monotone binary relations on the value set. Then admits quantifier elimination in the language (with and as constants) and is dp-minimal. Moreover, is distal if and only if the size of the quotients is bounded.
Proof.
Note that any infinite strictly descending chain of subgroups of must have trivial intersection. Moreover, every non-trivial subgroup of is of the form for some and hence is a good valuation in the sense of Definition 4.1.
Moreover, and is dense. Hence Proposition 4.27 implies the quantifier elimination result. Dp-minimality follows by Theorem 4.29 and the claim about distality follows by Theorem 4.32. ∎
In case of the -adic valuation Alouf and d’Elbée proved in Theorem 1.1 of [2] that has quantifier elimination in the language where
Conant [4] showed that the structure is a minimal proper expansion of , i.e. there is no proper intermediate expansion. Alouf and d’Elbée proved the same for . We will show that this does not hold true for arbitrary valuations.
Proposition 5.2.
Fix distinct primes and put . For fix , set , and recursively define
Set to be the valuation corresponding to and let be the valuation corresponding to . Then is definable in .
Proof.
If , then there is a unique such that . Let . If , then
If , then can be determined using and . ∎
Corollary 5.3.
Let be as in Proposition 5.2. Then there are many valuations such that is definable in . Only countably many of those can be definable in .
Proof.
There are many valuations as in Proposition 5.2 and is definable in each . On the other hand, has only countably many definable sets. ∎
Remark 5.4.
Note that by Theorem 4.32 all these structures are distal. Hence not even all dp-minimal distal expansions by valuations are minimal expansions.
The fact that expansions by arbitrary valuations are dp-minimal allows us to construct other non-trivial examples: For let denote the valuation corresponding to the sequence .
Proposition 5.5.
Let and be coprime positive integers. Then the expansion
is dp-minimal.
Proof.
We have
Hence the relation is definable in the dp-minimal structure . ∎
It seems unlikely that is definable from .
The induced structure on the index set seems to be important. If it is not o-minimal and is a definable infinite and co-infinite subset, then the set
is definable. It is not clear if is definable in .
If the induced structure on is o-minimal, then must be constant for all sufficiently large (in some elementary extension). If is finite, then the size of the quotients is bounded and hence we are in the distal case.
Conjecture 5.6.
Let be distal. Then the following are equivalent:
-
(a)
is a minimal expansion of ,
-
(b)
there is a prime such that for almost all ,
-
(c)
is interdefinable with a -adic valuation for some prime ,
-
(d)
the -induced structure on the value set of is o-minimal for all -definable valuations .
Proposition 5.7.
If (a) implies (d), then 5.6 holds.
Proof.
We already know (b) (c) (a) and by assumption (a) (d) holds. Hence (d) (b) remains to be shown.
Let be distal and assume (d). Then there is such that
for almost all . Therefore and are interdefinable and we may assume .
If where and are coprime, then is interdefinable with the valuation such that alternates between and . Then the induced structure on the value set of is not o-minimal. This contradicts (d).
Hence we may assume for some prime and . If , then the -adic valuation is definable by
Now the set contradicts (d).
Hence must be a prime. This shows (b). ∎
There are non-distal candidates for minimal expansions:
Question 5.8.
Let be an enumeration of the primes such that each prime appears exactly once and let be the valuation corresponding to . Then the induced structure on the value set is o-minimal. Is a minimal expansion of ?
We end this section with the observation that the -adic valuations have a limit theory:
Proposition 5.9.
For each prime let denote the -adic valuation on . Then the corresponding limit theory exists, i.e.
does not depend on the choice of the non-principal ultrafilter .
Proof.
We fix the common -theory
of these ultraproducts. Note that the predicate fails for all and the predicate holds true for all . Thus they are quantifier free 0-definable after naming the successor function on . Therefore has quantifier elimination after naming by Theorem 4.8 (because has quantifier elimination). The constants in generate a subgroup that is isomorphic to . An element must have valuation in all models of . Therefore all models of have isomorphic substructures and hence is complete. ∎
5.2 Multiple valuations
If is a non-empty set of primes, then Alouf and d’Elbée proved that the structure has dp-rank exactly . We will generalize this result to expansions of by arbitrary valuations which involve disjoint sets of primes.
