This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On the proof of elimination of imaginaries in algebraically closed valued fields

Will Johnson

ACVF is the theory of non-trivially valued algebraically closed valued fields. This theory is the model companion of the theory of valued fields. ACVF does not have elimination of imaginaries in the home sort (the valued field sort). Nevertheless, Haskell, Hrushovski, and Macpherson in [1] were able to find a collection of “geometric sorts” in which elimination of imaginaries holds.

Let KK be a model of ACVF, with valuation ring 𝒪\mathcal{O} and residue field kk. A lattice in KnK^{n} is an 𝒪\mathcal{O}-submodule ΛKn\Lambda\subseteq K^{n} isomorphic to 𝒪n\mathcal{O}^{n}. Let SnS_{n} denote the set of lattices in KnK^{n}. This is an interpretable set; it can be identified with GLn(K)/GLn(𝒪)GL_{n}(K)/GL_{n}(\mathcal{O}). For each lattice ΛKn\Lambda\subseteq K^{n}, let resΛ\operatorname{res}\Lambda denote Λ𝒪k\Lambda\otimes_{\mathcal{O}}k, a kk-vector space of dimension nn. Let TnT_{n} be

Tn=ΛSnresΛ={(Λ,x):ΛSn,xresΛ}.T_{n}=\bigcup_{\Lambda\in S_{n}}\operatorname{res}\Lambda=\{(\Lambda,x):\Lambda\in S_{n},~x\in\operatorname{res}\Lambda\}.

This set is again interpretable.

The main result of [1] is the following:

Theorem (Haskell, Hrushovski, Macpherson).

ACVF eliminates imaginaries relative to the sorts KK, {Sn:n1}\{S_{n}:n\geq 1\} and {Tn:n1}\{T_{n}:n\geq 1\}.

The proof in [1] is long and technical, and we aim to give a more straightforward proof. Our proof is a variant of Hrushovski’s shorter proof in [3], except that our strategy for coding definable types is different—see 4.2. We also give a new proof that finite sets of modules can be coded in the geometric sorts—see 4.3.

The proof of elimination of imaginaries given here aims to be more conceptual and less technical than previous proofs. We prove no new results. We include many details that are well-known at this point, for the sake of being self-contained.

1 Review of ACVF

1.1 Notation

In a model of ACVF, KK is the home sort (the valued field), 𝒪K\mathcal{O}\subseteq K is the valuation ring, 𝔐\mathfrak{M} is the maximal ideal in 𝒪\mathcal{O}, k=𝒪/𝔐k=\mathcal{O}/\mathfrak{M} is the residue field, Γ=K×/𝒪×\Gamma=K^{\times}/\mathcal{O}^{\times} is the valuation group, res:𝒪k\operatorname{res}:\mathcal{O}\to k is the residue map, and val:KΓ{+}\operatorname{val}:K\to\Gamma\cup\{+\infty\} is the valuation. The value group is written additively, and ordered so that

𝒪={xK:val(x)0}.\mathcal{O}=\{x\in K:\operatorname{val}(x)\geq 0\}.

A lattice in KnK^{n} is an 𝒪\mathcal{O}-submodule Λ\Lambda of KnK^{n} which is free of rank nn, i.e., isomorphic to 𝒪n\mathcal{O}^{n}. If Λ\Lambda is a lattice, resΛ\operatorname{res}\Lambda will denote Λ/𝔐Λ=Λ𝒪k\Lambda/\mathfrak{M}\Lambda=\Lambda\otimes_{\mathcal{O}}k. This is always an nn-dimensional kk-vector space. We will use the following interpretable sets:

  • SnS_{n}, the set of lattices in KnK^{n}.

  • TnT_{n}, the set of pairs (Λ,ξ)(\Lambda,\xi), where ΛSn\Lambda\in S_{n} and ξresΛ\xi\in\operatorname{res}\Lambda

  • Rn,R_{n,\ell}, the set of pairs (Λ,V)(\Lambda,V), where ΛSn\Lambda\in S_{n} and VV is an \ell-dimensional subspace of resΛ\operatorname{res}\Lambda.

Each of these sets is easily interpretable in ACVF. Our main goal will be to prove that ACVF has elimination of imaginaries in the sorts KK and Rn,R_{n,\ell}. In §5, we will note how this implies elimination of imaginaries in K,SnK,S_{n}, and TnT_{n}, the standard “geometric sorts” of [1]. But until then, the term “geometric sorts” will mean the sorts KK and Rn,R_{n,\ell}.

When working in an abstract model-theoretic context, the monster model will be denoted 𝕄\mathbb{M}. If a definable set or other entity XX has a code in 𝕄eq\mathbb{M}^{eq}, the code will be denoted X\ulcorner X\urcorner. Unless stated otherwise, “definable” will mean “interpretable.”

1.2 Basic facts

We assume without proof the following well-known facts about ACVF. Many of these are discussed in [5].

  • Models of ACVF are determined up to elementary equivalence by characteristic and residue characteristic, which must be (0,0)(0,0), (p,p)(p,p), or (0,p)(0,p) for some prime pp.

  • ACVF has quantifier elimination in the language with one sort KK, with the ring structure on KK, and with a binary predicate for the relation val(x)val(y)\operatorname{val}(x)\geq\operatorname{val}(y).

  • C-minimality: Every definable subset DK1D\subseteq K^{1} is a boolean combination of open and closed balls (including points). More precisely, DD can be written as a disjoint union of “swiss cheeses,” where a swiss cheese is a ball with finitely many proper subballs removed. There is a canonical minimal way of decomposing DD as a disjoint union of swiss cheeses. All the balls involved in this decomposition are algebraic over the code for DD.

  • The theory ACVF does not have the independence property. That is, ACVF is NIP.

  • The value group Γ\Gamma is o-minimal, in the sense that every definable subset of Γ1\Gamma^{1} is a finite union of points and intervals with endpoints in Γ{±}\Gamma\cup\{\pm\infty\}. (In fact, Γ\Gamma is a stably embedded pure divisible ordered abelian group.)

  • The residue field kk is strongly minimal, hence stable and stably embedded. Moreover, every definable subset of knk^{n} is coded by a tuple from kk. (In fact, kk is a stably embedded pure algebraically closed field.)

The first two points are due to Robinson [7], and the third is due to Holly [2]. The last three points are easy consequences of C-minimality111The final point uses the following fact: if TT is a strongly minimal theory, in which acl()\operatorname{acl}(\emptyset) is infinite and finite sets of tuples are coded by tuples, then TT eliminates imaginaries. This is Lemma 1.6 in [6]., and the last two can also be seen from the quantifier elimination result in the three-sorted language discussed in [5] and [1].

1.3 Valued KK-vector spaces

Let KK be an arbitrary valued field. Following Section 2.5 of [3],

Definition 1.1.

A valued KK-vector space is a KK-vector space VV and a set Γ(V)\Gamma(V) together with the following structure:

  • A total ordering on Γ(V)\Gamma(V)

  • An action

    +:Γ(K)×Γ(V)Γ(V)+:\Gamma(K)\times\Gamma(V)\to\Gamma(V)

    of Γ(K)=Γ\Gamma(K)=\Gamma on Γ(V)\Gamma(V), strictly order-preserving in each variable (hence free)

  • A surjective map val:V{0}Γ(V)\operatorname{val}:V\setminus\{0\}\to\Gamma(V), such that

    val(w+v)min(val(w),val(v))\operatorname{val}(w+v)\geq\min(\operatorname{val}(w),\operatorname{val}(v))
    val(αv)=val(α)+val(v)\operatorname{val}(\alpha\cdot v)=\operatorname{val}(\alpha)+\operatorname{val}(v)

    for w,vVw,v\in V and αK\alpha\in K, with the usual convention that val(0)=+>Γ(V)\operatorname{val}(0)=+\infty>\Gamma(V).

This is merely a variation on the notion of a normed vector space over a field with an absolute value.

Remark 1.2.

If dimKV\dim_{K}V is finite, then the action of Γ(K)\Gamma(K) on Γ(V)\Gamma(V) has finitely many orbits. In fact,

|Γ(V)/Γ(K)|dimKV.|\Gamma(V)/\Gamma(K)|\leq\dim_{K}V.
Proof.

Let v1,,vnv_{1},\ldots,v_{n} be non-zero vectors with val(vn)\operatorname{val}(v_{n}) in different orbits of Γ(K)\Gamma(K). We will show that the viv_{i} are linearly independent. If not, let w1,,wmw_{1},\ldots,w_{m} be a minimal subset which is linearly dependent. Then ixiwi=0\sum_{i}x_{i}w_{i}=0 for some xiK×x_{i}\in K^{\times}. But by assumption,

val(xiwi)=val(xi)+val(wi)val(xj)+val(wj)=val(xjwj)\operatorname{val}(x_{i}w_{i})=\operatorname{val}(x_{i})+\operatorname{val}(w_{i})\neq\operatorname{val}(x_{j})+\operatorname{val}(w_{j})=\operatorname{val}(x_{j}w_{j})

for any iji\neq j. By the ultrametric inequality in VV, ixiwi\sum_{i}x_{i}w_{i} cannot be zero, a contradiction. ∎

For the rest of this section, we will assume that all valued KK-vector spaces VV have value group Γ(V)=Γ(K)=Γ\Gamma(V)=\Gamma(K)=\Gamma, since the goal is Theorem 1.5.

If VV and WW are two such valued KK vector spaces, we can form a “direct sum” VWV\oplus W by setting

val(v,w)=min(val(v),val(w)).\operatorname{val}(v,w)=\min(\operatorname{val}(v),\operatorname{val}(w)).

For example, KnK^{\oplus n} is a valued KK-vector space with underlying vector space KnK^{n}, with value group Γ(K)\Gamma(K), and with valuation map given by

val(x1,,xn)=min(val(x1),,val(xn)).\operatorname{val}(x_{1},\ldots,x_{n})=\min(\operatorname{val}(x_{1}),\ldots,\operatorname{val}(x_{n})).

