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

Equidistribution of polynomially bounded o-minimal curves in homogeneous spaces

Michael Bersudsky, Nimish A. Shah, and Hao Xing The Ohio State University, Columbus, OH 43210, USA [email protected], [email protected], and [email protected]
Abstract.

We extend Ratner’s theorem on equidistribution of individual orbits of unipotent flows on finite volume homogeneous spaces of Lie groups to trajectories of non-contracting curves definable in polynomially bounded o-minimal structures.

To be precise, let φ:[0,)SL(n,)\varphi:[0,\infty)\to\text{SL}(n,\mathbb{R}) be a continuous map whose coordinate functions are definable in a polynomially bounded o-minimal structure; for example, rational functions. Suppose that φ\varphi is non-contracting; that is, for any linearly independent vectors v1,,vkv_{1},\ldots,v_{k} in n\mathbb{R}^{n}, φ(t).(v1vk)↛0\varphi(t).(v_{1}\wedge\cdots\wedge v_{k})\not\to 0 as tt\to\infty. Then, there exists a unique smallest subgroup HφH_{\varphi} of SL(n,)\text{SL}(n,\mathbb{R}) generated by unipotent one-parameter subgroups such that φ(t)Hφg0Hφ\varphi(t)H_{\varphi}\to g_{0}H_{\varphi} in SL(n,)/Hφ\text{SL}(n,\mathbb{R})/H_{\varphi} as tt\to\infty for some g0SL(n,)g_{0}\in\text{SL}(n,\mathbb{R}).

Let GG be a closed subgroup of SL(n,)\text{SL}(n,\mathbb{R}) and Γ\Gamma be a lattice in GG. Suppose that φ([0,))G\varphi([0,\infty))\subset G. Then HφGH_{\varphi}\subset G, and for any xG/Γx\in G/\Gamma, the trajectory {φ(t)x:t[0,T]}\{\varphi(t)x:t\in[0,T]\} gets equidistributed with respect to the measure g0μLxg_{0}\mu_{Lx} as TT\to\infty, where LL is a closed subgroup of GG such that Hx¯=Lx\overline{Hx}=Lx and LxLx admits a unique LL-invariant probability measure, denoted by μLx\mu_{Lx}.

A crucial new ingredient in this work is proving that for any finite-dimensional representation VV of SL(n,)\text{SL}(n,\mathbb{R}), there exist T0>0T_{0}>0, C>0C>0, and α>0\alpha>0 such that for any vGv\in G, the map tφ(t)vt\mapsto\lVert\varphi(t)v\rVert is (C,α)(C,\alpha)-good on [T0,)[T_{0},\infty).

1. Introduction

Let GG be a Lie group and Γ\Gamma be a lattice in GG. We say that a probability measure μ\mu on X:=G/ΓX:=G/\Gamma is homogeneous if there exists an xXx\in X such that the orbit HxHx is closed in XX and μ\mu is the (unique) HH-invariant probability measure on the orbit HxHx. We also call the orbit HxHx a periodic orbit of HH and μ=:μHx\mu=:\mu_{Hx} the homogeneous probability measure supported on that periodic orbit. Marina Ratner, in proving Raghunathan’s conjectures stated in [Dan81], showed that a Borel probability measure μ\mu which is invariant and ergodic under the action of a AdG\text{Ad}_{G}-unipotent one-parameter subgroup is homogeneous, see [Rat91]. Using this measure classification and non-divergence property of unipotent orbits due to Margulis and Dani [Mar71, Dan86], Ratner [Rat91a] proved the following equidistribution result for unipotent flows: Let U:={u(t):t}U:=\{u(t):t\in\mathbb{R}\} be a one-parameter AdG\text{Ad}_{G}-unipotent subgroup of GG and xXx\in X. Then there exists a closed subgroup HH of GG containing UU such that HxH_{x} is a periodic orbit, and for any fCc(X)f\in C_{c}(X), 1T0Tf(u(t)x)𝑑tXf𝑑μHx\frac{1}{T}\int_{0}^{T}f(u(t)x)dt\to\int_{X}f\,d\mu_{Hx} as TT\to\infty.

