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

Algebraic and o-minimal flows beyond the cocompact case

Spencer Dembner and Hunter Spink
Abstract.

Let XnX\subset\mathbb{C}^{n} be an algebraic variety, and let Λn\Lambda\subset\mathbb{C}^{n} be a discrete subgroup whose real and complex spans agree. We describe the topological closure of the image of XX in n/Λ\mathbb{C}^{n}/\Lambda, thereby extending a result of Peterzil-Starchenko in the case when Λ\Lambda is cocompact.

We also obtain a similar extension when XnX\subset\mathbb{R}^{n} is definable in an o-minimal structure with no restrictions on Λ\Lambda, and as an application prove the following conjecture of Gallinaro: for a closed semi-algebraic XnX\subset\mathbb{C}^{n} (such as a complex algebraic variety) and exp:n()n\exp:\mathbb{C}^{n}\to(\mathbb{C}^{*})^{n} the coordinate-wise exponential map, we have exp(X)¯=exp(X)i=1mexp(Ci)𝕋i\overline{\exp(X)}=\exp(X)\cup\bigcup_{i=1}^{m}\exp(C_{i})\cdot\mathbb{T}_{i} where 𝕋i()n\mathbb{T}_{i}\subset(\mathbb{C}^{*})^{n} are positive-dimensional compact real tori and CinC_{i}\subset\mathbb{C}^{n} are semi-algebraic.

1. Introduction

Let VV be a vector space over \mathbb{R} or \mathbb{C}, ΛV\Lambda\subset V a discrete subgroup, and XVX\subset V a closed subset. Let π:VV/Λ\pi\colon V\to V/\Lambda denote the projection map. Ullmo-Yafaev [5] observed that as RR\to\infty, the closures

π({xX:xR})¯\overline{\pi(\{x\in X:\|x\|\geq R\})}

form what we may call a “flow” of sets, which converge to a closed limiting set Fl(X){\operatorname{Fl}}(X) exactly containing those cosets v+Λv+\Lambda which an unbounded sequence of points xiXx_{i}\in X can approach in Euclidean distance. Evidently we have

π(X)¯=π(X)Fl(X).\overline{\pi(X)}=\pi(X)\cup{\operatorname{Fl}}(X).

In the case where XnX\subset\mathbb{C}^{n} is a complex curve and Λn\Lambda\subset\mathbb{C}^{n} is cocompact, [5, Theorem 2.4] shows Fl(X){\operatorname{Fl}}(X) consists of finitely many translates of real subtori 𝕋n/Λ\mathbb{T}\subset\mathbb{C}^{n}/\Lambda. This was later extended to arbitrary discrete subgroups by Dinh-Vu [2].

When XX is a higher-dimensional variety, an example of Peterzil-Starchenko [4, Section 8] shows this description of Fl(X){\operatorname{Fl}}(X) doesn’t hold, even if Λ\Lambda is cocompact. However, they show that in this cocompact case, replacing translated subtori by algebraic families of translated subtori is enough [4, Theorem 1.1].

Our main result shows the cocompactness restriction can be removed when the real and complex spans of Λ\Lambda agree. For 𝔽=\mathbb{F}=\mathbb{R} or \mathbb{C} and SS a subset of an 𝔽\mathbb{F}-vector space VV, let 𝔽S\mathbb{F}S denote its 𝔽\mathbb{F}-linear span.

Theorem 1.1.

Let XnX\subset\mathbb{C}^{n} be an algebraic variety, LnL\subset\mathbb{C}^{n} a \mathbb{C}-subspace, Λn\Lambda\subset\mathbb{C}^{n} a discrete subgroup with Λ=Λ=L\mathbb{R}\Lambda=\mathbb{C}\Lambda=L, and π:nn/Λ\pi:\mathbb{C}^{n}\to\mathbb{C}^{n}/\Lambda the quotient map. Then there are algebraic varieties C1,,CmnC_{1},\ldots,C_{m}\subset\mathbb{C}^{n} with dimCi<dimX\dim_{\mathbb{C}}C_{i}<\dim_{\mathbb{C}}X and positive-dimensional compact real tori 𝕋1,,𝕋mL/Λ\mathbb{T}_{1},\ldots,\mathbb{T}_{m}\subset L/\Lambda such that

Fl(X)=i=1m(π(Ci)+𝕋i).{\operatorname{Fl}}(X)=\bigcup_{i=1}^{m}(\pi(C_{i})+\mathbb{T}_{i}).

In particular, π(X)¯=π(X)i=1m(π(Ci)+𝕋i)\overline{\pi(X)}=\pi(X)\cup\bigcup_{i=1}^{m}(\pi(C_{i})+\mathbb{T}_{i}).

Remark 1.2.

As in [4], we will be able to take C1,,CmC_{1},\ldots,C_{m} to only depend on XX and LL, and take positive-dimensional \mathbb{C}-subspaces V1,,VmLV_{1},\ldots,V_{m}\subset L only depending on XX and LL such that 𝕋i=π(Vi)¯\mathbb{T}_{i}=\overline{\pi(V_{i})}.

When Λ\Lambda is cocompact, [4] also shows that we may assume CiC_{i} is a finite collection of points if 𝕋i\mathbb{T}_{i} is inclusion-maximal among 𝕋1,,𝕋m\mathbb{T}_{1},\ldots,\mathbb{T}_{m}, but this is false in this more general setting even for X=2X=\mathbb{C}^{2} and Λ=(i)×{0}\Lambda=(\mathbb{Z}\oplus i\mathbb{Z})\times\{0\}.

The condition that the \mathbb{R}-span and \mathbb{C}-span of Λ\Lambda are equal is essential, even without the dimension conditions, as the following example of [2] shows.

Example 1.3 ([2, Example in Section 4]).

Let XX be the hypersurface z=xyeiπ/4(y+1)z=xye^{-i\pi/4}(y+1) in 3\mathbb{C}^{3}, and let Λ=3\Lambda=\mathbb{Z}^{3}. Letting p:/p\colon\mathbb{C}\to\mathbb{C}/\mathbb{Z} be the projection and defining

L={s:arg(s2+s)=π/4 or 5π/4}{s:s2+s=0},L=\{s\in\mathbb{C}:\arg(s^{2}+s)=\pi/4\text{ or }5\pi/4\}\cup\{s\in\mathbb{C}:s^{2}+s=0\},

we have

Fl(X)=[(/)×p(L)×(/)][p(eiπ/4)×(/)×(/)].{\operatorname{Fl}}(X)=[(\mathbb{C}/\mathbb{Z})\times p(L)\times(\mathbb{C}/\mathbb{Z})]\cup[p(e^{i\pi/4}\mathbb{R})\times(\mathbb{C}/\mathbb{Z})\times(\mathbb{C}/\mathbb{Z})].