If VV and WW are two subspaces of a valued KK-vector space, say that VV and WW are perpendicular if VW=V\cap W=\emptyset and V+WV+W is isomorphic to VWV\oplus W. In other words, VV and WW are perpendicular if val(v+w)=min(val(v),val(w))\operatorname{val}(v+w)=\min(\operatorname{val}(v),\operatorname{val}(w)) for every vVv\in V and wWw\in W.

Recall that a valued field KK is spherically complete if every descending sequence of balls in KK has non-empty intersection. If VV is a valued KK-vector space, a ball in VV is a set of the form

{val(xa)γ} or {val(xa)>γ}\{\operatorname{val}(x-a)\geq\gamma\}\text{ or }\{\operatorname{val}(x-a)>\gamma\}

for aVa\in V and γΓ(V)\gamma\in\Gamma(V). We say that VV is spherically complete if every descending sequence of balls in VV has a non-empty intersection.

Remark 1.3.
  1. 1.

    If VV and WW are spherically complete, so is VWV\oplus W, because the balls in VWV\oplus W are of the form B1×B2B_{1}\times B_{2}, with B1B_{1} a ball in VV and B2B_{2} a ball in WW.

  2. 2.

    If VV is a subspace of a valued KK-vector space WW, and aWa\in W, then the intersection of any ball in WW with a+Va+V is either empty or a ball in the affine subspace a+Va+V.

  3. 3.

    If VV is a spherically complete subspace of WW and aWa\in W, and \mathcal{F} is the family of closed balls in WW centered at the origin which intersect a+Va+V, then V\mathcal{F}\cap V is a nested chain of balls in a+Va+V, so it has a non-empty intersection. Equivalently, the following set has a maximum:

    {val(a+v):vV}.\{\operatorname{val}(a+v):v\in V\}.

    That is, some element of a+Va+V is maximally close to 0.

Lemma 1.4.

Let WW be a valued KK-vector space. Let VV be a subspace. Suppose that aWVa\in W\setminus V is maximally close to 0 among elements of a+Va+V. Then KaK\cdot a is perpendicular to VV.

Proof.

We need to show that

val(v+αa)=min(val(v),val(αa))\operatorname{val}(v+\alpha a)=\min(\operatorname{val}(v),\operatorname{val}(\alpha a)) (1)

for vVv\in V and αK\alpha\in K. Replacing vv and α\alpha with α1v\alpha^{-1}v and α1α\alpha^{-1}\alpha changes both sides of (1) by the same amount, so we may assume that α=0\alpha=0 or α=1\alpha=1.

The α=0\alpha=0 case is trivial. Suppose that α=1\alpha=1; we want to show val(v+a)=min(val(v),val(a))\operatorname{val}(v+a)=\min(\operatorname{val}(v),\operatorname{val}(a)). If val(v)val(a)\operatorname{val}(v)\neq\operatorname{val}(a), then val(v+a)=min(val(v),val(a))\operatorname{val}(v+a)=\min(\operatorname{val}(v),\operatorname{val}(a)) by the ultrametric inequality. In the case where val(v)=val(a)\operatorname{val}(v)=\operatorname{val}(a), the ultrametric inequality only implies

val(v+a)min(val(v),val(a))=val(a).\operatorname{val}(v+a)\geq\min(\operatorname{val}(v),\operatorname{val}(a))=\operatorname{val}(a). (2)

But val(v+a)val(a)\operatorname{val}(v+a)\leq\operatorname{val}(a), by the assumption on aa. So equality holds in (2). ∎

Theorem 1.5.

Suppose KK is spherically complete, VV is an nn-dimensional KK-vector space, and Γ(K)=Γ(V)\Gamma(K)=\Gamma(V). Then VV is isomorphic to KnK^{\oplus n}. In other words, there is a basis {v1,,vn}V\{v_{1},\ldots,v_{n}\}\subseteq V such that

val(x1v1++xnvn)=min1inval(xi) for every xKn.\operatorname{val}(x_{1}v_{1}+\cdots+x_{n}v_{n})=\min_{1\leq i\leq n}\operatorname{val}(x_{i})\text{ for every $\vec{x}\in K^{n}$}.

In [3], Hrushovski calls {v1,,vn}\{v_{1},\ldots,v_{n}\} a “separating basis.”

Proof.

Proceed by induction on dimKV\dim_{K}V. The one-dimensional case is easy. Let VV^{\prime} be a codimension 1 subspace. By induction, VV^{\prime} is isomorphic to K(n1)K^{\oplus(n-1)}, so VV^{\prime} is spherically complete. Choose some a0VVa_{0}\in V\setminus V^{\prime} and let aa be an element of a0+Va_{0}+V^{\prime} maximally close to 0. By Lemma 1.4, KaK\cdot a is perpendicular to VV^{\prime}. Thus VVKK(n1)K=KnV\cong V^{\prime}\oplus K\cong K^{\oplus(n-1)}\oplus K=K^{\oplus n}. ∎

1.4 Definable submodules of KnK^{n}

We now return to the setting of ACVF.

Recall that every model of ACVF is elementarily equivalent to a spherical complete one.222This is well-known, and discussed in [5]. In the pure characteristic case, one can use fields of Hahn series. In the mixed characteristic case, one can use metric ultrapowers of p\mathbb{C}_{p}.

Theorem 1.6.

Let KK be a model of ACVF. Let VV be a definable KK-vector space, with dimKV<\dim_{K}V<\infty. Let NVN\subseteq V be a definable 𝒪\mathcal{O}-submodule. Then NN is isomorphic to Kn1×𝒪n2×𝔐n3K^{n_{1}}\times\mathcal{O}^{n_{2}}\times\mathfrak{M}^{n_{3}} for some n1,n2,n3<nn_{1},n_{2},n_{3}<n.

Proof.

We are trying to prove a conjunction of first-order sentences, so we may replace KK with an elementarily equivalent model. Therefore, we may assume KK is spherically complete.

Replacing VV with the KK-span of NN, we may assume that VV is the KK-span of NN. Similarly, if WW denotes the largest KK-vector space contained in NN, then by quotienting out WW, we may assume that NN contains no nontrivial KK-vector spaces. Now αK×αN=V\bigcup_{\alpha\in K^{\times}}\alpha N=V and αK×αN=0\bigcap_{\alpha\in K^{\times}}\alpha N=0.

For any nonzero vVv\in V, let

val(v)=sup{val(α):vαN}=inf{val(α):vαN}.\operatorname{val}(v)=\sup\{\operatorname{val}(\alpha):v\in\alpha N\}=\inf\{\operatorname{val}(\alpha):v\notin\alpha N\}.

This is well-defined by o-minimality of Γ\Gamma, and one easily checks that

val(βv)=val(β)+val(v).\operatorname{val}(\beta v)=\operatorname{val}(\beta)+\operatorname{val}(v). (3)
val(v)>0vN\operatorname{val}(v)>0\implies v\in N (4)
val(v)<0vN\operatorname{val}(v)<0\implies v\notin N (5)

for all βK\beta\in K, vVv\in V. We claim that val:VΓ\operatorname{val}:V\to\Gamma makes VV into a valued KK-vector space. Given (3), we merely need to check the ultrametric inequality

val(v+w)min(val(v),val(w)).\operatorname{val}(v+w)\geq\min(\operatorname{val}(v),\operatorname{val}(w)).

If this failed, then multiplying everything by an appropriate scalar, we would get

val(v+w)<0<min(val(v),val(w)).\operatorname{val}(v+w)<0<\min(\operatorname{val}(v),\operatorname{val}(w)).

But then v,wNv,w\in N and v+wNv+w\notin N, contradicting the fact that NN is a module.

So val:VΓ\operatorname{val}:V\to\Gamma makes VV into a valued KK-vector space. By Theorem 1.5, we can assume that VV is KnK^{\oplus n}. Then (4-5) mean the following for xKn\vec{x}\in K^{n}:

  • If val(xi)>0\operatorname{val}(x_{i})>0 for every ii, then xN\vec{x}\in N. In other words, 𝔐nN\mathfrak{M}^{n}\subseteq N.

  • If val(xi)<0\operatorname{val}(x_{i})<0 for some ii, then xN\vec{x}\notin N. In other words, N𝒪nN\subseteq\mathcal{O}^{n}.

So NN is sandwiched between 𝒪n\mathcal{O}^{n} and 𝔐n\mathfrak{M}^{n}. But the possibilities for NN then correspond to the submodules of 𝒪n/𝔐n\mathcal{O}^{n}/\mathfrak{M}^{n}, i.e., the kk-subspaces of knk^{n}. These are easy to deal with.

Specifically, note that N/𝔐nN/\mathfrak{M}^{n} is a kk-subspace of 𝒪n/𝔐n=kn\mathcal{O}^{n}/\mathfrak{M}^{n}=k^{n}. Let γ\gamma be an element of GLn(k)GL_{n}(k) sending N/𝔐nknN/\mathfrak{M}^{n}\subseteq k^{n} to k×0nknk^{\ell}\times 0^{n-\ell}\subseteq k^{n} for =dimkN/𝔐n\ell=\dim_{k}N/\mathfrak{M}^{n}. Then γ\gamma can be lifted to some γGLn(𝒪)\gamma^{\prime}\in GL_{n}(\mathcal{O}), because 𝒪\mathcal{O} is a local ring. If N=γ(N)N^{\prime}=\gamma^{\prime}(N), then N/𝔐nN^{\prime}/\mathfrak{M}^{n} is k×0n=(𝒪×𝔐n)/𝔐nk^{\ell}\times 0^{n-\ell}=(\mathcal{O}^{\ell}\times\mathfrak{M}^{n-\ell})/\mathfrak{M}^{n}. So N=𝒪×𝔐nKnN^{\prime}=\mathcal{O}^{\ell}\times\mathfrak{M}^{n-\ell}\subseteq K^{n}. But NN^{\prime} and NN are isomorphic. ∎

Let ModnMod_{n} denote the set of definable submodules of KnK^{n}. The theorem implies that the elements of ModnMod_{n} fall into finitely many definable families. Consequently, we get the following

Corollary 1.7.