More generally, for a continuous curve φ:[0,)G\varphi:[0,\infty)\to G, xXx\in X, and T>0T>0, consider the probability measure μT,φ,x\mu_{T,\varphi,x} on G/ΓG/\Gamma defined by:

(1.1) μT,φ,x(f)=1T0Tf(φ(t)x)𝑑t,fCc(X).\mu_{T,\varphi,x}(f)=\frac{1}{T}\int_{0}^{T}f(\varphi(t)x)\,dt,~{}\forall f\in C_{c}(X).

In [Sha96] it was shown that the equidistribution of unipotent flows due to Ratner generalizes to polynomial curves in GG. Namely, if GSL(n,)G\leq\text{SL}(n,\mathbb{R}) and φ:G\varphi:\mathbb{R}\to G is a map whose coordinate functions are polynonmial, then the measures μT,φ,x\mu_{T,\varphi,x} converge to φ(0)μHx\varphi(0)\mu_{Hx} as TT\to\infty, where μHx\mu_{Hx} is supported on a periodic orbit HxHx, for HGH\leq G being the smallest closed subgroup for which {φ(0)1φ(t):t}H\{\varphi(0)^{-1}\varphi(t):t\in\mathbb{R}\}\subset H such that the orbit HxHx is periodic.

More recently, Peterzil and Strachenko [PS18] studied such questions in the following setting: Let GG be a closed subgroup of the group of upper triangle unipotent subgroup of SL(n,)\text{SL}(n,\mathbb{R}) and Γ\Gamma be a (cocompact) lattice in GG. They proved that if SGS\subset G is definable in an o-minimal structure, then there exists a definable set SGS^{\prime}\subset G whose image in G/ΓG/\Gamma is the closure of the image of SS in G/ΓG/\Gamma. Moreover, they proved the following [PS18, Theorem 1.6]: Let φ:[0,)G\varphi:[0,\infty)\to G be a curve definable in a polynomially bounded o-minimal structure. Let xG/Γx\in G/\Gamma. If {φ(t)x:t0}\{\varphi(t)x:t\geq 0\} is dense in G/ΓG/\Gamma, then μT,φ,x\mu_{T,\varphi,x} converges to the GG-invariant probability on G/ΓG/\Gamma as TT\to\infty.

Our goal in this paper is to generalize the above equidistribution results for curves in an affine algebraic group such that the curves are definable in polynomially bounded o-minimal structures and satisfy an additional non-contraction condition.

We now proceed to give some basic definitions and describe our results.

1.1. O-minimal curves

We recall the basic definition of o-minimal structures. For more details on o-minimal structures, we refer the readers to [DM96, Dri98]. We will recall the required results when needed.

We only deal with o-minimal structures on the field of real numbers. The definition below is borrowed from an article of Dries and Miller [DM96].

Definition 1.1.

An o-minimal structure 𝒮\mathscr{S} on the real field is a sequence of families of sets

𝒮nsubsets of n,n,\mathscr{S}_{n}\subseteq\text{subsets of }\mathbb{R}^{n},~{}n\in\mathbb{N},

such that the following requirements are satisfied:

  1. (1)

    For each nn\in\mathbb{N}, 𝒮n\mathscr{S}_{n} is a boolean algebra of sets, and n𝒮n\mathbb{R}^{n}\in\mathscr{S}_{n}.

  2. (2)

    For each nn\in\mathbb{N}, 𝒮n\mathscr{S}_{n} contains the diagonals {(x1,,xn):xi=xj},\{(x_{1},...,x_{n}):x_{i}=x_{j}\}, for all 1i<jn1\leq i<j\leq n.

  3. (3)

    For each nn\in\mathbb{N}, if A𝒮nA\in\mathscr{S}_{n}, then A×𝒮n+1A\times\mathbb{R}\in\mathscr{S}_{n+1} and ×A𝒮n+1\mathbb{R}\times A\in\mathscr{S}_{n+1}.

  4. (4)

    For each nn\in\mathbb{N}, if A𝒮n+1A\in\mathscr{S}_{n+1}, then π(A)𝒮n\pi(A)\in\mathscr{S}_{n}, where π:n+1n\pi:\mathbb{R}^{n+1}\to\mathbb{R}^{n} is the projection to the first nn-coordinates.