But one can show that any translate of a proper subset C+V3C+V\subset\mathbb{C}^{3} with CC an algebraic variety and VV a real vector subspace cannot intersect ×L×\mathbb{C}\times L\times\mathbb{C} in a full-dimensional set. So if π(X)¯=π(X)i=1m(π(Ci)+𝕋i)\overline{\pi(X)}=\pi(X)\cup\bigcup_{i=1}^{m}(\pi(C_{i})+\mathbb{T}_{i}), we obtain a contradiction by applying the Baire category theorem to the countable union π1(π(X)¯)=iCi+π1(𝕋i)\pi^{-1}(\overline{\pi(X)})=\bigcup_{i}C_{i}+\pi^{-1}(\mathbb{T}_{i}) of translates for i=1,,mi=1,\ldots,m of proper subsets of 3\mathbb{C}^{3} of the form Ci+ViC_{i}+V_{i}, where ViV_{i} is the connected component of π1(𝕋i)\pi^{-1}(\mathbb{T}_{i}) at 0, obtains a contradiction.

Gallinaro [3, Question 6.3.2], in work on Zilber’s exponential-algebraic closedness conjecture (an equivalent formulation of Zilber’s quasi-minimality conjecture [6] as established by Bays-Kirby [1, Theorem 1.5]), has conjectured that such a decomposition should be possible if the CiC_{i} are taken semi-algebraic. Our second main result, an o-minimal version of 1.1 with no restrictions on Λ\Lambda generalizing [4, Theorem 1.2], confirms this conjecture.

Theorem 1.4.

Fix an o-minimal structure om\mathbb{R}_{om} extending the field \mathbb{R}. Let XnX\subset\mathbb{R}^{n} be a definable closed set, LnL\subset\mathbb{R}^{n} an \mathbb{R}-subspace, and Λn\Lambda\subset\mathbb{R}^{n} a discrete subgroup with Λ=L\mathbb{R}\Lambda=L. Then for π:nn/Λ\pi:\mathbb{R}^{n}\to\mathbb{R}^{n}/\Lambda the quotient map, the following is true. There are definable closed sets C1,,CmnC_{1},\ldots,C_{m}\subset\mathbb{R}^{n} with dimCi<dimX\dim C_{i}<\dim X and compact positive-dimensional real tori 𝕋1,,𝕋mL/Λ\mathbb{T}_{1},\ldots,\mathbb{T}_{m}\subset L/\Lambda such that

Fl(X)=i=1m(π(Ci)+𝕋i).{\operatorname{Fl}}(X)=\bigcup_{i=1}^{m}(\pi(C_{i})+\mathbb{T}_{i}).

In particular, π(X)¯=π(X)i=1m(π(Ci)+𝕋i)\overline{\pi(X)}=\pi(X)\cup\bigcup_{i=1}^{m}(\pi(C_{i})+\mathbb{T}_{i}).

Remark 1.5.

As in [4], we will be able to take C1,,CmC_{1},\ldots,C_{m} to only depend on XX and LL, and take positive-dimensional \mathbb{R}-subspaces V1,,VmLV_{1},\ldots,V_{m}\subset L only depending on XX and LL such that 𝕋i=π(Vi)¯\mathbb{T}_{i}=\overline{\pi(V_{i})}.

In the cocompact case [4] also shows that we may take CiC_{i} to be compact if 𝕋i\mathbb{T}_{i} is inclusion-maximal amongst 𝕋1,,𝕋m\mathbb{T}_{1},\ldots,\mathbb{T}_{m}, but this is false in this more general setting even for X=2X=\mathbb{R}^{2} and Λ=×{0}\Lambda=\mathbb{Z}\times\{0\}.

Corollary 1.6 ([3, Question 6.3.2]).

Let XnX\subset\mathbb{C}^{n} be a closed semi-algebraic set (such as a complex algebraic variety), and exp:n()n\exp:\mathbb{C}^{n}\to(\mathbb{C}^{*})^{n} the map (z1,,zn)(exp(z1),,exp(zn))(z_{1},\ldots,z_{n})\mapsto(\exp(z_{1}),\ldots,\exp(z_{n})). Then there are semi-algebraic subsets C1,,CmnC_{1},\ldots,C_{m}\subset\mathbb{C}^{n} with dimCi<dimX\dim C_{i}<\dim X and compact positive-dimensional real tori 𝕋1,,𝕋m()n\mathbb{T}_{1},\ldots,\mathbb{T}_{m}\subset(\mathbb{C}^{*})^{n} such that

exp(X)¯=exp(X)i=1mexp(Ci)𝕋i.\overline{\exp(X)}=\exp(X)\cup\bigcup_{i=1}^{m}\exp(C_{i})\cdot\mathbb{T}_{i}.
Proof.

We apply the theorem with Λ=nn\Lambda=\mathbb{Z}^{n}\subset\mathbb{C}^{n} in the semi-algebraic o-minimal structure alg\mathbb{R}_{alg}, since exp\exp is surjective with kernel n\mathbb{Z}^{n}, so exp\exp is identified with the quotient map nn/n\mathbb{C}^{n}\to\mathbb{C}^{n}/\mathbb{Z}^{n}. ∎

To prove the cocompact case, the main technical tool introduced in [4] is a definable collection of “asymptotic flats” 𝒜(X)\mathcal{A}^{\mathbb{R}}(X) (or 𝒜(X)\mathcal{A}^{\mathbb{C}}(X)) associated to XX which encode the linear behaviour of sequences of points of XX. Roughly speaking, limiting sequences of points are encoded by realizations of XX in a saturated extension \mathcal{R} of \mathbb{R} (or in 𝒞:=i\mathcal{C}:=\mathcal{R}\oplus i\mathcal{R}), and we associate to such a point the smallest flat defined over \mathbb{R} (or \mathbb{C}) which is infinitesimally close to this realization.

The key idea in our proof is that if we decompose our ambient vector space as LLL\oplus L^{\perp}, then a sequence of points x1,x2,x_{1},x_{2},\ldots limiting to a coset of Λ\Lambda has convergent LL^{\perp}-component, and the corresponding asymptotic flat would then have linear part contained in L{0}L\oplus\{0\}. This motivates us to consider just the sub-family of flats of 𝒜(X)\mathcal{A}(X) with no linear part in the LL^{\perp} direction, and by doing so we can carry out a very similar argument in both the complex algebraic and o-minimal settings.

Acknowledgements. We thank Y. Peterzil, S. Starchenko, and F. Gallinaro for helpful answers to questions.