The set ModnMod_{n} is interpretable.

2 Generalities on Definable Types

Work in an arbitrary theory TT, with monster model 𝕄\mathbb{M}. By “CC-definable type,” we will mean CC-definable type over the monster, as opposed to some smaller model, unless stated otherwise. By “definable type,” we mean a CC-definable type for some C𝕄C\subseteq\mathbb{M}.

In this section we review some well-known facts about definable types. We omit many of the proofs, which are usually straightforward.

2.1 Operations on definable types

If pp is a CC-definable type and ff is a CC-definable function, there is a unique CC-definable type fpf_{*}p which is characterized by the following property:

ap|Bf(a)fp|B, for all small BC and all a.a\models p|B\implies f(a)\models f_{*}p|B\text{, for all small $B\supseteq C$ and all $a$.}

The choice of CC does not matter—if pp and ff are CC^{\prime}-definable for some other set CC^{\prime}, then the resulting fpf_{*}p is the same. The type fpf_{*}p is called the pushforward of pp along ff.

If pp and qq are two CC-definable types, there is a unique CC-definable type pqp\otimes q which is characterized by the following property:

(a,b)pq|B(ap|Bb)(bq|B), for all small BC and all a,b.(a,b)\models p\otimes q|B\iff(a\models p|Bb)\wedge(b\models q|B)\text{, for all small $B\supseteq C$ and all $a,b$.}

Again, p(x)q(y)p(x)\otimes q(y) does not depend on the choice of CC. The product operation is associative:

(p(x)q(y))r(z)=p(x)(q(y)r(z)),(p(x)\otimes q(y))\otimes r(z)=p(x)\otimes(q(y)\otimes r(z)),

but commutativity

p(x)q(y)=?q(y)p(x)p(x)\otimes q(y)\stackrel{{\scriptstyle?}}{{=}}q(y)\otimes p(x)

can fail.

Remark 2.1.

If f,gf,g are definable functions and p,qp,q are definable types, then fpgq=(f×g)(pq)f_{*}p\otimes g_{*}q=(f\times g)_{*}(p\otimes q), where f×gf\times g sends (x,y)(x,y) to (f(x),g(y))(f(x),g(y)).

2.2 Generically stable types

Now assume that TT is NIP. (This includes the case of ACVF.)

Definition 2.2.

A definable type p(x)p(x) is generically stable if p(x)q(y)=q(y)p(x)p(x)\otimes q(y)=q(y)\otimes p(x) for every definable type q(y)q(y).

For other equivalent definitions of generic stability, see Section 3 of [4].

Definition 2.3.

Let ff be a CC-definable function and pp be a CC-definable type. Abusing terminology significantly, say that pp is dominated along ff if

f(a)fp|Bap|B for all small BC and all a.f(a)\models f_{*}p|B\implies a\models p|B\text{ for all small $B\supseteq C$ and all $a$.}

Note that the converse implication holds by definition of fpf_{*}p. Unlike the previous definitions, this does depend on the choice of CC. In the cases we care about, CC will be \emptyset.

Remark 2.4.

Suppose pp is dominated along ff, and qq is some other definable type. If BB is a set over which everything is defined and over which the domination holds, then

(f(a),b)fpq|B(a,b)pq|B(f(a),b)\models f_{*}p\otimes q|B\implies(a,b)\models p\otimes q|B (6)

We will use the following basic facts about generically stable types:

Theorem 2.5.
(a)

Products of generically stable types are generically stable.

(b)

Pushforwards of generically stable types are generically stable.

(c)

If pp is dominated along ff and fpf_{*}p is generically stable, then pp is generically stable.

(d)

If pp and qq are generically stable, dominated along ff and gg, respectively, then pqp\otimes q is dominated along f×gf\times g.

(e)

To check generic stability, it suffices to show that pp commutes with itself, i.e., p(x1)p(x2)=p(x2)p(x1)p(x_{1})\otimes p(x_{2})=p(x_{2})\otimes p(x_{1}).

Proof.
(a)

If pp and qq are generically stable, and rr is arbitrary, then pqr=prq=rpqp\otimes q\otimes r=p\otimes r\otimes q=r\otimes p\otimes q.

(b)

Suppose pp is generically stable, ff is a definable function, and qq is arbitrary. Then p(x)q(y)=q(y)p(x)p(x)\otimes q(y)=q(y)\otimes p(x). Pushing both sides forwards along (f×id)(f\times id) and applying Remark 2.1, we get that fp(x)q(y)=q(y)fp(x)f_{*}p(x^{\prime})\otimes q(y)=q(y)\otimes f_{*}p(x^{\prime}).

(c)

Let qq be another invariant type; we will show that p(x)q(y)=q(y)p(x)p(x)\otimes q(y)=q(y)\otimes p(x). Let BB be a set over which p,q,fp,q,f are defined. Let (b,a)(b,a) realize qp|Bq\otimes p|B. By Remark 2.1, (b,f(a))qfp|B(b,f(a))\models q\otimes f_{*}p|B. Since fpf_{*}p is generically stable, (f(a),b)fpq|B(f(a),b)\models f_{*}p\otimes q|B. By (6), (a,b)pq|B(a,b)\models p\otimes q|B. So pqp\otimes q and qpq\otimes p agree when restricted to the arbitrary set BB.

(d)

Let BB be a sufficiently big set. Suppose that (f(a),g(b))fpgq|B(f(a),g(b))\models f_{*}p\otimes g_{*}q|B. We need to show that (a,b)pq|B(a,b)\models p\otimes q|B. By (6), (a,g(b))pgq|B(a,g(b))\models p\otimes g_{*}q|B. By generic stability of pp, (g(b),a)gqp|B(g(b),a)\models g_{*}q\otimes p|B. By (6) again, (b,a)qp|B(b,a)\models q\otimes p|B. By generic stability again, (a,b)pq|B(a,b)\models p\otimes q|B.

(e)

Suppose p(x)p(x) commutes with itself, but p(x)q(y)q(y)p(x)p(x)\otimes q(y)\neq q(y)\otimes p(x). Choose some formula ϕ(x;y;c)\phi(x;y;c) which is in p(x)q(y)p(x)\otimes q(y) and not in q(y)p(x)q(y)\otimes p(x). We will prove that ϕ(x;y,z)\phi(x;y,z) has the independence property. Let nn be arbitrary. Let a1,,an,b,an+1,,a2na_{1},\ldots,a_{n},b,a_{n+1},\ldots,a_{2n} realize pnqpnp^{\otimes n}\otimes q\otimes p^{\otimes n} restricted to cc. Then ϕ(ai;b;c)in\models\phi(a_{i};b;c)\iff i\leq n, by choice of ϕ(x;y;c)\phi(x;y;c). The fact that pp commutes with itself implies that all permutations of (a1,,a2n)(a_{1},\ldots,a_{2n}) have the same type over cc. Therefore, for each permutation π\pi of {1,,2n}\{1,\ldots,2n\}, we can find a bπb_{\pi} such that ϕ(ai;bπ;c)\phi(a_{i};b_{\pi};c) holds iff π(i)n\pi(i)\leq n. It follows that for any S{1,,n}S\subseteq\{1,\ldots,n\}, we can find a bSb_{S} such that ϕ(ai;bS;c)\phi(a_{i};b_{S};c) holds if and only if iSi\in S. As nn was arbitrary, TT has the independence property, a contradiction.

2.3 Definable types in ACVF

Now work in ACVF. Recall that ACVF is NIP. We will make use of several definable types:

  • If BB is an open or closed ball in the home sort, then there is a complete type pB(x)p_{B}(x) over 𝕄\mathbb{M} which says that xBx\in B and xx is not in any strictly smaller balls. This type is called the generic type of BB. Completeness follows from CC-minimality. This type is definable, essentially because if BB^{\prime} is any other ball, then the formula xBx\in B^{\prime} is in pB(x)p_{B}(x) if and only if BBB^{\prime}\supseteq B. If CC is any set of parameters over which BB is defined, then pB|Cp_{B}|C says precisely that xx is in BB, and xx is not in any acleq(C)\operatorname{acl}^{eq}(C)-definable proper subball of BB.

  • There is also a type pk(x)p_{k}(x) which says that xx is in the residue field, and is not algebraic over 𝕄\mathbb{M}. This is called the generic type of the residue field, and is definable because kk is strongly minimal. If CC is any set of parameters, pk|Cp_{k}|C says precisely that xkx\in k and xacleq(C)x\notin\operatorname{acl}^{eq}(C).

  • The valuation ring 𝒪\mathcal{O} is a closed ball, so it has a generic type p𝒪p_{\mathcal{O}}. Over any set of parameters CC, p𝒪(x)p_{\mathcal{O}}(x) says that x𝒪x\in\mathcal{O}, and that xx is not in any acleq(C)\operatorname{acl}^{eq}(C)-definable proper subballs of 𝒪\mathcal{O}. Every proper subball of 𝒪\mathcal{O} is contained in a unique one of the form res1(α)\operatorname{res}^{-1}(\alpha), for αk\alpha\in k. Consequently, p𝒪|Cp_{\mathcal{O}}|C equivalently says that x𝒪x\in\mathcal{O} and that xres1(α)x\notin\operatorname{res}^{-1}(\alpha) for any αacleq(C)\alpha\in\operatorname{acl}^{eq}(C). Equivalently,

    xp𝒪|Cres(x)pk|Cx\models p_{\mathcal{O}}|C\iff\operatorname{res}(x)\models p_{k}|C

    Therefore, p𝒪p_{\mathcal{O}} is dominated along res\operatorname{res}, and resp𝒪=pk\operatorname{res}_{*}p_{\mathcal{O}}=p_{k}.