Let be a non-empty family of non-trivial valuations . For each set
We view together with these valuations as a multi-sorted structure with group sort and with a distinct value sort for each valuation . Now put
-
•
,
-
•
, and
-
•
for each valuation . Let be the monotone hull of as in Section 4.3 and set
to be the disjoint union of these languages.
Proposition 5.10.
Suppose the sets are pairwise disjoint. Then has quantifier elimination in the language .
Proof.
This is very similar to the proof of Theorem 4.8. Note that a multi-sorted version of Lemma 2.21 holds true in this setting. As in the proof of Theorem 4.8 it suffices to show the back-and-forth property for systems of linear congruences. Let be the system
where and are finite index sets for each for a finite subset . By an application of Lemma 4.19 we may assume that all the and have the same value which we denote by .
Let be a solution. We will show that we can assume that and all constants and are contained in : If is in , then all must be contained in since otherwise the congruences can not be satisfied. If is not contained in , then the congruence
does not impose any restrictions on and we can ignore it without changing the solutions in .
If is not contained in , then there is a constant (and hence in the language) such that . In that case the shifted system :
is solved by and all the constants and can be assumed to lie in . Thus we can replace by .
Hence we may assume that is a system of linear congruences in the subgroup . We have
and the valuations involve disjoint sets of primes. Therefore the system can be solved independently for each valuation . This is done as in the proof of Theorem 4.8. ∎
Theorem 5.11.
Suppose the sets are pairwise disjoint. Then
Proof.
is shown exactly as in the case of the -adic valuations which was done by Alouf and d’Elbée (Theorem 1.2 of [2]).
Now assume . As in the proof of Theorem 4.29 this is witnessed by mutually indiscernible sequences (in the group sort) and a singleton in the group sort such that no sequence is indiscernible over . As argued in Theorem 4.29, the fact that a sequence is not indiscernible over must be witnessed by an atomic -formula which involves a valuation.
Since , there must be two sequences and for which this witnessing formula involves the same valuation . This is a contradiction because is dp-minimal by Theorem 4.29. ∎
6 Further results
6.1 Uniformly definable families of finite-index subgroups of dp-minimal groups
The classification of NIP profinite groups by Macpherson and Tent in [19] yields information about uniformly definable families of finite index subgroups in arbitrary NIP groups (see Theorem 8.7 in [3]). We will do the same in the dp-minimal case. The arguments are almost identical to those in Section 8 of [3] (see also Remark 5.5 in [19]), we only need to make sure that the construction presented there preserves dp-minimality.
Let be a group and let be a family of normal subgroups of finite index such that
We view as an -structure . Let be the projection maps. Then is a neighborhood basis at the identity. Therefore we may view as an -structure .
Lemma 6.1.
Let be an -saturated elementary extension of . Then
and is a neighborhood basis for the identity consisting of open normal subgroups.
Proof.
By elementarity we have for all . Using this and elementarity it is easy to see that
Now write
and let be the natural homomorphism. Clearly . It remains to show that is surjective.
Fix and consider the partial type
Given finite, there is such that . Then for all . Hence is finitely satisfiable and as is -saturated, there exists such that for all and hence .
The family is a neighborhood basis for the identity consisting of open normal subgroups. ∎
Lemma 6.2.
If has NIP, then has NIP. If moreover is dp-minimal, then is dp-minimal.
Proof.
Since has NIP, every uniformly definable family of subgroups contains only finitely many subgroups of each finite index. Let be an -saturated elementary extension. Then is externally definable (since ). If is dp-minimal, then is dp-minimal by Remark 2.16. By the above lemma the structure is interpretable as a quotient in and hence is NIP (resp. dp-minimal). ∎
Let be a group and let be a formula. Set to be the family of all normal subgroups which are finite intersections of conjugates of -definable subgroups of finite index. Note that every -definable subgroup of finite index contains some . The profinite group naturally becomes an -structure .
Proposition 6.3.
Let and be as above. If is NIP, then is NIP. If moreover is dp-minimal, then is dp-minimal in the group sort.
Proof.
By Baldwin-Saxl finite intersections of conjugates of -definable subgroups are uniformly definable by some formula . The set is definable. Put . Then . Since is closed under intersections, it follows that is externally definable. Let be the equivalence relation defined by . Now apply the previous lemma to the structure . ∎
By Proposition 3.3 every dp-minimal profinite group has an open abelian subgroup. Now Proposition 6.3 implies the following:
Proposition 6.4.