2. Model theory setup

We closely follow the setup of [4], which we recall briefly. Fix a cardinal κ>2ω\kappa>2^{\omega}. \mathbb{R} will be considered as an ordered field over the language 0,1,+,×,<\langle 0,1,+,\times,<\rangle. Let full\mathbb{R}_{full} be the expansion of the field \mathbb{R} by all subsets of n\mathbb{R}^{n} for all nn, full\mathcal{L}_{full} the corresponding language, and full\mathcal{R}_{full} a κ\kappa-saturated extension. Denote by \mathcal{R} the underlying field. Fix an o-minimal expansion om\mathbb{R}_{om} of \mathbb{R} with corresponding language om\mathcal{L}_{om}, and let om\mathcal{R}_{om} be the corresponding reduct fullom\mathcal{R}_{full}\upharpoonright\mathcal{L}_{om} of full\mathcal{R}_{full}, which is also κ\kappa-saturated.

We denote μ\mu_{\mathcal{R}}\subset\mathcal{R} for the set of infinitesimal elements and 𝒪=μ\mathcal{O}_{\mathcal{R}}=\mathcal{R}\oplus\mu_{\mathcal{R}} the set of bounded elements (neither of which are definable in om\mathcal{R}_{om}). We denote st:𝒪\operatorname{st}:\mathcal{O}_{\mathcal{R}}\to\mathbb{R} the standard part map, and for a subset SnS\subset\mathcal{R}^{n} we denote

st(S):=st(S𝒪n).\operatorname{st}(S):=\operatorname{st}(S\cap\mathcal{O}_{\mathcal{R}}^{n}).

The complex numbers =i\mathbb{C}=\mathbb{R}\oplus i\mathbb{R} will be considered as a structure with signature 0,1,+,×\langle 0,1,+,\times\rangle. We denote by 𝒞=i\mathcal{C}=\mathcal{R}\oplus i\mathcal{R}, the set of infinitesimals μ𝒞=μiμ\mu_{\mathcal{C}}=\mu_{\mathcal{R}}\oplus i\mu_{\mathcal{R}}, and the set of bounded elements 𝒪𝒞=𝒪i𝒪\mathcal{O}_{\mathcal{C}}=\mathcal{O}_{\mathcal{R}}\oplus i\mathcal{O}_{\mathcal{R}}. With these identifications, we may similarly define st:𝒪𝒞𝒞\operatorname{st}:\mathcal{O}_{\mathcal{C}}\to\mathcal{C} and st(S)=st(S𝒪𝒞n)\operatorname{st}(S)=\operatorname{st}(S\cap\mathcal{O}_{\mathcal{C}}^{n}) for a subset S𝒞nS\subset\mathcal{C}^{n}. Through the identification 𝒞=i\mathcal{C}=\mathcal{R}\oplus i\mathcal{R}, we may consider 𝒞full\mathcal{C}_{full} the expansion of \mathbb{C} by all subsets of n\mathbb{C}^{n} for all nn.

Denote val=0,1,+,,𝒪\mathcal{L}_{val}=\langle 0,1,+,\cdot,\mathcal{O}\rangle the language of a valued field, with 𝒪\mathcal{O} a unary predicate for the valuation ring, and 𝒞val\mathcal{C}_{val} the val\mathcal{L}_{val}-structure of the valued field 𝒞\mathcal{C} augmented with the set 𝒪𝒞\mathcal{O}_{\mathcal{C}} of bounded elements. Note also that μ𝒞\mu_{\mathcal{C}} is definable in 𝒞val\mathcal{C}_{val} by the formula x=0x1𝒪𝒞x=0\wedge x^{-1}\not\in\mathcal{O}_{\mathcal{C}}. Note that 𝒞val\mathcal{C}_{val} is not a reduct of 𝒞full\mathcal{C}_{full} and in particular not κ\kappa-saturated, since for example the finitely-satisfiable collection of 2ω2^{\omega} formulas x𝒪𝒞c{xc+μ𝒞}x\in\mathcal{O}_{\mathcal{C}}\cup\bigcup_{c\in\mathbb{C}}\{x\not\in c+\mu_{\mathcal{C}}\} is not satisfiable.

We will denote (𝔽,)(\mathbb{F},\mathcal{F}) to be either (,)(\mathbb{R},\mathcal{R}) or (,𝒞)(\mathbb{C},\mathcal{C}).

If \mathcal{L} is one of the languages om,full\mathcal{L}_{om},\mathcal{L}_{full} when working in the real setting or val,full\mathcal{L}_{val},\mathcal{L}_{full} when working in the complex setting, then an \mathcal{L}-definable partial type p(x1,,xn)p(x_{1},\ldots,x_{n}) is a finitely satisfiable collection of formulas over the language \mathcal{L} with free variables x1,,xnx_{1},\ldots,x_{n} and constants in AA. We denote by p()np(\mathcal{F})\subset\mathcal{F}^{n} the set of realizations of pp over \mathcal{F}.

Suppose we are in one of the κ\kappa-saturated structures om,full\mathcal{R}_{om},\mathcal{R}_{full}, or 𝒞full\mathcal{C}_{full}, and consider types pp defined over a set of constants AA with |A|<κ|A|<\kappa. Two such types p,qp,q are considered equivalent if for every finite p0pp_{0}\subset p there exists a finite q0qq_{0}\subset q such that q0()p0()q_{0}(\mathcal{F})\subset p_{0}(\mathcal{F}) and vice versa. By κ\kappa-saturation, two such types p,qp,q are equivalent if and only if p()=q()p(\mathcal{F})=q(\mathcal{F}). If p,qp,q are any two types, then the sum p+qp+q is defined by

p+q={y,z with x=y+z and f(y) and g(z) both true}fp,gq.p+q=\{\exists y,z\in\mathcal{F}\text{ with }x=y+z\text{ and }f(y)\text{ and }g(z)\text{ both true}\}_{f\in p,g\in q}.

By κ\kappa-saturation, for two such types we have (p+q)()=p()+q()(p+q)(\mathcal{F})=p(\mathcal{F})+q(\mathcal{F}), and consequently for three such types p,q,rp,q,r, the types (p+q)+r(p+q)+r and p+(q+r)p+(q+r) are equivalent.

For SnS\subset\mathcal{F}^{n}, we can consider SS as an full\mathcal{F}_{full}-definable subset of n\mathcal{F}^{n}. By abuse of notation we also use SS to denote the partial type of all full\mathcal{F}_{full}-formulas with constants in 𝔽\mathbb{F} satisfied by SS, and we write S:=S()S^{\sharp}:=S(\mathcal{F}) for the set of realizations of SS over \mathcal{F}. Also by abuse of notation, we identify μ\mu_{\mathcal{R}} with the partial type {r<x<r}r>0\{-r<x<r\}_{r\in\mathbb{R}_{>0}} in om\mathcal{R}_{om} or full\mathcal{R}_{full} and μ𝒞\mu_{\mathcal{C}} with the partial type {r<Re(x),Im(x)<r}r>0\{-r<Re(x),Im(x)<r\}_{r\in\mathbb{R}_{>0}} in 𝒞full\mathcal{C}_{full}, so that μ()=μ\mu_{\mathcal{R}}(\mathcal{R})=\mu_{\mathcal{R}} and μ𝒞(𝒞)=μ𝒞\mu_{\mathcal{C}}(\mathcal{C})=\mu_{\mathcal{C}}. These abuses of notation are consistent with the sum notation for types, as by κ\kappa-saturatedness for S,TnS,T\subset\mathcal{F}^{n} we have

(S+T)=S+T, and (S+μ𝒞n)()=S+μ𝒞n.\displaystyle(S+T)^{\sharp}=S^{\sharp}+T^{\sharp},\text{ and }(S+\mu_{\mathcal{C}}^{n})(\mathcal{F})=S^{\sharp}+\mu_{\mathcal{C}}^{n}.