The type pkp_{k} is generically stable. To see this, use (e) of Theorem 2.5 and stability of kk.

Since pkp_{k} is generically stable, so is p𝒪p_{\mathcal{O}}, by Theorem 2.5(c). If BB is any closed ball, then there is an affine transformation f(x)=ax+bf(x)=ax+b sending 𝒪\mathcal{O} to BB, and pB=fp𝒪p_{B}=f_{*}p_{\mathcal{O}}. By Theorem 2.5(b), each pBp_{B} is generically stable.

Let p𝒪np_{\mathcal{O}^{n}} be p𝒪np_{\mathcal{O}}^{\otimes n}. We think of p𝒪np_{\mathcal{O}^{n}} as the generic type of the lattice 𝒪n\mathcal{O}^{n}. By Theorem 2.5, p𝒪np_{\mathcal{O}^{n}} is generically stable, and is dominated along the map (x1,,xn)(res(x1),,res(xn))(x_{1},\ldots,x_{n})\mapsto(\operatorname{res}(x_{1}),\ldots,\operatorname{res}(x_{n})). Also, the pushforward along this map is pknp_{k}^{\otimes n}, the generic type of knk^{n}.

The generic type of knk^{n} is stabilized by the action of GLn(k)GL_{n}(k), so by domination, the generic type of 𝒪n\mathcal{O}^{n} is stabilized by GLn(𝒪)GL_{n}(\mathcal{O}). In light of this, the following definition does not depend on the choice of gg:

Definition 2.6.

Let Λ\Lambda be a lattice in KnK^{n}. The generic type pΛp_{\Lambda} of Λ\Lambda is gp𝒪ng_{*}p_{\mathcal{O}^{n}}, where g:KnKng:K^{n}\to K^{n} is any linear map sending 𝒪n\mathcal{O}^{n} to Λ\Lambda.

Moreover, pΛp_{\Lambda} is Λ\ulcorner\Lambda\urcorner-definable. Note that pΛp_{\Lambda} is a generically stable type, because it is a pushforward of a generically stable type.

2.4 Left transitivity

Now return to an arbitrary theory TT. Work in TeqT^{eq}, so that acl\operatorname{acl} and dcl\operatorname{dcl} mean acleq\operatorname{acl}^{eq} and dcleq\operatorname{dcl}^{eq}.

Lemma 2.7.

Suppose CBC\subseteq B are small sets and a1,a2a_{1},a_{2} are tuples (possibly infinite, but small). If tp(a1/B)\operatorname{tp}(a_{1}/B) is CC-definable and tp(a2/Ba1)\operatorname{tp}(a_{2}/Ba_{1}) is Ca1Ca_{1}-definable, then tp(a2a1/B)\operatorname{tp}(a_{2}a_{1}/B) is CC-definable.

Proof.

Naming the parameters from CC, we may assume C=C=\emptyset. Let ϕ(x2,x1;y)\phi(x_{2},x_{1};y) be a formula; we must produce a ϕ\phi-definition (over \emptyset) for tp(a2a1/B)\operatorname{tp}(a_{2}a_{1}/B). Since tp(a2/Ba1)\operatorname{tp}(a_{2}/Ba_{1}) is a1a_{1}-definable, the ϕ(x2;x1,y)\phi(x_{2};x_{1},y)-type of a2a_{2} over Ba1Ba_{1} has a definition ψ(y,a1)\psi(y,a_{1}). In particular, for every tuple bb from BB,

ϕ(a2,a1,b)ψ(b,a1).\models\phi(a_{2},a_{1},b)\leftrightarrow\psi(b,a_{1}).

Meanwhile, since tp(a1/B)\operatorname{tp}(a_{1}/B) is 0-definable, there is a formula χ(y)\chi(y) such that for every bb in BB,

ψ(b,a1)χ(b).\models\psi(b,a_{1})\leftrightarrow\chi(b).

Thus, for every bb in BB,

ϕ(a2,a1,b)χ(b),\models\phi(a_{2},a_{1},b)\leftrightarrow\chi(b),

so χ(y)\chi(y) is the ϕ(x2,x1;y)\phi(x_{2},x_{1};y)-definition of tp(a2a1/B)\operatorname{tp}(a_{2}a_{1}/B). ∎

Lemma-Definition 2.8.

The following are equivalent for A,B,CA,B,C small sets of parameters.

  1. 1.

    The *-type tp(A/BC)\operatorname{tp}(A/BC) has a global CC-definable extension.

  2. 2.

    For every small set of parameters DD, there is a CC-definable extension of tp(A/BC)\operatorname{tp}(A/BC) to a *-type over BCDBCD.

  3. 3.

    For every small set of parameters DD, there is DBCDD^{\prime}\equiv_{BC}D such that tp(A/BCD)\operatorname{tp}(A/BCD^{\prime}) is CC-definable.

  4. 4.

    For some small model MBCM\supseteq BC, tp(A/M)\operatorname{tp}(A/M) is CC-definable.

We denote these equivalent conditions by A|CdefBA\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{def}}_{C}B.

Proof.
(1\implies2)

The restriction of a global CC-definable type to BCDBCD is CC-definable.

(2\implies3)

Given DD, (2) implies that there is ABCAA^{\prime}\equiv_{BC}A such that tp(A/BCD)\operatorname{tp}(A^{\prime}/BCD) is CC-definable. Choose DD^{\prime} such that ADBCADA^{\prime}D\equiv_{BC}AD^{\prime}. Then tp(A/BCD)\operatorname{tp}(A/BCD^{\prime}) is CC-definable and DBCDD^{\prime}\equiv_{BC}D.

(3\implies4)

Applying (3) to DD a small model containing BCBC, we get a small model DD^{\prime} containing BCBC such that tp(A/BCD)=tp(A/D)\operatorname{tp}(A/BCD^{\prime})=\operatorname{tp}(A/D^{\prime}) is CC-definable.

(4\implies1)

CC-definable types over models have unique CC-definable extensions to elementary extensions. This is true even for *-types.

Lemma 2.9.

If a1|CdefBa_{1}\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{def}}_{C}B and a2|Ca1defBa_{2}\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{def}}_{Ca_{1}}B then a2a1|CdefBa_{2}a_{1}\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{def}}_{C}B, so |def\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{def}} satisfies left-transitivity.

Proof.

We use condition (3) of Lemma-Definition 2.8. Let DD be a small set of parameters. Since a1|CdefBa_{1}\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{def}}_{C}B, there is DBCDD^{\prime}\equiv_{BC}D such that tp(a1/BCD)\operatorname{tp}(a_{1}/BCD^{\prime}) is CC-definable. As a2|Ca1defBa_{2}\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{def}}_{Ca_{1}}B there is D′′Ca1BDD^{\prime\prime}\equiv_{Ca_{1}B}D^{\prime} such that tp(a2/BCa1D′′)\operatorname{tp}(a_{2}/BCa_{1}D^{\prime\prime}) is Ca1Ca_{1}-definable. Note that tp(a1/BCD′′)\operatorname{tp}(a_{1}/BCD^{\prime\prime}) is CC-definable. By Lemma 2.7, it follows that tp(a2a1/BCD′′)\operatorname{tp}(a_{2}a_{1}/BCD^{\prime\prime}) is CC-definable. Since D′′BCDBCDD^{\prime\prime}\equiv_{BC}D^{\prime}\equiv_{BC}D, we have verified a2a1|CdefBa_{2}a_{1}\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{def}}_{C}B using condition (3). ∎

Definition 2.10.

If A,B,CA,B,C are small sets of parameters, we will write A|CadefBA\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{adef}}_{C}B to mean A|acl(C)defBA\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{def}}_{\operatorname{acl}(C)}B. (Recall that acl(C)\operatorname{acl}(C) means acleq(C)\operatorname{acl}^{eq}(C).)

In other words, a|CadefBa\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{adef}}_{C}B if tp(a/BC)\operatorname{tp}(a/BC) can be extended to a type which is almost CC-definable, that is, acl(C)\operatorname{acl}(C)-definable. In a stable theory, |adef\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{adef}} is exactly nonforking independence.

Lemma 2.11.

If A|CadefBA\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{adef}}_{C}B, then acl(AC)|CadefB\operatorname{acl}(AC)\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{adef}}_{C}B.

Proof.

Replacing CC with acl(C)\operatorname{acl}(C), we may assume that C=acl(C)C=\operatorname{acl}(C). By condition (4) of Lemma-Definition 2.8, there is a small model MM containing BCBC such that tp(A/M)\operatorname{tp}(A/M) is CC-definable. We need to show that tp(acl(AC)/M)\operatorname{tp}(\operatorname{acl}(AC)/M) is CC-definable. This is equivalent to showing that for each acl(AC)\operatorname{acl}(AC)-definable set XX, there is some CC-definable set XX^{\prime} such that XM=XMX\cap M=X^{\prime}\cap M.

Given such an XX, let X1,,XnX_{1},\ldots,X_{n} be the conjugates of XX over ACAC. Let EE be the equivalence relation

xEyi=1n(xXiyXi)xEy\leftrightarrow\bigwedge_{i=1}^{n}(x\in X_{i}\leftrightarrow y\in X_{i})

Then EE is ACAC-definable. Since tp(AC/M)\operatorname{tp}(AC/M) is CC-definable, the restriction E=EME^{\prime}=E\cap M of EE to MM is CC-definable. Since EE has finitely many equivalence classes, so does EE^{\prime}, and hence each equivalence class of EE^{\prime} is CC-definable, as acl(C)=C\operatorname{acl}(C)=C. But XMX\cap M is a union of finitely many EE^{\prime}-equivalence classes, so XMX\cap M is CC-definable. ∎

Lemma 2.12.