  5. (5)

    𝒮3\mathscr{S}_{3} contains the graph of addition {(x,y,x+y):x,y}\{(x,y,x+y):x,y\in\mathbb{R}\} and the graph of multiplication {(x,y,xy):x,y}\{(x,y,xy):x,y\in\mathbb{R}\}.

  6. (6)

    𝒮1\mathscr{S}_{1} consists exactly of the finite unions of intervals (unbounded included) and singeltons.

We say that a set AnA\subset\mathbb{R}^{n} is definable in 𝒮\mathscr{S} if A𝒮nA\in\mathscr{S}_{n}. A function f:Bmf:B\to\mathbb{R}^{m} with BnB\subseteq\mathbb{R}^{n} is said to be definable if its graph is in 𝒮n+m\mathscr{S}_{n+m}.

  • An o-minimal structure is polynomially bounded, if for every definable function f:[0,)f:[0,\infty)\to\mathbb{R} there exists an rr\in\mathbb{R} such that f(x)=O(xr)f(x)=O(x^{r}) as xx\to\infty.

In many regards, functions definable in a polynomially bounded o-minimal structures behave like rational functions. We note the following result which we will use throughout the paper:

  • By [Mil94a]: If f:[0,)f:[0,\infty)\to\mathbb{R} definable in a polynomially bounded o-minimal structure 𝒮\mathscr{S}, then either ff is eventually constantly zero, or there exists r0r\neq 0 such that limtf(t)/tr=c0\lim_{t\to\infty}f(t)/t^{r}=c\neq 0. In the latter case, we define

    (1.2) deg(f):=r.\deg(f):=r.

    If ff is eventually constantly zero, we define deg(f)=\deg(f)=-\infty.

The following allows for constructing many examples of functions definable in a polynomially bounded o-minimal structure.

  • By [Mil94], there is a polynomially bounded o-minimal structure 𝒮\mathscr{S} such that the functions f(t)=trf(t)=t^{r} (t>0t>0), for rr\in\mathbb{R}, and all real analytic functions restricted to a compact set are definable in 𝒮\mathscr{S}.

  • If a finite collection of functions is definable, then every function in the algebra they generate is also definable in the same structure.

  • If f,gf,g are definable functions, then h(t):=f(t)/g(t)h(t):=f(t)/g(t), where tt is such that g(t)0g(t)\neq 0, is definable in the same structure.

  • Suppose that I,JI,J\subseteq\mathbb{R} and f:If:I\to\mathbb{R} and g:JIg:J\to I are definable. Then, the composition fgf\circ g is definable in the same structure.

  • Let II\subseteq\mathbb{R}. If f:If:I\to\mathbb{R} is definable and injective, then its inverse function is definable in the same structure.

For example, using the above one sees that

x(tan(π2xx+1)x2+2x+1+arcsin(x2x2+3)x4)π,x>0,x\mapsto\left(\tan\left(\frac{\pi}{2}\frac{x}{x+1}\right)\frac{x^{\sqrt{2}}+2}{x+1}+\arcsin\left(\frac{x^{2}}{x^{2}+3}\right)x^{4}\right)^{\pi},x>0,

is definable in a polynomially bounded o-minimal structure.

Remark 1.2.

For our equidistribution results, we will be mainly concerned with polynomially bounded o-minimal structures. As was noted in [PS18], the class of polynomially bounded o-minimal structures is a natural class to study equidistribution in view of the following. By [Mil94a], if an o-minimal structure is not polynomially bounded, then the exponential function is definable. As a consequence, log(t)\log(t) will also be definable. Notice that the measures on the circle defined by

1T0Tf(log(t+1)+)𝑑t,fC(/),\frac{1}{T}\int_{0}^{T}f(\log(t+1)+\mathbb{Z})\,dt,~{}f\in C(\mathbb{R}/\mathbb{Z}),

do not converge as TT\to\infty. Namely, for

φ(t):=[1log(t+1)01],\varphi(t):=\begin{bmatrix}1&\log(t+1)\\ 0&1\end{bmatrix},

and xx being the identity coset in SL(2,)/SL(2,)\text{SL}(2,\mathbb{R})/\text{SL}(2,\mathbb{Z}), the measures μT,φ,x\mu_{T,\varphi,x} will not converge.