Let be a dp-minimal group and let be a formula. Let be the family of all normal subgroups which are finite intersections of conjugates of -definable subgroups of finite index. If is infinite, then there is such that for all the quotient is abelian.
Proof.
The profinite group is dp-minimal and therefore is virtually abelian by Proposition 3.3. Since the quotients are preserved, this implies the proposition. ∎
Remark 6.5.
By Theorem 3.4 there are essentially two types of dp-minimal profinite groups. This will also be seen in the abelian quotients in the statement of Proposition 6.4.
6.2 Strong homogeneity of profinite groups
Jarden and Lubotzky [8] showed that two elementarily equivalent profinite groups are isomorphic if one of them is finitely generated. This was generalized to strongly complete profinite groups by Helbig [7]. A profinite group is strongly complete if all subgroups of finite index are open. The tools used by Helbig and the construction in Section 6.1 give a proof for strong homogeneity.
Let be a profinite group and suppose is a neighborhood basis at the identity consisting of open normal subgroups. Let be the group language expanded by a family of unary predicates . We consider as an structure by setting . Note that if is an elementary expansion, then there is a natural -structure on the quotient .
Lemma 6.7.
Let a profinite group equipped with an structure as above. Let be an elementary extension of in the language . Then the composition
is an -isomorphism.
Proof.
The lemma follows from the same arguments as Lemma 6.1. ∎
Proposition 6.8.
Let and be profinite groups as structures such that the predicates encode neighborhood bases at the identity consisting of open normal subgroups in both groups. Suppose is a subset and is an elementary map with respect to the language . Then extends to an -isomorphism between and .
Proof.
Let be a common strongly -homogeneous elementary extension of and . We can find such that . Since is an -automorphism, it induces an automorphism of . Now use Lemma 6.7 to get the desired isomorphism between and . ∎
The following observation in Remark 3.12 in [7] is a consequence of Theorem 2 in [18] and Corollary 52.12 in [11]:
Theorem 6.9.
Let be a profinite group. Then the following are equivalent:
-
(a)
is strongly complete.
-
(b)
For each finite group there exists a group word such that and is open in .
Recall that a group word has finite width if for some . We will make use of the following result:
Proposition 6.10 (Proposition 5.2(b) of [23]).
Let be a profinite group. If is a group word, then is closed in if and only if has finite width in .
Proposition 6.11.
Let and be profinite groups. Let be a subset of and let be an elementary map. If one of the groups is strongly complete, then extends to an isomorphism.
Proof.
By Theorem 6.9 and Proposition 6.10 strong completeness is a first-order property among profinite groups. For each finite group there is a group word such that , is open in , and is open in . Note that by Proposition 6.10 and elementary equivalence of and , and are definable by the same formula without parameters.
If is an open normal subgroup of then and hence . Therefore the family is a neighborhood basis at the identity.
Hence we may consider and as -structures where the predicates are given by . By Proposition 6.8 extends to an isomorphism. ∎
6.3 A result on families of subgroups of NTP2 groups
By Theorem 1.1 of [19] a full profinite group is NIP if and only if it is NTP2. Since the structure of these groups is determined by the lattice of subgroups, this only depends on a single formula. We will show a version for formulas in NTP2 groups. We will use the following lemma by Macpherson and Tent on groups in NTP2:
Lemma 6.12 (Lemma 4.3 in [19]).
Let be an -definable group in a structure with NTP2 theory, and a formula implying . Then there is such that the following holds. Suppose that is a subgroup of , is an epimorphism to the Cartesian product of the groups , and is for each the composition of with the canonical projection . Suppose also that for each , there is a subgroup and group with , such that finite intersections of the groups are uniformly definable by instances of . Then .
Proposition 6.13.
Let be an NTP2 group and let be a formula such that . Suppose that the family consists of normal subgroups of and is closed under finite intersections. Then has NIP.
Proof.
We aim to show that satisfies the Baldwin-Saxl condition. Let be instances of and fix as in Lemma 6.12. Now set and . Note that and we have if . Now set
We then have
Now set and assume that all are non-trivial. Let
be the natural projection and let be the projection on . Set and put . Then
Hence Lemma 6.12 implies .