If a1|CadefBa_{1}\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{adef}}_{C}B and a2|Ca1adefBa_{2}\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{adef}}_{Ca_{1}}B, then a2a1|CadefBa_{2}a_{1}\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{adef}}_{C}B, so |adef\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{adef}} satisfies left-transitivity.

Proof.

By Lemma 2.11, we know that acl(a1C)|acl(C)defB\operatorname{acl}(a_{1}C)\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{def}}_{\operatorname{acl}(C)}B. We are given a2|acl(Ca1)defBa_{2}\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{def}}_{\operatorname{acl}(Ca_{1})}B. Combining these using Lemma 2.9, we conclude that a2acl(Ca1)|acl(C)defBa_{2}\operatorname{acl}(Ca_{1})\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{def}}_{\operatorname{acl}(C)}B. This easily implies a2a1|acl(C)defBa_{2}a_{1}\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{def}}_{\operatorname{acl}(C)}B, as desired. ∎

3 An Abstract Criterion for Elimination of Imaginaries

We state a sufficient condition for a theory TT to have elimination of imaginaries, extracted from [3].

Theorem 3.1.

Let TT be a theory, with home sort KK (meaning 𝕄eq=dcleq(K)\mathbb{M}^{eq}=\operatorname{dcl}^{eq}(K)). Let 𝒢\mathcal{G} be some collection of sorts. If the following conditions all hold, then TT has elimination of imaginaries in the sorts 𝒢\mathcal{G}.

  • For every non-empty definable set XK1X\subseteq K^{1}, there is an acleq(X)\operatorname{acl}^{eq}(\ulcorner X\urcorner)-definable type in XX.

  • Every definable type in KnK^{n} has a code in 𝒢\mathcal{G} (possibly infinite). That is, if pp is any (global) definable type in KnK^{n}, then the set p\ulcorner p\urcorner of codes of the definitions of pp is interdefinable with some (possibly infinite) tuple from 𝒢\mathcal{G}.

  • Every finite set of finite tuples from 𝒢\mathcal{G} has a code in 𝒢\mathcal{G}. That is, if SS is a finite set of finite tuples from 𝒢\mathcal{G}, then S\ulcorner S\urcorner is interdefinable with a tuple from 𝒢\mathcal{G}.

Proof.

Assume the three conditions.

Claim 3.2.

For every non-empty definable set XKnX\subseteq K^{n}, there is an acleq(X)\operatorname{acl}^{eq}(\ulcorner X\urcorner)-definable type in XX.

Proof.

We proceed by induction on nn, the base case n=1n=1 being given. Suppose n>1n>1. Take XKnX\subseteq K^{n}. Let C=XC=\ulcorner X\urcorner. Let π:KnKn1\pi:K^{n}\twoheadrightarrow K^{n-1} be the projection onto the first n1n-1 coordinates. Then π(X)\pi(X) is CC-definable, so by induction, there is an acleq(C)\operatorname{acl}^{eq}(C)-definable type in π(X)\pi(X). Let a1a_{1} realize this type. Then a1|CadefCa_{1}\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{adef}}_{C}C.

Let Y={yK1|(a1,y)X}Y=\{y\in K^{1}|(a_{1},y)\in X\}, so YY is essentially Xπ1(a1)X\cap\pi^{-1}(a_{1}). Then YY is Ca1Ca_{1}-definable and non-empty. By assumption, there is an acleq(Ca1)\operatorname{acl}^{eq}(Ca_{1})-definable type in YY. Let a2a_{2} realize this type; then a2Ya_{2}\in Y and a2|Ca1adefCa_{2}\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{adef}}_{Ca_{1}}C. Since a1|CadefCa_{1}\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{adef}}_{C}C, it follows that a1a2|CadefCa_{1}a_{2}\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{adef}}_{C}C by Lemma 2.12. By definition of |adef\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{adef}}, there is an acleq(C)\operatorname{acl}^{eq}(C)-definable type p(x1,x2)p(x_{1},x_{2}) such that a1a2p|acleq(C)a_{1}a_{2}\models p|\operatorname{acl}^{eq}(C). As a2Ya_{2}\in Y, the tuple a1a2a_{1}a_{2} is in XX, so pp is an acleq(C)\operatorname{acl}^{eq}(C)-definable type in XX. ∎

Let ee be any imaginary. Then there is some nn and some 0-definable equivalence relation EE on KnK^{n} such that ee is a code for some EE-equivalence class XX. By the claim, there is an acleq(e)\operatorname{acl}^{eq}(e)-definable type pp in XX. Then edcleq(p)e\in\operatorname{dcl}^{eq}(\ulcorner p\urcorner), because XX is the unique EE-equivalence class in which the type p\ulcorner p\urcorner lives. By the second assumption, there is some small tuple t0𝒢t_{0}\subseteq\mathcal{G} such that p\ulcorner p\urcorner is interdefinable with t0t_{0}. Thus edcleq(t0)e\in\operatorname{dcl}^{eq}(t_{0}) and t0acleq(e)t_{0}\in\operatorname{acl}^{eq}(e). By compactness, we can find some finite tuple tt from 𝒢\mathcal{G} such that edcleq(t)e\in\operatorname{dcl}^{eq}(t) and tacleq(e)t\in\operatorname{acl}^{eq}(e). Write ee as f(t)f(t) for some 0-definable function ff. Let SS be the (finite) set of conjugates of tt over ee. Then SS is ee-definable. Moreover, f(t)=ef(t^{\prime})=e for any tSt^{\prime}\in S, so ee is S\ulcorner S\urcorner-definable. Hence ee and S\ulcorner S\urcorner are interdefinable. By the third hypothesis, S\ulcorner S\urcorner has a code in 𝒢\mathcal{G}. So ee has a code in 𝒢\mathcal{G}. As ee was arbitrary, TT has elimination of imaginaries down to the sorts in 𝒢\mathcal{G}. ∎

The conditions in the theorem are sufficient but not necessary for elimination of imaginaries to hold. Namely, the first condition has nothing to do with 𝒢\mathcal{G}, and happens to fail in p\mathbb{Q}_{p}, even if 𝒢\mathcal{G} is chosen to be all of peq\mathbb{Q}_{p}^{eq}.

3.1 Examples

We sketch how to use Theorem 3.1 to verify elimination of imaginaries in ACF and DCF0 in the home sort KK (so 𝒢\mathcal{G} is merely {K}\{K\}). The first condition follows from stability: if XX is any non-empty definable set, then the formula xXx\in X does not fork over X\ulcorner X\urcorner. If pp is a global type which does not fork over X\ulcorner X\urcorner and contains this formula, then pp is an acleq(X)\operatorname{acl}^{eq}(\ulcorner X\urcorner)-definable type in XX.

For the second condition, one must check that every type has a code (possibly infinite) in the home sort. If pp is a type in KnK^{n}, then there is a minimal Zariski-closed or Kolchin-closed set VV containing pp, and pp and VV have the same code. The second condition thus reduces to coding Zariski-closed sets or Kolchin-closed sets, respectively. So does the third condition, since any finite subset of KnK^{n} is Zariski-closed and Kolchin-closed. Now, to code a Zariski-closed or Kolchin-closed set VV, we merely need to code the ideal II of polynomials or differential polynomials which vanish on VV. In the ACF case, this reduces to coding, for each d<ωd<\omega, the intersection of II with the space of degree d\leq d polynomials in K[X1,,Xn]K[X_{1},\ldots,X_{n}]. Something similar happens in DCF. So the problem reduces to coding linear subspaces of KmK^{m} for various mm.

But this is doable, by the following basic and general fact:

Lemma 3.3.

Let KK be any field. Let VV be a subspace of KnK^{n}. Then VV can be coded by a tuple in KK, and VV and Kn/VK^{n}/V have V\ulcorner V\urcorner-definable bases.

Proof.

Let m=dimVm=\dim V. By linear algebra, there is some coordinate projection π:KnKm\pi:K^{n}\to K^{m} such that the restriction of π\pi to VV is an isomorphism VKmV\to K^{m}. Then the preimage of the standard basis under this isomorphism is a V\ulcorner V\urcorner-definable basis for VV. This basis is a code for VV. Meanwhile, if we push the standard basis of KnK^{n} forward to Kn/VK^{n}/V, then some subset of this will be a basis for Kn/VK^{n}/V, and will be definable over the parameters (such as V\ulcorner V\urcorner) that were used to interpret the set Kn/VK^{n}/V. ∎

For the case of ACVF, the coding of definable types will be done similarly. But in addition to coding subspaces of KnK^{n}, we will also need to code definable ways of turning KnK^{n} into a valued KK-vector space. The third condition of Theorem 3.1 will be verified using the coding of definable types.

4 Elimination of imaginaries in ACVF

In this section, we prove that ACVF has elimination of imaginaries in the sorts KK, Rn,R_{n,\ell}, by applying Theorem 3.1. Recall that we are referring to these as the geometric sorts. We say that an object has a geometric code if it has a code in these sorts.

4.1 Coding modules

Recall that Rn,R_{n,\ell} is the set of pairs (Λ,V)(\Lambda,V) where Λ\Lambda is a lattice in KnK^{n} and VV is an \ell-dimensional subspace of resΛ:=Λ𝒪k=Λ/𝔐Λ\operatorname{res}\Lambda:=\Lambda\otimes_{\mathcal{O}}k=\Lambda/\mathfrak{M}\Lambda.

For fixed Λ\Lambda, the poset of kk-subspaces of resΛ\operatorname{res}\Lambda is isomorphic to the poset of 𝒪\mathcal{O}-submodules between 𝔐Λ\mathfrak{M}\Lambda and Λ\Lambda. Moreover, \ell-dimensional subspaces correspond exactly to 𝒪\mathcal{O}-submodules isomorphic to 𝒪×𝔐n\mathcal{O}^{\ell}\times\mathfrak{M}^{n-\ell}. So we could equivalently define Rn,R_{n,\ell} to be the set of all 𝒪\mathcal{O}-submodules of KnK^{n} isomorphic to 𝒪×𝔐n\mathcal{O}^{\ell}\times\mathfrak{M}^{n-\ell}. Under this identification,

=0nRn,\bigcup_{\ell=0}^{n}R_{n,\ell}

is the space of all open bounded definable 𝒪\mathcal{O}-submodules of KnK^{n}, by Theorem 1.6.