1.2. o-minimal curves in affine algebraic groups

In this paper we will denote by 𝐆\mathbf{G} an affine algebraic group defined over \mathbb{R}, and by 𝐆()\mathbf{G}(\mathbb{R}) the \mathbb{R}-points of 𝐆\mathbf{G}. We view 𝐆()\mathbf{G}(\mathbb{R}) as a Lie group in the standard way via the subspace topology obtained by embedding algebraically 𝐆()SL(n,)M(n,)\mathbf{G}(\mathbb{R})\hookrightarrow\text{SL}(n,\mathbb{R})\subset\text{M}(n,\mathbb{R}).

Definition 1.3.

Let G:=𝐆()G:=\mathbf{G}(\mathbb{R}), and let [G]\mathbb{R}[G] be the coordinate ring of real polynomial functions on GG. We say that a curve φ:[0,)G\varphi:[0,\infty)\to G is definable in an o-minimal structure 𝒮\mathscr{S}, if the map tf(φ(t))t\mapsto f(\varphi(t)) on [0,)[0,\infty) is definable in 𝒮\mathscr{S}, for all f[G]f\in\mathbb{R}[G].

We observe that if VV is a finite dimensional rational representation, and φ:[0,)G\varphi:[0,\infty)\to G is definable in 𝒮\mathscr{S}, then the curve tφ(t).vt\mapsto\varphi(t).v is definable in 𝒮\mathscr{S} for all vVv\in V. In particular, if GG is embedded in SL(n,)M(n,)\text{SL}(n,\mathbb{R})\subset\mathrm{M}(n,\mathbb{R}), namely

(1.3) φ(t)=[φi,j(t)]1i,jn,t[0,),\varphi(t)=[\varphi_{i,j}(t)]_{1\leq i,j\leq n},~{}t\in[0,\infty),

then φ\varphi is definable in 𝒮\mathscr{S} if and only if each φi,j:[0,)\varphi_{i,j}:[0,\infty)\to\mathbb{R} is definable in 𝒮\mathscr{S}.

1.2.1. The non-contraction property

The following introduces a certain “non-contraction” property which narrows down the polynomially bounded o-minimal curves to a class of curves whose orbits on rational linear representations of GG have the (C,α)(C,\alpha)-good growth property introduced by Kleinbock and Margulis [KM98].

Definition 1.4.

We say that a definable curve φ:[0,)G\varphi:[0,\infty)\to G in an affine algebraic group G=𝐆()G=\mathbf{G}(\mathbb{R}) is non-contracting in GG if for all finite dimensional representations VV of GG defined over \mathbb{R}, it holds that limtφ(t).v0\lim_{t\to\infty}\varphi(t).v\neq 0 for all vV{0}v\in V\smallsetminus\{0\}.

Remark 1.5.

Suppose that GG is a unipotent group and φ:[0,)G\varphi:[0,\infty)\to G is a definable curve. Then φ\varphi is non-contracting. This is so because, for any algebraic action of an algebraic unipotent group on a finite-dimensional vector space, every orbit of a nonzero vector is Zariski closed, and in particular, the closure of the orbit does not contain the origin, see [Bir71, Theorem 12.1].

We now give a practical criterion that allows one to verify the non-contraction property for curves in SL(n,)\text{SL}(n,\mathbb{R}). Consider the exterior representation of SL(n,)\text{SL}(n,\mathbb{R}) on kn\bigwedge^{k}\mathbb{R}^{n} defined by

g.(v1v2vk):=gv1gv2gvk.g.(v_{1}\wedge v_{2}\wedge\cdots\wedge v_{k}):=gv_{1}\wedge gv_{2}\wedge\cdots\wedge gv_{k}.
Proposition 1.6.