Finally, by κ\kappa-saturatedness, for any S𝔽nS\subset\mathbb{F}^{n} we have

st(S)=S¯,\operatorname{st}(S^{\sharp})=\overline{S},

where S¯𝔽n\overline{S}\subset\mathbb{F}^{n} is the topological closure.

3. Flows via non-standard realizations

Note that

π1Fl(X)=R>0(XB(0,R))+Λ¯.\pi^{-1}{\operatorname{Fl}}(X)=\bigcap_{R\in\mathbb{R}_{>0}}\overline{(X-B(0,R))+\Lambda}.

We will give an alternate description of π1Fl(X)\pi^{-1}{\operatorname{Fl}}(X) via non-standard realizations.

Definition 3.1.

Let X𝔽nX\subset\mathbb{F}^{n} be an om\mathbb{R}_{om}-definable set for 𝔽=\mathbb{F}=\mathbb{R}, or a constructible set for 𝔽=\mathbb{F}=\mathbb{C}. We define the partial type

X𝔽={{xX and xR}R>0𝔽={xX and x𝒪𝒞n}𝔽=.X_{\infty}^{\mathbb{F}}=\begin{cases}\{x\in X\text{ and }\|x\|\geq R\}_{R\in\mathbb{R}_{>0}}&\mathbb{F}=\mathbb{R}\\ \{x\in X\text{ and }x\not\in\mathcal{O}_{\mathcal{C}}^{n}\}&\mathbb{F}=\mathbb{C}.\end{cases}

This is either an om\mathcal{R}_{om}-partial type with constants in \mathbb{R} or a 𝒞val\mathcal{C}_{val}-partial type with constants in \mathbb{C}.

Theorem 3.2.

For X𝔽nX\subset\mathbb{F}^{n} as above, we have X𝔽()=X𝒪nX_{\infty}^{\mathbb{F}}(\mathcal{F})=X^{\sharp}\setminus\mathcal{O}_{\mathcal{F}}^{n}, and

π1Fl(X)=st(X𝔽()+Λ).\pi^{-1}{\operatorname{Fl}}(X)=\operatorname{st}(X_{\infty}^{\mathbb{F}}(\mathcal{F})+\Lambda^{\sharp}).
Proof.

That X𝔽()=X𝒪nX_{\infty}^{\mathbb{F}}(\mathcal{F})=X^{\sharp}\setminus\mathcal{O}_{\mathcal{F}}^{n} is trivial when 𝔽=\mathbb{F}=\mathbb{C}, and when 𝔽=\mathbb{F}=\mathbb{R} this follows from the fact that 𝒪n\mathcal{O}_{\mathcal{R}}^{n} is by definition the set of bounded elements of n\mathcal{R}^{n}.

Now, for 𝔽=\mathbb{F}=\mathbb{C}, since by treating n\mathbb{C}^{n} as a real vector space and XX as a semi-algebraic (and in particular om\mathbb{R}_{om}-definable) set in 2n\mathbb{R}^{2n} we have X(𝒞)=X()X_{\infty}^{\mathbb{C}}(\mathcal{C})=X_{\infty}^{\mathbb{R}}(\mathcal{R}) by the previous paragraph, and st\operatorname{st} is unchanged under restriction of scalars, it suffices to prove the statement for =\mathcal{F}=\mathcal{R}.

If π(z)Fl(X)\pi(z)\in Fl(X), the type (X+Λ)(z+μn)(X_{\infty}^{\mathbb{R}}+\Lambda)\cup(z+\mu_{\mathcal{R}}^{n}) is finitely satisfied. Thus, by saturation of \mathcal{R}, there exists z(X+Λ)()=X()+Λz^{\prime}\in(X_{\infty}^{\mathbb{R}}+\Lambda)(\mathcal{R})=X_{\infty}^{\mathbb{R}}(\mathcal{R})+\Lambda^{\sharp} infinitesimally close to zz. Since st\operatorname{st} is insensitive to addition by an element of infinitesimals, we have zst(X()+Λ)z\in\operatorname{st}(X_{\infty}^{\mathbb{R}}(\mathcal{R})+\Lambda^{\sharp}). Conversely, if α(X()+Λ)𝒪n\alpha\in(X_{\infty}^{\mathbb{R}}(\mathcal{R})+\Lambda^{\sharp})\cap\mathcal{O}_{\mathcal{F}}^{n}, then since X()(XB(0,R))X_{\infty}^{\mathbb{R}}(\mathcal{R})\subset(X-B(0,R))^{\sharp}, we have st(α)st((XB(0,R))+Λ)=(XB(0,R))+Λ¯\operatorname{st}(\alpha)\in\operatorname{st}((X-B(0,R))^{\sharp}+\Lambda^{\sharp})=\overline{(X-B(0,R))+\Lambda} for all R>0R\in\mathbb{R}_{>0}. ∎

4. Asymptotic flats

Let GrAffi𝔽(𝔽n)\operatorname{GrAff}^{\mathbb{F}}_{i}(\mathbb{F}^{n}) be the Grassmannian of affine ii-dimensional flats of 𝔽n\mathbb{F}^{n}, and let

GrAff𝔽(𝔽n)=GrAff0𝔽(𝔽n)GrAffn𝔽(𝔽n).\operatorname{GrAff}^{\mathbb{F}}(\mathbb{F}^{n})=\operatorname{GrAff}^{\mathbb{F}}_{0}(\mathbb{F}^{n})\sqcup\ldots\sqcup\operatorname{GrAff}^{\mathbb{F}}_{n}(\mathbb{F}^{n}).

Note that GrAff0𝔽(𝔽n)=𝔽n\operatorname{GrAff}^{\mathbb{F}}_{0}(\mathbb{F}^{n})=\mathbb{F}^{n} and GrAffn𝔽(𝔽n)={{𝔽n}}\operatorname{GrAff}^{\mathbb{F}}_{n}(\mathbb{F}^{n})=\{\{\mathbb{F}^{n}\}\}.