In section 1.4, we saw that ModnMod_{n}, the set of all definable submodules of KnK^{n}, is interpretable. We now show that ModnMod_{n} can be embedded into the geometric sorts.

Lemma 4.1.

If MKnM\leq K^{n} is a definable submodule of KnK^{n}, then MM has a geometric code.

Proof.

Let V+V^{+} be the KK-span of NN, and let VV^{-} be the maximal KK-subspace of KnK^{n} contained in NN. By Lemma 3.3, the subspaces V+V^{+} and VV^{-} can be coded by a tuple cc from KK, and the quotient V+/VV^{+}/V^{-} has a cc-definable identification with KmK^{m}, for some mm. Then NN is interdefinable over cc with the image of N/VN/V^{-} in KmK^{m}. But this image will be an open bounded definable submodule, so it is an element of Rm,R_{m,\ell} for some \ell. ∎

4.2 Coding definable types

A definable type in KnK^{n} induces an ideal II in K[X1,,Xn]K[X_{1},\ldots,X_{n}] together with the structure of a valued KK-vector space on the quotient K[X]/IK[\vec{X}]/I. By quantifier elimination in the one-sorted language, these data completely determine the type. So the problem of finding codes for definable types reduces to the (easy) problem of coding subspaces, and the problem of coding valued vector space structures on KK-vector spaces.333We will be more explicit in the proof of Theorem 4.5 below.

At the risk of being overly pedantic…

Definition 4.2.

Let VV be a KK-vector space. A VVS structure on VV is a binary relation RR on VV such that there is a valued KK-vector space structure (V,Γ(V),)(V,\Gamma(V),\ldots) on VV for which xRyval(x)val(y)xRy\iff\operatorname{val}(x)\leq\operatorname{val}(y).

The VVS structures on VV are essentially the distinct ways of turning VV into a valued KK-vector space. Two valued KK-vector spaces WW and WW^{\prime} with the same underlying vector space VV yield the same VVS structure on VV iff they are identical up to an isomorphism of value groups Γ(W)Γ(W)\Gamma(W)\cong\Gamma(W^{\prime}).

If VV is a definable KK-vector space, it makes sense to say that a VVS structure RR is “definable,” meaning that RR is a definable subset of V×VV\times V. If RR is definable, then Γ(V)\Gamma(V) is interpretable and the map val:VΓ(V)\operatorname{val}:V\to\Gamma(V) and the action of Γ(K)\Gamma(K) on Γ(V)\Gamma(V) are all definable.

Theorem 4.3.

Let τ\tau be (the code for) a definable VVS structure on KmK^{m}. Then τ\tau is interdefinable with an element of the geometric sorts.

Proof.

Let VV be the associated valued KK-vector space. So KmK^{m} is the underlying vector space of VV and τ\tau is a code for the relation val(x)val(y)\operatorname{val}(x)\leq\operatorname{val}(y). The set Γ(V)\Gamma(V) is τ\tau-interpretable, as Km0K^{m}\setminus 0 modulo the equivalence relation val(x)val(y)val(y)val(x)\operatorname{val}(x)\leq\operatorname{val}(y)\wedge\operatorname{val}(y)\leq\operatorname{val}(x).

By Remark 1.2, Γ(V)\Gamma(V) consists of finitely many orbits under Γ(K)\Gamma(K).

If there was only one orbit, and if there was a canonical identification of Γ(V)\Gamma(V) with Γ(K)\Gamma(K), we could proceed as follows: let BB be the closed ball around 0 with valuative radius 0. The other closed balls around 0 are all of the form αB\alpha B, for αK\alpha\in K. The set BB is a definable 𝒪\mathcal{O}-submodule of KmK^{m}, so it has a geometric code. It determines τ\tau, however, because val(x)val(y)\operatorname{val}(x)\leq\operatorname{val}(y) if and only if every closed ball containing 0 and xx contains yy. So τ\tau and B\ulcorner B\urcorner would be interdefinable.

In general we have several orbits. The first order of business is finding a τ\tau-definable element in each one:

Claim 4.4.

Each orbit of Γ(K)\Gamma(K) on Γ(V)\Gamma(V) contains a τ\tau-definable element.

Proof.

For x,yΓ(V)x,y\in\Gamma(V), let xyx\gg y indicate that Γ(K)+x>Γ(K)+y\Gamma(K)+x>\Gamma(K)+y. Let xyx\sim y indicate that x≫̸yx\not\gg y and y≫̸xy\not\gg x. This is an equivalence relation. Each orbit of Γ(K)\Gamma(K) is in one \sim-equivalence class, so there are finitely many \sim-equivalence classes. Each \sim-equivalence class is a convex subset of the linear order Γ(V)\Gamma(V). Let C1>C2>>CC_{1}>C_{2}>\ldots>C_{\ell} be the distinct \sim-equivalence classes sorted in order from most positive to most negative.

For 0i0\leq i\leq\ell, let ViV_{i} be the set of vVv\in V such that val(v)Cj\operatorname{val}(v)\in C_{j} for some jij\leq i. Each ViV_{i} is a KK-vector space, yielding an ascending filtration

0=V0V1V=V.0=V_{0}\subseteq V_{1}\subseteq\cdots\subseteq V_{\ell}=V.

On ViVi1V_{i}\setminus V_{i-1}, the function val\operatorname{val} lands in CiC_{i} and factors through the quotient Vi/Vi1V_{i}/V_{i-1}, by the ultrametric inequality.

The equivalence relation \sim is τ\tau-definable. Since there are finitely many equivalence classes and they are totally ordered, each CiC_{i} is τ\tau-definable. Consequently, each ViV_{i} is τ\tau-definable. By Lemma 3.3, each quotient space Vi/Vi1V_{i}/V_{i-1} has a τ\tau-definable basis. In particular, there is a τ\tau-definable non-zero vector in Vi/Vi1V_{i}/V_{i-1}. Taking its valuation, we get a τ\tau-definable element of CiC_{i}. We have shown:

Each CiC_{i} contains a τ\tau-definable element. (7)

Next, for x,yΓ(V)x,y\in\Gamma(V) let xyx\approx y indicate that x+γ>y>xγx+\gamma>y>x-\gamma for all positive γΓ(K)\gamma\in\Gamma(K). This is again a τ\tau-definable equivalence relation. If xx and yy are in the same orbit, but are not equal, then xyx\not\approx y. Consequently, each \approx-equivalence class contains at most one point from each orbit, so each \approx-equivalence class is finite. This implies that if xyx\approx y, then xx and yy are interalgebraic over τ\tau. In light of the total ordering, they are actually interdefinable over τ\tau.

Let xx be an arbitrary element of Γ(V)\Gamma(V). We will show that the orbit Γ(K)+x\Gamma(K)+x contains a τ\tau-definable element. By (7), xyx\sim y for some τ\tau-definable element yy. The set

{γΓ(K):γ+xy}\{\gamma\in\Gamma(K):\gamma+x\leq y\}

is non-empty, because x≫̸yx\not\gg y, and it is bounded above, because y≫̸xy\not\gg x. It is also definable, so it has a supremum γ0\gamma_{0}, by o-minimality of Γ(K)\Gamma(K). Then γ0+xy\gamma_{0}+x\approx y. The element γ0+x\gamma_{0}+x is interdefinable over τ\tau with the τ\tau-definable element yy, so it is itself τ\tau-definable. ∎

Given the claim, let γ1,,γn\gamma_{1},\ldots,\gamma_{n} be a set of τ\tau-definable orbit representatives. Let Bi={vKm:val(v)γi}B_{i}=\{v\in K^{m}:\operatorname{val}(v)\geq\gamma_{i}\}. Each BiB_{i} is a τ\tau-definable 𝒪\mathcal{O}-submodule of KmK^{m}, i.e., an element of ModmMod_{m}. The closed balls of VV containing 0 are exactly the sets of the form αBi\alpha B_{i} for 1in1\leq i\leq n and αK\alpha\in K. The family of closed balls containing 0 is enough to determine the VVS structure, so τ\tau is interdefinable with the tuple (B1,,Bn)(B_{1},\ldots,B_{n}). But by Lemma 4.1, each BiB_{i} has a geometric code. ∎

Theorem 4.5.

Let p(x)p(x) be a definable type in KnK^{n}. Then p(x)p(x) has a code in the geometric sorts.

Proof.

For each dd, let VdV_{d} be the space of polynomials in K[X1,,Xn]K[X_{1},\ldots,X_{n}] of degree d\leq d. This is a finite dimensional definable KK-vector space with a 0-definable basis. Let IdI_{d} be the set of Q(X)VdQ(X)\in V_{d} such that the formula Q(x)=0Q(x)=0 is in p(x)p(x). Let RdR_{d} be the set of pairs (Q1(X),Q2(X))(Q_{1}(X),Q_{2}(X)) in Vd×VdV_{d}\times V_{d} such that the formula val(Q1(x))val(Q2(x))\operatorname{val}(Q_{1}(x))\leq\operatorname{val}(Q_{2}(x)) is in p(x)p(x). Then IdI_{d} is a subspace of VdV_{d}, and RdR_{d} induces a definable VVS structure on the quotient space Vd/IdV_{d}/I_{d}. Quantifier elimination in the one-sorted language implies that pp is completely determined by the collection of all IdI_{d}’s and RdR_{d}’s for d<ωd<\omega. By Lemma 3.3, we can find codes Id\ulcorner I_{d}\urcorner in the home sort for the IdI_{d}’s. After naming these codes, each quotient space Vd/IdV_{d}/I_{d} has a definable basis, and can be definably identified with some power of KK. Then each RdR_{d} is interdefinable with a definable VVS structure on a power of KK. By Theorem 4.3, these VVS structures have codes cdc_{d} in the geometric sorts. Now the union of all the Id\ulcorner I_{d}\urcorner’s and cdc_{d}’s is a code for pp. ∎

4.3 Coding finite sets

In this section, we show that finite sets of tuples from the geometric sorts can be coded in the geometric sorts. We make use of the existence of geometric codes for definable types. For a more elementary but more complicated approach, see Proposition 3.4.1 in [1].