If , then there must be such that and hence can be written as in intersection of instances of . Inductively this shows that any intersection of instances of is an intersection of at most instances. Hence satisfies the Baldwin-Saxl condition.
This implies that has NIP: Otherwise we can find a constant and an indiscernible sequence such that
Set and take , all of them odd. By the Baldwin-Saxl lemma we may assume
By indiscernibility this implies
But this is a contradiction since . ∎
While it is clearly sufficient to assume that the -definable subgroups normalize each other, the above proof requires some normality assumption.
Question 6.14.
Does Proposition 6.13 hold even without any normality assumption?
References
- [1] Eran Alouf. On dp-minimal expansions of the integers, 2020. arXiv:2001.11480 [math.LO].
- [2] Eran Alouf and Christian D’Elbée. A new dp-minimal expansion of the integers. The Journal of Symbolic Logic, 84(2):632–663, 2019.
- [3] Tim Clausen and Katrin Tent. Some model theory of profinite groups. Lectures in Model Theory, Münster Lectures in Mathematics, EMS Publishing, 2016.
- [4] Gabriel Conant. There are no intermediate structures between the group of integers and presburger arithmetic. The Journal of Symbolic Logic, 83(1):187–207, 2018.
- [5] Gabriel Conant and Anand Pillay. Stable groups and expansions of . Fundamenta Mathematicae, 01 2016.
- [6] J. D. Dixon, M. P. F. Du Sautoy, A. Mann, and D. Segal. Analytic Pro-P Groups. Cambridge Studies in Advanced Mathematics. Cambridge University Press, 2 edition, 1999.
- [7] Patrick Helbig. On small profinite groups. Journal of Group Theory, 20(5), 2017.
- [8] Moshe Jarden and Alexander Lubotzky. Elementary equivalence of profinite groups. Bulletin of The London Mathematical Society, 40:887–896, 07 2008.
- [9] Franz-Viktor Kuhlmann and Salma Kuhlmann. Ax-Kochen-Ershov principles for valued and ordered vector spaces. In W. Charles Holland, editor, Ordered Algebraic Structures, pages 237–259, Dordrecht, 1997. Kluwer.
- [10] Fares Maalouf. Espaces vectoriels c-minimaux. J. Symbolic Logic, 75(2):741–758, 06 2010.
- [11] Hanna Neumann. Varieties of Groups, volume 37 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 2. Folge. Springer, first edition, 1967.
- [12] Daniel Palacín and Rizos Sklinos. On superstable expansions of free abelian groups. Notre Dame J. Formal Log., 59:157–169, 2018.
- [13] Luis Ribes and Pavel Zalesskii. Profinite Groups, volume 40 of A Series of Modern Surveys in Mathematics. Springer, second edition, 2000.
- [14] Pierre Simon. On dp-minimal ordered structures. J. Symb. Log., 76:448–460, 2011.
- [15] Pierre Simon. Distal and non-distal nip theories. Ann. Pure Appl. Log., 164:294–318, 2013.
- [16] Pierre Simon. A Guide to NIP Theories. Cambridge, 2015.
- [17] Patrick Simonetta. An example of a c-minimal group which is not abelian-by-finite. Proceedings of the American Mathematical Society, 131, 12 2003.
- [18] M.G. Smith and John Wilson. On subgroups of finite index in compact hausdorff groups. Archiv der Mathematik, 80:123–129, 04 2003.
- [19] Katrin Tent and Dugald Macpherson. Profinite groups with NIP theory and p-adic analytic groups. Bull. Lond. Math. Soc., 48(6):1037–1049, 2016.
- [20] Katrin Tent and Martin Ziegler. A Course in Model Theory. Lecture Notes in Logic. Cambridge University Press, 2012.
- [21] Minh Tran and Erik Walsberg. A family of dp-minimal expansions of , 2017. arXiv:1711.04390 [math.LO].
- [22] Erik Walsberg. Dp-minimal expansions of via dense pairs via Mordell-Lang, 2020. arXiv:2004.06847 [math.LO].
- [23] J.S. Wilson. Profinite Groups. Clarendon Press, 1998.
- [24] E. I. Zelmanov. On periodic compact groups. Israel Journal of Mathematics, 77:83–95, 1992.