For αn\alpha\in\mathcal{F}^{n}, we denote by Aα𝔽A_{\alpha}^{\mathbb{F}} the minimal 𝔽\mathbb{F}-flat under inclusion such that α(Aα𝔽)#+μn\alpha\in(A_{\alpha}^{\mathbb{F}})^{\#}+\mu_{\mathcal{F}}^{n} (that this exists is [4, Proposition 4.2] when 𝔽=\mathbb{F}=\mathbb{C} and as noted in [4, Section 7.1], the proof also works for 𝔽=\mathbb{F}=\mathbb{R}). For 0in0\leq i\leq n define 𝒜i𝔽(X):={Aα𝔽:αX and dim𝔽(Aα𝔽)=i}GrAffi𝔽(𝔽n)\mathcal{A}_{i}^{\mathbb{F}}(X):=\{A_{\alpha}^{\mathbb{F}}:\alpha\in X^{\sharp}\text{ and }\dim_{\mathbb{F}}(A_{\alpha}^{\mathbb{F}})=i\}\subset{\operatorname{GrAff}}_{i}^{\mathbb{F}}(\mathbb{F}^{n}) and

𝒜𝔽(X):={Aα𝔽:αX}=i=0nAα,i𝔽(X)GrAffn𝔽.\mathcal{A}^{\mathbb{F}}(X):=\{A_{\alpha}^{\mathbb{F}}:\alpha\in X^{\sharp}\}=\bigsqcup_{i=0}^{n}A_{\alpha,i}^{\mathbb{F}}(X)\subset\operatorname{GrAff}^{\mathbb{F}}_{n}.

By [4, Theorems 4.5 and 7.1] this is a definable subset of GrAff𝔽(𝔽n)\operatorname{GrAff}^{\mathbb{F}}(\mathbb{F}^{n}), and by definition Aα𝔽𝒜0𝔽(X)A_{\alpha}^{\mathbb{F}}\in\mathcal{A}_{0}^{\mathbb{F}}(X) if and only if α𝒪n\alpha\in\mathcal{O}^{n}, in which case Aα𝔽=st(α)A_{\alpha}^{\mathbb{F}}=\operatorname{st}(\alpha) (see [4, Lemma 4.8]). In particular 𝒜0𝔽(X)=X¯\mathcal{A}_{0}^{\mathbb{F}}(X)=\overline{X}. Write 𝒜pos𝔽(X):=i=1n𝒜i𝔽(X)\mathcal{A}_{pos}^{\mathbb{F}}(X):=\bigsqcup_{i=1}^{n}\mathcal{A}^{\mathbb{F}}_{i}(X).

A slight refinement of the arguments in [4] shows that if Λ\Lambda is cocompact in 𝔽n\mathbb{F}^{n}, then

Fl(X)=A𝒜pos𝔽(X)π(A)¯.{\operatorname{Fl}}(X)=\bigcup_{A\in\mathcal{A}_{pos}^{\mathbb{F}}(X)}\overline{\pi(A)}.

We will adapt this now to lattices Λ\Lambda with Λ=L\mathbb{R}\Lambda=L an 𝔽\mathbb{F}-linear subspace of 𝔽n\mathbb{F}^{n}.

Let GrAffi𝔽(𝔽n;L)GrAffi𝔽(𝔽n)\operatorname{GrAff}_{i}^{\mathbb{F}}(\mathbb{F}^{n};L)\subset\operatorname{GrAff}_{i}^{\mathbb{F}}(\mathbb{F}^{n}) be the subset of affine ii-dimensional flats AA of 𝔽n\mathbb{F}^{n} such that the linear part of AA is contained in LL, and

GrAff𝔽(𝔽n;L)=GrAff0𝔽(𝔽n;L)GrAffdim𝔽(L)𝔽(𝔽n;L)\operatorname{GrAff}^{\mathbb{F}}(\mathbb{F}^{n};L)=\operatorname{GrAff}_{0}^{\mathbb{F}}(\mathbb{F}^{n};L)\sqcup\ldots\sqcup\operatorname{GrAff}^{\mathbb{F}}_{\dim_{\mathbb{F}}(L)}(\mathbb{F}^{n};L)

For 0idim𝔽(L)0\leq i\leq\dim_{\mathbb{F}}(L), we define 𝒜i𝔽(X;L):=𝒜i𝔽(X)GrAffi𝔽(𝔽n;L)\mathcal{A}^{\mathbb{F}}_{i}(X;L):=\mathcal{A}^{\mathbb{F}}_{i}(X)\cap\operatorname{GrAff}_{i}^{\mathbb{F}}(\mathbb{F}^{n};L) and

𝒜𝔽(X;L):=𝒜𝔽(X)GrAff𝔽(𝔽n;L).\mathcal{A}^{\mathbb{F}}(X;L):=\mathcal{A}^{\mathbb{F}}(X)\cap\operatorname{GrAff}^{\mathbb{F}}(\mathbb{F}^{n};L).

Clearly this is a definable subset, and 𝒜𝔽(X;L)=X\mathcal{A}^{\mathbb{F}}(X;L)=X. Write 𝒜pos𝔽(X;L):=𝒜pos𝔽(X)GrAff𝔽(𝔽n;L)\mathcal{A}_{pos}^{\mathbb{F}}(X;L):=\mathcal{A}^{\mathbb{F}}_{pos}(X)\cap\operatorname{GrAff}^{\mathbb{F}}(\mathbb{F}^{n};L).

Theorem 4.1.

If Λ𝔽n\Lambda\subset\mathbb{F}^{n} is a discrete subgroup and L𝔽nL\subset\mathbb{F}^{n} is an 𝔽\mathbb{F}-linear subspace such that Λ=L\mathbb{R}\Lambda=L, then

Fl(X)=A𝒜pos𝔽(X;L)π(A)¯.{\operatorname{Fl}}(X)=\bigcup_{A\in\mathcal{A}^{\mathbb{F}}_{pos}(X;L)}\overline{\pi(A)}.

Before proving the theorem, we start by proving some lemmas.

Lemma 4.2.

Suppose that L=𝔽k×{0}nkL=\mathbb{F}^{k}\times\{0\}^{n-k}. Then for αn\alpha\in\mathcal{F}^{n}, we have Aα𝔽GrAffi(𝔽n;L)A_{\alpha}^{\mathbb{F}}\in\operatorname{GrAff}_{i}(\mathbb{F}^{n};L) if and only if α=(α1,α2)k×𝒪nk\alpha=(\alpha_{1},\alpha_{2})\in\mathcal{F}^{k}\times\mathcal{O}_{\mathcal{F}}^{n-k}, in which case Aα𝔽𝔽k×{st(α2)}A_{\alpha}^{\mathbb{F}}\subset\mathbb{F}^{k}\times\{\operatorname{st}(\alpha_{2})\}. In particular,

𝒜pos𝔽(X;L)={Aα𝔽:αX𝔽()(k×𝒪nk)}.\mathcal{A}_{pos}^{\mathbb{F}}(X;L)=\{A_{\alpha}^{\mathbb{F}}:\alpha\in X_{\infty}^{\mathbb{F}}(\mathcal{F})\cap(\mathcal{F}^{k}\times\mathcal{O}_{\mathcal{F}}^{n-k})\}.
Proof.