Definition 4.6.

If XX is some set, SymnX\operatorname{Sym}^{n}X will denote the nn-fold symmetric product of XX, that is, XnX^{n} modulo the action of the nnth symmetric group. The natural map XnSymnXX^{n}\to\operatorname{Sym}^{n}X will be denoted σ\sigma, so that

σ(x1,,xn)=σ(y1,,yn)\sigma(x_{1},\ldots,x_{n})=\sigma(y_{1},\ldots,y_{n})

if and only if there is a permutation π\pi of nn such that xi=yπ(i)x_{i}=y_{\pi(i)} for i=1,,ni=1,\ldots,n.

Definition 4.7.

A 0-definable map π:XY\pi:X\to Y has definable lifting if for every bYb\in Y there is a bb-definable type pbp_{b} in π1(b)\pi^{-1}(b). (In particular, π\pi must be surjective.) Say that π\pi has generically stable lifting if it has definable lifting and pbp_{b} can be taken to be generically stable.

In both cases, we can easily modify the map bpbb\mapsto p_{b} to be automorphism equivariant, so that if σAut(𝕄/)\sigma\in\operatorname{Aut}(\mathbb{M}/\emptyset), then pσ(b)=σ(pb)p_{\sigma(b)}=\sigma(p_{b}) for every bb. Conversely, if there is an automorphism equivariant map bpbb\mapsto p_{b} from elements of YY to definable (resp. generically stable) types in XX, such that pbp_{b} is in π1(b)\pi^{-1}(b), then π\pi has definable (resp. generically stable) lifting—the automorphism invariance ensures that pbp_{b} is bb-definable.

If π:XY\pi:X\to Y has definable lifting, and qq is a CC-definable type in YY for some parameters CC, then q=πpq=\pi_{*}p for some CC-definable type pp in XX. Indeed, if MM is a model containing CC, and bb realizes q|Mq|M, and aπ1(b)a\in\pi^{-1}(b) realizes pb|Mbp_{b}|Mb, then a|CbdefMa\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{def}}_{Cb}M and b|CdefMb\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{def}}_{C}M, so a|CdefMa\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{def}}_{C}M by left-transitivity of |def\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{def}}. Thus ap|Ma\models p|M for some CC-definable type pp. By definition of pushforward, π(a)=b\pi(a)=b realizes πp|M\pi_{*}p|M. Then the CC-definable types πp\pi_{*}p and qq have the same restriction to a model containing CC, so they are equal.

From this, we draw the following conclusion:

Observation 4.8.

If π:XY\pi:X\to Y has definable lifting, and definable types in XX have codes in the geometric sorts, then so do definable types in YY.

Indeed, let qq be a definable type in YY. Then qq is q\ulcorner q\urcorner-definable, so q=πpq=\pi_{*}p for some q\ulcorner q\urcorner-definable type pp. But then pp and qq are interdefinable, and pp has a geometric code.

We make two other useful observations:

Observation 4.9.

If π:XY\pi:X\to Y and π:XY\pi^{\prime}:X^{\prime}\to Y^{\prime} both have definable lifting, then so does the product map π×π:X×XY×Y\pi\times\pi^{\prime}:X\times X^{\prime}\to Y\times Y^{\prime}. Indeed, if ff and ff^{\prime} are the automorphism-equivariant maps witnessing definable lifting, then (b,b)f(b)f(b)(b,b^{\prime})\mapsto f(b)\otimes f^{\prime}(b^{\prime}) witnesses definable lifting for the product map π×π\pi\times\pi^{\prime}. The same statement holds with “generically stable lifting” in place of “definable lifting,” by Theorem 2.5(a).

Observation 4.10.

If π:XY\pi:X\to Y has generically stable lifting, then so does Symnπ:SymnXSymnY\operatorname{Sym}^{n}\pi:\operatorname{Sym}^{n}X\to\operatorname{Sym}^{n}Y. Indeed, suppose that bpbb\mapsto p_{b} is the automorphism equivariant map from elements of YY to generically stable types in XX. Then

σ(b1,,bn)σ(pb1pbn)\sigma(b_{1},\ldots,b_{n})\mapsto\sigma_{*}(p_{b_{1}}\otimes\cdots\otimes p_{b_{n}})

is a well-defined automorphism-equivariant map from elements of SymnY\operatorname{Sym}^{n}Y to generically stable types in SymnX\operatorname{Sym}^{n}X, witnessing generically stable lifting for Symnπ\operatorname{Sym}^{n}\pi. Generic stability ensures that

σ(pb1pbn)=σ(pbπ(1)pbπ(n))\sigma_{*}(p_{b_{1}}\otimes\cdots\otimes p_{b_{n}})=\sigma_{*}(p_{b_{\pi(1)}}\otimes\cdots\otimes p_{b_{\pi(n)}})

for any b1,,bnb_{1},\ldots,b_{n} and any permutation π\pi of {1,,n}\{1,\ldots,n\}.

Lemma 4.11.

Definable lifting holds for the map Symn(Km×𝒪)Symn(Km×k)\operatorname{Sym}^{n}(K^{m}\times\mathcal{O}^{\ell})\to\operatorname{Sym}^{n}(K^{m}\times k^{\ell}), induced by the componentwise residue 𝒪k\mathcal{O}^{\ell}\to k^{\ell}.

Proof.

First note that the residue map res:𝒪k\operatorname{res}:\mathcal{O}\to k has definable lifting: to each residue αk\alpha\in k we associate the generic type of the open ball res1(α)\operatorname{res}^{-1}(\alpha). By Observation 4.9, the componentwise residue map 𝒪nkn\mathcal{O}^{n}\to k^{n} has definable lifting. It follows that the map Symn𝒪Symnk\operatorname{Sym}^{n}\mathcal{O}\to\operatorname{Sym}^{n}k has definable lifting. Indeed, there is a commutative diagram

Symn𝒪\textstyle{\operatorname{Sym}^{n}\mathcal{O}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sim}𝒪n\textstyle{\mathcal{O}^{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Symnk\textstyle{\operatorname{Sym}^{n}k\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sim}kn\textstyle{k^{n}}

in which the vertical maps are induced by the residue map 𝒪k\mathcal{O}\to k, and the horizontal maps are induced by the elementary symmetric polynomials. These horizontal maps are bijections because kk and KK are algebraically closed and 𝒪\mathcal{O} is integrally closed. So definable lifting holds for Symn𝒪Symnk\operatorname{Sym}^{n}\mathcal{O}\to\operatorname{Sym}^{n}k.

Now consider the map Symn(Km×𝒪)Symn(Km×k)\operatorname{Sym}^{n}(K^{m}\times\mathcal{O}^{\ell})\to\operatorname{Sym}^{n}(K^{m}\times k^{\ell}). Sweeping the set/multiset distinction under the rug, let SS be a finite subset of Km×kK^{m}\times k^{\ell}. Let TkT\subseteq k be the set of all elements of kk appearing in SS. By definable lifting of Symi𝒪Symik\operatorname{Sym}^{i}\mathcal{O}\to\operatorname{Sym}^{i}k, there is a finite subset T𝒪T^{\prime}\subseteq\mathcal{O} such that T|SdefS\ulcorner T^{\prime}\urcorner\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{def}}_{\ulcorner S\urcorner}\ulcorner S\urcorner and such that TT^{\prime} lifts TT, meaning that the residue map 𝒪k\mathcal{O}\to k restricts to a bijection TTT^{\prime}\to T. Let f:TTf:T\to T^{\prime} be the inverse, a partial section of the residue map. Let SKm×𝒪S^{\prime}\subseteq K^{m}\times\mathcal{O}^{\ell} be obtained by applying ff componentwise to SS. Then SS^{\prime} maps onto SS via the componentwise residue map Km×𝒪Km×kK^{m}\times\mathcal{O}^{\ell}\to K^{m}\times k^{\ell}. Moreover, SS^{\prime} is definable over ST\ulcorner S\urcorner\ulcorner T^{\prime}\urcorner, and so

S|SdefS.\ulcorner S^{\prime}\urcorner\operatorname*{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}^{\textrm{def}}_{\ulcorner S\urcorner}\ulcorner S\urcorner.

Therefore SSymn(Km×𝒪)\ulcorner S^{\prime}\urcorner\in\operatorname{Sym}^{n}(K^{m}\times\mathcal{O}^{\ell}) lifts SSymn(Km×k)\ulcorner S\urcorner\in\operatorname{Sym}^{n}(K^{m}\times k^{\ell}), and realizes an S\ulcorner S\urcorner-definable type. ∎

Corollary 4.12.

Definable types in Symn(Km×k)\operatorname{Sym}^{n}(K^{m}\times k^{\ell}) have geometric codes.

Proof.

The set Symn(Km×𝒪)\operatorname{Sym}^{n}(K^{m}\times\mathcal{O}^{\ell}) embeds into Symn(Km+)\operatorname{Sym}^{n}(K^{m+\ell}), which 0-definably embeds into some power KNK^{N} by elimination of imaginaries in ACF. By Theorem 4.5, definable types in KNK^{N} have geometric codes. Therefore, definable types in Symn(Km×𝒪)\operatorname{Sym}^{n}(K^{m}\times\mathcal{O}^{\ell}) have geometric codes. By Observation 4.9 and Lemma 4.11, the Corollary follows. ∎

Theorem 4.13.