Suppose that φ:[0,)SL(n,)\varphi:[0,\infty)\to\text{SL}(n,\mathbb{R}) is a continuous curve definable curve in an o-minimal structure. Then φ\varphi is non-contracting if and only if for all 1kn11\leq k\leq n-1 and linearly independent vectors v1,,vknv_{1},...,v_{k}\in\mathbb{R}^{n}, it holds that

(1.4) limtφ(t).(v1vk)0.\lim_{t\to\infty}\varphi(t).(v_{1}\wedge\cdots\wedge v_{k})\neq 0.

Due to the previous result, the following provides criteria for the non-contraction of definable curves in algebraic groups generated by unipotent one-parameter subgroups.

Proposition 1.7.

Let G=𝐆()G=\mathbf{G}(\mathbb{R}) be such that the identity component G0G^{0} is generated by unipotent one-parameter subgroups. Suppose ρ:GSL(n,)\rho:G\to\text{SL}(n,\mathbb{R}) is rational homomorphism with a finite kernel. Then, for any polynomially bounded definable curve φ:[0,)G\varphi:[0,\infty)\to G, if ρφ\rho\circ\varphi is non-contracting in SL(n,)\text{SL}(n,\mathbb{R}), then φ\varphi is non-contracting in GG.

1.3. The main equidistribution result

In Section 2.2, we will show that for a given continuous definable non-contracting curve φ:[0,)G\varphi:[0,\infty)\to G in an algebraic group G=𝐆()G=\mathbf{G}(\mathbb{R}), there exists a unique subgroup generated by one-parameter unipotent subgroups of GG, denoted by HφGH_{\varphi}\leq G and called the hull of φ\varphi, such that φ([0,))/Hφ\varphi([0,\infty))/H_{\varphi} is contained in a compact subset of G/HφG/H_{\varphi}. Also, we can choose a bounded definable curve β:[0,)G\beta:[0,\infty)\to G, which we refer to as a correcting curve such that β(t)φ(t)Hφ\beta(t)\varphi(t)\in H_{\varphi} for all tt; see Theorem 2.4 and Definition 2.5. We note that limtβ(t)=βG\lim_{t\to\infty}\beta(t)=\beta_{\infty}\in G by the Monotonicity Theorem [DM96, Theorem 4.1].

Theorem 1.8.

Let G:=𝐆()G:=\mathbf{G}(\mathbb{R}). Suppose that 𝒢G\mathcal{G}\leq G is a connected, closed Lie subgroup with a lattice Γ𝒢\Gamma\leq\mathcal{G}. Suppose that φ:[0,)G\varphi:[0,\infty)\to G is a continuous, unbounded, non-contracting curve in GG which is definable in a polynomially bounded o-minimal structure. Suppose that φ([0,))𝒢\varphi\left([0,\infty)\right)\subset\mathcal{G}. Let HφH_{\varphi} be the hull of φ\varphi in GG and let β\beta be a definable correcting curve in GG, so that βφHφ\beta\varphi\subset H_{\varphi}, see Definition 2.5. Then Hφ𝒢H_{\varphi}\subset\mathcal{G}, β([0,))𝒢\beta([0,\infty))\subset\mathcal{G}, and the following holds: Fix x0𝒢/Γx_{0}\in\mathcal{G}/\Gamma. Then,

limTμT,βφ,x0=μHφx0¯,\lim_{T\to\infty}\mu_{T,\beta\varphi,x_{0}}=\mu_{\overline{H_{\varphi}x_{0}}},

in the weak-\ast topology, where, in view of Ratner’s theorem, μHφx0¯\mu_{\overline{H_{\varphi}x_{0}}} is the LL-invariant probability measure on Hφx0¯=Lx0\overline{H_{\varphi}x_{0}}=Lx_{0} for a closed connected subgroup LL of 𝒢\mathcal{G}. In particular,

limTμT,φ,x0=β1μHφx0¯,\lim_{T\to\infty}\mu_{T,\varphi,x_{0}}=\beta_{\infty}^{-1}\mu_{\overline{H_{\varphi}x_{0}}},

where β=limtβ(t)\beta_{\infty}=\lim_{t\to\infty}\beta(t).