If Aα𝔽GrAffi(𝔽n;L)A_{\alpha}^{\mathbb{F}}\in\operatorname{GrAff}_{i}(\mathbb{F}^{n};L), then Aα𝔽𝔽k×{a}A_{\alpha}^{\mathbb{F}}\subset\mathbb{F}^{k}\times\{a\} for some a𝔽nka\in\mathbb{F}^{n-k}, so

α(Aα𝔽)+μn(𝔽k×{a})+μn=k×{a}+μnk×𝒪nk.\alpha\in(A_{\alpha}^{\mathbb{F}})^{\sharp}+\mu_{\mathcal{F}}^{n}\subset(\mathbb{F}^{k}\times\{a\})^{\sharp}+\mu_{\mathcal{F}}^{n}=\mathcal{F}^{k}\times\{a\}+\mu_{\mathcal{F}}^{n}\subset\mathcal{F}^{k}\times\mathcal{O}_{\mathcal{F}}^{n-k}.

Conversely, if αk×𝒪nk\alpha\in\mathcal{F}^{k}\times\mathcal{O}_{\mathcal{F}}^{n-k}, then

αk×{st(α2)}+μn=(𝔽k×{st(α2)})+μn,\alpha\in\mathcal{F}^{k}\times\{\operatorname{st}(\alpha_{2})\}+\mu_{\mathcal{F}}^{n}=(\mathbb{F}^{k}\times\{\operatorname{st}(\alpha_{2})\})^{\sharp}+\mu_{\mathcal{F}}^{n},

so Aα𝔽𝔽k×{st(α2)}=L+(0,st(α2))A_{\alpha}^{\mathbb{F}}\subset\mathbb{F}^{k}\times\{\operatorname{st}(\alpha_{2})\}=L+(0,\operatorname{st}(\alpha_{2})) and this shows Aα𝔽GrAffi(𝔽n;L)A_{\alpha}^{\mathbb{F}}\in\operatorname{GrAff}_{i}(\mathbb{F}^{n};L). ∎

Definition 4.3.

For αn\alpha\in\mathcal{F}^{n}, we define pαp_{\alpha} to be the complete 𝒞val\mathcal{C}_{val}-type of α\alpha with \mathbb{C}-constants if 𝔽=\mathbb{F}=\mathbb{C} and to be the complete om\mathcal{R}_{om}-type of α\alpha with \mathbb{R}-constants if 𝔽=\mathbb{F}=\mathbb{R}.

Lemma 4.4.

For α=(α1,α2)k×𝒪nk\alpha=(\alpha_{1},\alpha_{2})\in\mathcal{F}^{k}\times\mathcal{O}_{\mathcal{F}}^{n-k} and a=st(α2)a=\operatorname{st}(\alpha_{2}), then pα()k×{a}+μnp_{\alpha}(\mathcal{F})\subset\mathcal{F}^{k}\times\{a\}+\mu_{\mathcal{F}}^{n}. Also, if αX()\alpha\in X_{\infty}(\mathcal{F}) then pα()X()p_{\alpha}(\mathcal{F})\subset X_{\infty}(\mathcal{F}), so in particular

αX𝔽()(k×𝒪nk)pα()X𝔽()(k×𝒪nk).\alpha\in X_{\infty}^{\mathbb{F}}(\mathcal{F})\cap(\mathcal{F}^{k}\times\mathcal{O}_{\mathcal{F}}^{n-k})\implies p_{\alpha}(\mathcal{F})\subset X_{\infty}^{\mathbb{F}}(\mathcal{F})\cap(\mathcal{F}^{k}\times\mathcal{O}_{\mathcal{F}}^{n-k}).
Proof.

k×{a}+μn=(𝔽k×{a})+μn\mathcal{F}^{k}\times\{a\}+\mu_{\mathcal{F}}^{n}=(\mathbb{F}^{k}\times\{a\})^{\sharp}+\mu_{\mathcal{F}}^{n} corresponds to a partial type that α\alpha satisfies by construction. Similarly for X𝔽X_{\infty}^{\mathbb{F}} (noting that this partial type is definable over 𝒞val\mathcal{C}_{val} with \mathbb{C}-coefficients when 𝔽=\mathbb{F}=\mathbb{C} and over om\mathbb{R}_{om} with \mathbb{R}-coefficients when 𝔽=\mathbb{F}=\mathbb{R}). ∎

Lemma 4.5.

For Aα𝔽GrAff(𝔽n;L)A_{\alpha}^{\mathbb{F}}\in\operatorname{GrAff}(\mathbb{F}^{n};L) we have Aα𝔽+Λ¯=st(pα()+Λ)\overline{A_{\alpha}^{\mathbb{F}}+\Lambda}=\operatorname{st}(p_{\alpha}(\mathcal{F})+\Lambda^{\sharp}).

Proof.

After a linear change of coordinates, we may assume that ΛL=𝔽k×{0}\Lambda\subset L=\mathbb{F}^{k}\times\{0\}. Then by 4.2 we have α=(α1,α2)k×𝒪nk\alpha=(\alpha_{1},\alpha_{2})\in\mathcal{F}^{k}\times\mathcal{O}_{\mathcal{F}}^{n-k}, and writing a=st(α2)a=\operatorname{st}(\alpha_{2}) we have Aα𝔽𝔽k×{a}A_{\alpha}^{\mathbb{F}}\subset\mathbb{F}^{k}\times\{a\}, so the left hand side is contained in 𝔽k×{a}\mathbb{F}^{k}\times\{a\}. Additionally, ΛL=k×{0}nk\Lambda^{\sharp}\subset L^{\sharp}=\mathcal{F}^{k}\times\{0\}^{n-k} and by 4.4 we also have pα()k×{a}+μnp_{\alpha}(\mathcal{F})\subset\mathcal{F}^{k}\times\{a\}+\mu_{\mathcal{F}}^{n}, so the right hand side is also contained in 𝔽k×{a}\mathbb{F}^{k}\times\{a\}.