Let GG be any geometric sort. Then elements of SymnG\operatorname{Sym}^{n}G have geometric codes.

Proof.

We claim that there is some map GGG^{\prime}\to G with generically stable lifting, such that GG^{\prime} embeds (0-definably) into a product Km×kK^{m}\times k^{\ell}. Assuming this is true, we get a map

Symn(G)Symn(G)\operatorname{Sym}^{n}(G^{\prime})\to\operatorname{Sym}^{n}(G)

with generically stable (hence definable) lifting, and Symn(G)\operatorname{Sym}^{n}(G^{\prime}) embeds into Symn(Km×k)\operatorname{Sym}^{n}(K^{m}\times k^{\ell}). Corollary 4.12 ensures that definable types in Symn(G)\operatorname{Sym}^{n}(G^{\prime}) have geometric codes, so by definable lifting, definable types in Symn(G)\operatorname{Sym}^{n}(G) have geometric codes. In particular, looking at constant types in Symn(G)\operatorname{Sym}^{n}(G), we see that elements of Symn(G)\operatorname{Sym}^{n}(G) have geometric codes.

It remains to find GGG^{\prime}\to G. The property of generically stable lifting is closed under taking products (Observation 4.9), and GG is a product of KK’s and Rn,R_{n,\ell}’s, so it suffices to consider the case G=KG=K or G=Rn,G=R_{n,\ell}. For G=KG=K, we take the identity map G=KK=GG^{\prime}=K\to K=G. The case G=Rn,G=R_{n,\ell} remains. Let Rn,~\widetilde{R_{n,\ell}} be the set of triples (b,Λ,V)(\vec{b},\Lambda,V), where Λ\Lambda is a lattice in KnK^{n}, b\vec{b} is a lattice basis, and VV is an \ell-dimensional kk-subspace of resΛ:=Λ/𝔐Λ\operatorname{res}\Lambda:=\Lambda/\mathfrak{M}\Lambda. Then the canonical map

Rn,~Rn,\widetilde{R_{n,\ell}}\to R_{n,\ell}
(b,Λ,V)(Λ,V)(\vec{b},\Lambda,V)\mapsto(\Lambda,V)

has generically stable lifting. Indeed, given Λ\Lambda, any realization of pΛnp_{\Lambda}^{\otimes n} will be a basis of Λ\Lambda, where pΛp_{\Lambda} is the generic type of Λ\Lambda.

Moreover, we can embed Rn,~\widetilde{R_{n,\ell}} into a product of KK’s and kk’s. Let Grn,Gr_{n,\ell} denote the set of \ell-dimensional kk-subspaces of knk^{n}. Then there is a 0-definable map

Rn,~Kn2×Grn,\widetilde{R_{n,\ell}}\to K^{n^{2}}\times Gr_{n,\ell}
(b,Λ,V)(b,W),(\vec{b},\Lambda,V)\mapsto(\vec{b},W),

where WW is the image of VV under the identification of resΛ\operatorname{res}\Lambda with knk^{n} induced by the basis b\vec{b}. This map is an injection, and Grn,Gr_{n,\ell} can be embedded in a power of kk by algebraic geometry, or elimination of imaginaries in ACF. ∎

4.4 Putting everything together

Theorem 4.14.

ACVF has elimination of imaginaries in the geometric sorts, i.e., in KK and the Rn,R_{n,\ell}.

Proof.

This follows by Theorem 3.1. The second condition is Theorem 4.5. The third condition is Theorem 4.13. The first condition of Theorem 3.1 can be verified as follows: Let DD be a one-dimensional definable set. Then DD can be written as a disjoint union of acleq(D)\operatorname{acl}^{eq}(\ulcorner D\urcorner)-definable “swiss cheeses.” If B(B1Bn)B\setminus(B_{1}\cup\cdots\cup B_{n}) is one of these swiss cheeses, then the generic type pBp_{B} of BB is in B(B1Bn)B\setminus(B_{1}\cup\cdots\cup B_{n}), hence in DD. Since BB is acleq(D)\operatorname{acl}^{eq}(\ulcorner D\urcorner)-definable, so is pBp_{B}. ∎

5 Reduction to the standard geometric sorts

Haskell, Hrushovski, and Macpherson showed that ACVF has elimination of imaginaries in the sorts K,Sn,TnK,S_{n},T_{n}. To deduce this result from Theorem 4.14, we need to code the Rm,R_{m,\ell} sorts into the SnS_{n} and TnT_{n}. This is done in 2.6.4 of [1], but for the sake of completeness, we quickly recall the arguments here. Recall that SnS_{n} is the set of lattices in KnK^{n}, and TnT_{n} is the union of resΛ\operatorname{res}\Lambda as Λ\Lambda ranges over SnS_{n}.

First of all, we can easily code the Rm,R_{m,\ell} in terms of Rn,0(=Sn)R_{n,0}(=S_{n}) and Rn,1R_{n,1} (which is roughly a projectivized version of TnT_{n}). Indeed, if Λ\Lambda is a lattice in KnK^{n}, and VV is an \ell-dimensional subspace in res(V)\operatorname{res}(V), then VV can be coded by a one-dimensional subspace (namely V\bigwedge^{\ell}V) in

res(Λ)=res(Λ),\bigwedge^{\ell}\operatorname{res}(\Lambda)=\operatorname{res}(\bigwedge^{\ell}\Lambda),

and Λ\bigwedge^{\ell}\Lambda is a lattice in Kn\bigwedge^{\ell}K^{n}. So to code an element of Rn,R_{n,\ell}, we can use the underlying lattice in Rn,0R_{n,0}, and then an element in RN,1R_{N,1}, where N=dimKnN=\dim\bigwedge^{\ell}K^{n}.

Now to code an element of Rn,1R_{n,1} in terms of the SnS_{n} and TnT_{n}, we proceed by induction. Let Λ\Lambda be a lattice in KnK^{n}. Let π\pi be the projection onto the first coordinate. Then π(Λ)\pi(\Lambda) is free444Finitely generated torsion-free 𝒪\mathcal{O}-modules are always free., so we have a split exact sequence

0ΛΛπ(Λ)00\to\Lambda^{\prime}\to\Lambda\to\pi(\Lambda)\to 0

where Λ\Lambda^{\prime} is a lattice in Kn1K^{n-1}. Since this sequence is split exact, it remains split exact after tensoring with kk. So

0res(Λ)res(Λ)res(π(Λ))00\to\operatorname{res}(\Lambda^{\prime})\to\operatorname{res}(\Lambda)\to\operatorname{res}(\pi(\Lambda))\to 0

is exact. Let VV be a one dimensional subspace of res(Λ)\operatorname{res}(\Lambda). If VV sits inside res(Λ)\operatorname{res}(\Lambda^{\prime}), then VV is interdefinable with a one-dimensional subspace of res(Λ)\operatorname{res}(\Lambda^{\prime}), so can be coded in the true geometric sorts by induction.

Otherwise, VV maps isomorphically onto the one-dimensional kk-space res(π(Λ))\operatorname{res}(\pi(\Lambda)). Then to code VV, it suffices to code the inverse map res(π(Λ))Vres(Λ)\operatorname{res}(\pi(\Lambda))\to V\hookrightarrow\operatorname{res}(\Lambda). But because all the 𝒪\mathcal{O}-modules in sight are free,

Homk(res(π(Λ)),res(Λ))=Homk(π(Λ)k,Λk)=Hom𝒪(π(Λ),Λ)k.\operatorname{Hom}_{k}(\operatorname{res}(\pi(\Lambda)),\operatorname{res}(\Lambda))=\operatorname{Hom}_{k}(\pi(\Lambda)\otimes k,\Lambda\otimes k)=\operatorname{Hom}_{\mathcal{O}}(\pi(\Lambda),\Lambda)\otimes k.

And Hom𝒪(π(Λ),Λ)\operatorname{Hom}_{\mathcal{O}}(\pi(\Lambda),\Lambda) is a lattice in HomK(K,Kn)Kn\operatorname{Hom}_{K}(K,K^{n})\cong K^{n}. So a map from res(π(Λ))\operatorname{res}(\pi(\Lambda)) to res(Λ)\operatorname{res}(\Lambda) can be coded by an element of TnT_{n}.

Acknowledgments.

The author would like to thank Tom Scanlon, Ehud Hrushovski, and the participants in the 2014 Berkeley Model Theory Seminar. Also, this material is based upon work supported by the National Science Foundation Graduate Research Fellowship under Grant No. DGE 1106400. Any opinions, findings, and conclusions or recommendations expressed in this material are those of the author and do not necessarily reflect the views of the National Science Foundation.

References

  • [1] Deirdre Haskell, Ehud Hrushovski, and Dugald Macpherson. Definable sets in algebraically closed valued fields: elimination of imaginaries. J. reine angew. Math., 597:175–236, 2006.
  • [2] Jan E. Holly. Canonical forms for definable subsets of algebraically closed and real closed valued fields. J. Symbolic Logic, 60(3):843–860, September 1995.
  • [3] Ehud Hrushovski. Imaginaries and definable types in algebraically closed valued fields. preprint, 2014. arXiv:1403.7326.
  • [4] Ehud Hrushovski and Anand Pillay. On NIP and invariant measures. J. Eur. Math. Soc., 13(4):1005–1061, 2011.
  • [5] Dugald Macpherson. Model theory of valued fields. http://www1.maths.leeds.ac.uk/Pure/staff/macpherson/modnetval4.pdf, May 2008.
  • [6] Anand Pillay. Model theory of algebraically closed fields. In Elisabeth Bouscaren, editor, Model Theory and Algebraic Geometry, pages 61–84. Springer, 1998.
  • [7] Abraham Robinson. Complete theories. Studies in Logic and the Foundations of Mathematics. North-Holland, 1956.