Remark 1.9.

the non-contraction assumption is necessary for Theorem 1.8 to hold in the above generality. We note that if φ\varphi is a contracting curve in SL(n,)\text{SL}(n,\mathbb{R}), then there exists xSL(n,)/SL(n,)x\in\text{SL}(n,\mathbb{R})/\text{SL}(n,\mathbb{Z}) such that limtφ(t)x=\lim_{t\to\infty}\varphi(t)x=\infty. To see the existence of such xx, since φ\varphi is contracting, by Proposition 1.6, there exists linearly independent vectors {v1,,vk}n\{v_{1},...,v_{k}\}\subseteq\mathbb{R}^{n} such that φ(t).(v1vk)0\varphi(t).(v_{1}\wedge\cdots\wedge v_{k})\to 0 as tt\to\infty. Identify SL(n,)/SL(n,)\text{SL}(n,\mathbb{R})/\text{SL}(n,\mathbb{Z}) with the space of unimodular lattices, and let xx be a lattice which includes {v1,,vk}\{v_{1},...,v_{k}\}. Then by Minkowski’s theorem, the length of the shortest nonzero vector(s) of the discrete subgroup Span{φ(t)v1,,φ(t)vk}φ(t)x0\text{Span}_{\mathbb{Z}}\{\varphi(t)v_{1},...,\varphi(t)v_{k}\}\subseteq\varphi(t)x_{0} goes to zero as tt\to\infty. By Mahler’s criterion, the curve φ(t)x0\varphi(t)x_{0} diverges to \infty in the space of unimodular lattices.

In the next result, proved in Section 7, we give an example of a definable curve in SL(n+1,)\text{SL}(n+1,\mathbb{R}) whose trajectory diverges in every irreducible representation of SL(n+1,)\text{SL}(n+1,\mathbb{R}); in other words, it is non-contracting, and its hull is SL(n+1,)\text{SL}(n+1,\mathbb{R}). As a consequence, every trajectory of such a curve in the space of unimodular lattice in n+1\mathbb{Z}^{n+1} gets equidistributed (Corollary 1.11).

Proposition 1.10.

Let f0,,fnf_{0},...,f_{n}, and h1,,hnh_{1},\ldots,h_{n} be continuous real functions on [0,)[0,\infty) definable in a polynomially bounded o-minimal structure. For t0t\geq 0, define

(1.5) φ(t):=[f0(t)f1(t)fn(t)h1(t)1hn(t)1]\varphi(t):=\begin{bmatrix}f_{0}(t)&f_{1}(t)&\cdots&f_{n}(t)\\ &h_{1}(t)^{-1}&&\\ &&\ddots&\\ &&&h_{n}(t)^{-1}\end{bmatrix}

Suppose that the functions hih_{i}’s and fif_{i}’s defining φ\varphi satisfy the following conditions:

  1. (1)

    f0=h1hnf_{0}=h_{1}\cdots h_{n}, degf0=n\deg f_{0}=n, and degh1deghn>0\deg h_{1}\geq\cdots\geq\deg h_{n}>0.

  2. (2)

    deg(f0+g)n\deg(f_{0}+g)\geq n for any gg in the linear span of f1,,fnf_{1},\ldots,f_{n}.

  3. (3)

    For any i{1,,n}i\in\{1,\ldots,n\}, deg(fi+g)>ni\deg(f_{i}+g)>n-i for any gg in the liner span of fi+1,,fnf_{i+1},\ldots,f_{n}.

Then the hull of φ\varphi is Hφ=SL(n+1,)H_{\varphi}=\text{SL}(n+1,\mathbb{R}).

For example, let f0(t)=tnf_{0}(t)=t^{n}. Suppose f1,,fnf_{1},\dots,f_{n} are rational functions such that degf1>>degfn\deg f_{1}>\cdots>\deg f_{n}, and degfin\deg f_{i}\neq n and degfi>ni\deg f_{i}>n-i for all 1in1\leq i\leq n. Let hi(t)=trih_{i}(t)=t^{r_{i}} for some r1rn>0r_{1}\geq\cdots\geq r_{n}>0 such that i=1nri=n\sum_{i=1}^{n}r_{i}=n. Then, these functions satisfy the conditions of Proposition 1.10.

Using Theorem 1.8, we obtain the following:

Corollary 1.11.