Let Λ{0}k×𝔽nk\Lambda^{\prime}\subset\{0\}^{k}\times\mathbb{F}^{n-k} be a discrete subgroup with Λ={0}k×𝔽nk\mathbb{R}\Lambda^{\prime}=\{0\}^{k}\times\mathbb{F}^{n-k}. Then (Λ+Λ)=𝔽n\mathbb{R}(\Lambda+\Lambda^{\prime})=\mathbb{F}^{n} so by [4, Proposition 5.1] in the complex case and [4, Proposition 7.3] in the real case, we have

Aα𝔽+Λ¯+Λ=Aα𝔽+Λ+Λ¯=st(pα()+(Λ+Λ))=st(pα()+Λ+(Λ)).\overline{A_{\alpha}^{\mathbb{F}}+\Lambda}+\Lambda^{\prime}=\overline{A_{\alpha}^{\mathbb{F}}+\Lambda+\Lambda^{\prime}}=\operatorname{st}(p_{\alpha}(\mathcal{F})+(\Lambda+\Lambda^{\prime})^{\sharp})=\operatorname{st}(p_{\alpha}(\mathcal{F})+\Lambda^{\sharp}+(\Lambda^{\prime})^{\sharp}).

Because pα(),Λk×𝒪nkp_{\alpha}(\mathcal{F}),\Lambda^{\sharp}\subset\mathcal{F}^{k}\times\mathcal{O}_{\mathcal{F}}^{n-k}, for the last nkn-k coordinates of the sum pα()+Λ+(Λ)p_{\alpha}(\mathcal{F})+\Lambda^{\sharp}+(\Lambda^{\prime})^{\sharp} to be bounded we need the element from (Λ)(\Lambda^{\prime})^{\sharp} to lie in {0}×𝒪nk\{0\}\times\mathcal{O}_{\mathcal{F}}^{n-k}, which in particular means it is infinitesimally close to an element of Λ\Lambda^{\prime}. Therefore,

Aα𝔽+Λ¯+Λ=st(pα()+Λ)+Λ.\overline{A_{\alpha}^{\mathbb{F}}+\Lambda}+\Lambda^{\prime}=\operatorname{st}(p_{\alpha}(\mathcal{F})+\Lambda^{\sharp})+\Lambda^{\prime}.

Intersecting both sides of the equality with 𝔽k×{a}\mathbb{F}^{k}\times\{a\}, we obtain the desired equality. ∎

Proof of 4.1.

We may suppose that ΛL=𝔽k×{0}nk\Lambda\subset L=\mathbb{F}^{k}\times\{0\}^{n-k}. Then since Λk×{0}nk\Lambda^{\sharp}\subset\mathcal{F}^{k}\times\{0\}^{n-k}, if a sum α+λ𝒪n\alpha+\lambda\in\mathcal{O}_{\mathcal{F}}^{n} for some λΛ\lambda\in\Lambda^{\sharp} then we must have αk×𝒪nk\alpha\in\mathcal{F}^{k}\times\mathcal{O}_{\mathcal{F}}^{n-k}. Therefore,

π1(Fl(X))=αX𝔽()st(α+Λ)=αX𝔽()(k×𝒪nk)st(α+Λ)\pi^{-1}({\operatorname{Fl}}(X))=\bigcup_{\alpha\in X_{\infty}^{\mathbb{F}}(\mathcal{F})}\operatorname{st}(\alpha+\Lambda^{\sharp})=\bigcup_{\alpha\in X_{\infty}^{\mathbb{F}}(\mathcal{F})\cap(\mathcal{F}^{k}\times\mathcal{O}_{\mathcal{F}}^{n-k})}\operatorname{st}(\alpha+\Lambda^{\sharp})
=αX𝔽()(k×𝒪nk)st(pα()+Λ)=Aα𝔽𝒜pos𝔽(X;L)Aα𝔽+Λ¯=\bigcup_{\alpha\in X_{\infty}^{\mathbb{F}}(\mathcal{F})\cap(\mathcal{F}^{k}\times\mathcal{O}_{\mathcal{F}}^{n-k})}\operatorname{st}(p_{\alpha}(\mathcal{F})+\Lambda^{\sharp})=\bigcup_{A_{\alpha}^{\mathbb{F}}\in\mathcal{A}^{\mathbb{F}}_{pos}(X;L)}\overline{A_{\alpha}^{\mathbb{F}}+\Lambda}

where the third equality is by 4.4. ∎

Remark 4.6.

In the cocompact case, every positive-dimensional algebraic set has a nontrivial asymptotic flat by [4, Lemma 4.8]. In contrast, 𝒜pos𝔽(X;L)\mathcal{A}^{\mathbb{F}}_{pos}(X;L) may be empty. For example, the parabola y=x2y=x^{2} in 2\mathbb{R}^{2} has no nontrivial asymptotic flats relative to the lattice Λ=×0\Lambda=\mathbb{Z}\times 0, because any unbounded solution is unbounded in both coordinates.

5. Neat families

In this section we take “definable” to mean defianble either over om\mathbb{R}_{om} in the language of ordered fields, or over \mathbb{C} in the language of fields, according to whether the setting is real or complex. By Chevalley’s theorem, in the latter case the definable sets are exactly constructible sets, i.e., Boolean combinations of algebraic sets.

For a flat AGrAff𝔽(𝔽n)A\in\operatorname{GrAff}^{\mathbb{F}}(\mathbb{F}^{n}), we write L(A)𝔽nL(A)\subset\mathbb{F}^{n} for the linear part of AA (the 𝔽\mathbb{F}-subspace such that we can write A=L(A)+cA=L(A)+c for some cc), and for a subset TGrAff𝔽(𝔽n)T\subset\operatorname{GrAff}^{\mathbb{F}}(\mathbb{F}^{n}), we define L(T)𝔽nL(T)\subset\mathbb{F}^{n} to be the smallest 𝔽\mathbb{F}-linear space containing L(A)L(A) for all ATA\in T.

Definition 5.1.

A “neat” family of flats is a definable family TGrAffi𝔽(𝔽n)T\subset\operatorname{GrAff}_{i}^{\mathbb{F}}(\mathbb{F}^{n}) such that

  1. (1)

    TT is a connected \mathbb{R}-submanifold.

  2. (2)

    For any nonempty open subset UTU\subset T we have L(U)=L(T)L(U)=L(T).

Theorem 5.2 ([4, Theorem 7.7 and Section 6.1]).

Every definable family TGrAff𝔽(𝔽n)T\subset\operatorname{GrAff}^{\mathbb{F}}(\mathbb{F}^{n}) can be written as a finite union of neat definable families.

Proof.

In the real setting, this is exactly what is proved in [4, Theorem 7.7]. In the complex setting, we may decompose TT into finitely many smooth pieces, and then further decompose each of those pieces into its finitely many irreducible components, so it suffices to show that if TT is smooth and irreducible then it is neat. A smooth irreducible set is a connected \mathbb{R}-submanifold, so it suffices to show that for any nonempty open subset UTU\subset T that L(U)=L(T)L(U)=L(T). Any nonempty open subset UTU\subset T is Zariski dense in TT by irreducibility of TT, so because the Zariski-closed set of those ATA\in T such that L(A)L(U)L(A)\subset L(U) contains UU, it must therefore be all of TT. Hence L(A)L(U)L(A)\subset L(U) for all ATA\in T, so L(T)L(U)L(T)\subset L(U), and L(U)=L(T)L(U)=L(T). ∎

By this theorem, we may break up the definable family of flats as 𝒜pos𝔽(X;L)=T1Tm\mathcal{A}_{pos}^{\mathbb{F}}(X;L)=T_{1}\cup\ldots\cup T_{m} where each TiT_{i} is a neat family in GrAffj(𝔽n;L)\operatorname{GrAff}_{j}(\mathbb{F}^{n};L) with 1jdim(L)1\leq j\leq\dim(L). We need the following proposition, the analogue of [4, Proposition 6.2 and Proposition 7.6] for GrAffj(𝔽n;L)\operatorname{GrAff}_{j}(\mathbb{F}^{n};L).

Proposition 5.3.

If TT is a neat family of jj-dimensional flats of 𝔽n\mathbb{F}^{n} in GrAffj(𝔽n;L)\operatorname{GrAff}_{j}(\mathbb{F}^{n};L), then ATπ(A)\bigcup_{A\in T}\pi(A) is topologically dense in ATπ(A+L(T))\bigcup_{A\in T}\pi(A+L(T)).

Proof.

The proof of [4, Proposition 7.6] carries over with essentially no modifications.

First we show the analogue of [4, Proposition 7.6(1)], that the set {AT:π(L(A))¯=π(L(T))¯}\{A\in T:\overline{\pi(L(A))}=\overline{\pi(L(T))}\} is topologically dense in TT. Indeed, there are only countably many closed real Lie subgroups of L/ΓL/\Gamma, so by the Baire Category theorem it suffices to show for any proper real Lie subgroup 𝕋π(L(T))¯\mathbb{T}\subsetneq\overline{\pi(L(T))} that the set {AT:π(L(A))𝕋}\{A\in T:\pi(L(A))\subset\mathbb{T}\} is nowhere dense in TT. This set is equal to {AT:L(A)W}\{A\in T:L(A)\subset W\}, where WW is the maximal 𝔽\mathbb{F}-linear subspace contained in π1(𝕋)\pi^{-1}(\mathbb{T}) (either the connected component WW^{\prime} of 0 in the real setting or WiWW^{\prime}\cap iW^{\prime} in the complex setting). Since this is definable, it suffices to show that it does not contain an open subset UU of TT. But since TT is neat, L(U)=L(T)L(U)=L(T), so not every L(A)L(A) for AUA\in U can be contained in WW since WW is a proper subspace of L(T)L(T).

Now we show ATπ(A)\bigcup_{A\in T}\pi(A) is topologically dense in ATπ(A+L(T))\bigcup_{A\in T}\pi(A+L(T)), the analogue of [4, Proposition 7.6(2)]. Indeed, by what we have just shown, there is a sequence A1,A2,AA_{1},A_{2},\ldots\to A such that π(L(Ai))¯=π(L(T))¯\overline{\pi(L(A_{i}))}=\overline{\pi(L(T))}, or equivalently π(Ai)¯=π(Ai+L(T))¯\overline{\pi(A_{i})}=\overline{\pi(A_{i}+L(T))}. Hence iπ(Ai)¯=iπ(Ai+L(T))¯\bigcup_{i}\overline{\pi(A_{i})}=\bigcup_{i}\overline{\pi(A_{i}+L(T))} contains π(A+L(T))¯\overline{\pi(A+L(T))} in its closure, and therefore iπ(A)\bigcup_{i}\pi(A) contains π(A+L(T))¯\overline{\pi(A+L(T))} in its closure. ∎

Now the proofs of 1.1 and 1.4 proceed identical to [4, Theorem 7.8]. We sketch how this works. For each ii, let L(Ti)L(T_{i})^{\perp} be a complementary 𝔽\mathbb{F}-subspace to L(Ti)L(T_{i}). For ATiA\in T_{i} the intersection (A+L(Ti))L(Ti)(A+L(T_{i}))\cap L(T_{i})^{\perp} is a single point cA,ic_{A,i}, and we let Ci=ATici,AL(Ti)C_{i}^{\prime}=\bigcup_{A\in T_{i}}c_{i,A}\subset L(T_{i})^{\perp} be the closure of the definable set of such points as we range over ATA\in T. Letting Ci=Ci¯C_{i}=\overline{C_{i}^{\prime}}, we have dim𝔽Ci=dim𝔽Ci\dim_{\mathbb{F}}C_{i}=\dim_{\mathbb{F}}C_{i}^{\prime}, and [4, Section 6.2 and 7.3.1, Proofs of Clause (i)] applied to CiC_{i}^{\prime} shows that with this construction, dim𝔽Ci=dim𝔽Ci<dim𝔽X\dim_{\mathbb{F}}C_{i}=\dim_{\mathbb{F}}C_{i}^{\prime}<\dim_{\mathbb{F}}X. Let Vi=L(Ti)V_{i}=L(T_{i}) and 𝕋i=π(Vi)¯\mathbb{T}_{i}=\overline{\pi(V_{i})}. Then we have (by 4.1 for the first equality, by 5.3 and the fact that Fl(X){\operatorname{Fl}}(X) is closed for the second equality, that A+L(Ti)=cA,i+ViA+L(T_{i})=c_{A,i}+V_{i} for the third equality, and that Fl(X){\operatorname{Fl}}(X) is closed for the fourth equality) that

Fl(X)=iATiπ(A)¯=iATiπ(A+L(Ti))¯=iATiπ(cA,i)+π(Vi)¯=iπ(Ci)+𝕋i.{\operatorname{Fl}}(X)=\bigcup_{i}\bigcup_{A\in T_{i}}\overline{\pi(A)}=\bigcup_{i}\bigcup_{A\in T_{i}}\overline{\pi(A+L(T_{i}))}=\bigcup_{i}\bigcup_{A\in T_{i}}\pi(c_{A,i})+\overline{\pi(V_{i})}=\bigcup_{i}\pi(C_{i})+\mathbb{T}_{i}.

References

  • [1] Martin Bays and Jonathan Kirby. Pseudo-exponential maps, variants, and quasiminimality. Algebra Number Theory, 12(3):493–549, 2018.
  • [2] Tien-Cuong Dinh and Duc-Viet Vu. Algebraic flows on commutative complex Lie groups. Comment. Math. Helv., 95(3):421–460, 2020.
  • [3] Francesco Paolo Gallinaro. Around exponential-algebraic closedness. April 2022.
  • [4] Ya’acov Peterzil and Sergei Starchenko. Algebraic and o-minimal flows on complex and real tori. Adv. Math., 333:539–569, 2018.
  • [5] Emmanuel Ullmo and Andrei Yafaev. Algebraic flows on abelian varieties. J. Reine Angew. Math., 741:47–66, 2018.
  • [6] B. I. Zilber. Generalized analytic sets. Algebra i Logika, 36(4):387–406, 478, 1997.