Let GG be a Lie group and Γ\Gamma be a lattice in GG. Suppose that ρ:SL(n+1,)G\rho:\text{SL}(n+1,\mathbb{R})\to G is a continuous homomorphism. Let φ:[0,)SL(n+1,)\varphi:[0,\infty)\to\text{SL}(n+1,\mathbb{R}) be a curve satisfying the conditions of Proposition 1.10. Then for any xG/Γx\in G/\Gamma, and any bounded continuous function ff on G/ΓG/\Gamma,

limT1T0Tf(φ(t)x0)𝑑t=μX(f),\lim_{T\to\infty}\frac{1}{T}\int_{0}^{T}f(\varphi(t)x_{0})dt=\mu_{X}(f),

where μX\mu_{X} is the homogeneous probability measure on the homogeneous space ρ(SL(n+1,))x¯\overline{\rho(\text{SL}(n+1,\mathbb{R}))x}.

1.4. Growth of o-minimal functions

We now describe the (C,α)(C,\alpha)-good property concerning the growth of certain families of functions definable in polyonomially bounded o-minimal structures. The property is well known for polynomials, and it is new in the o-minimal setting. It is key for the technique used in this paper.

We need the following definition.

Definition 1.12.

Let 𝒱\mathcal{V} be finite dimensional vector space spanned over \mathbb{R} by real functions definable in a polynomially bounded o-minimal structure. Identify 𝒱\mathcal{V} with N\mathbb{R}^{N} by choosing a basis for 𝒱\mathcal{V}. A subset 𝒱\mathscr{F}\subseteq\mathcal{V} is called a closed definable cone if it is a closed definable subset of N\mathbb{R}^{N} such that if vv\in\mathscr{F} then αv\alpha v\in\mathscr{F} for all scalars α0\alpha\geq 0.

Example 1.13.

Let ψ:[0,)GL(m,)\psi:[0,\infty)\to\text{GL}(m,\mathbb{R}) be a curve definable in a polynomially bounded o-minimal structure, and let \|\cdot\| be the Euclidean norm. Then, the vector space

𝒱:=Span{ψ(t)v2:vm}\mathcal{V}:=\text{Span}_{\mathbb{R}}\{\|\psi(t)\textbf{v}\|^{2}:\textbf{v}\in\mathbb{R}^{m}\}

is finite dimensional, and :={ψ(t)v2:vm}\mathscr{F}:=\{\|\psi(t)\textbf{v}\|^{2}:\textbf{v}\in\mathbb{R}^{m}\} is a closed definable cone.

Theorem 1.14.

((C,α)(C,\alpha)-good property). Let 𝒱\mathcal{V} be a finite-dimensional real vector space spanned by functions f:[0,)f:[0,\infty)\to\mathbb{R} definable in a polynomially bounded o-minimal structure. Suppose that 𝒱\mathscr{F}\subset\mathcal{V} is a closed definable cone such that for all f{0}f\in\mathscr{F}\smallsetminus\{0\} it holds that

limtf(t)0.\lim_{t\to\infty}f(t)\neq 0.

Then, there exist C>0C>0, α>0\alpha>0 and T01T_{0}\geq 1 such that for any f{0}f\in\mathscr{F}\smallsetminus\{0\}, ff is (C,α)(C,\alpha)-good in [0,)[0,\infty); that is, ϵ>0\forall\epsilon>0, and for every bounded interval I[T0,)I\subseteq[T_{0},\infty)

(1.6) |{tI:|f(t)|ϵ}|C(ϵfI)α|I|.|\{t\in I:|f(t)|\leq\epsilon\}|\leq C\left(\frac{\epsilon}{\|f\|_{I}}\right)^{\alpha}\cdot|I|.

We note the following result is an immediate consequence of Theorem 1.14.

Proposition 1.15.

Let G:=𝐆()G:=\mathbf{G}(\mathbb{R}) and let φ:[0,)G\varphi:[0,\infty)\to G be a non-contracting curve definable in a polynomially bounded o-minimal structure. Suppose that VV is a finite dimensional rational representation of GG over \mathbb{R}. Fix a norm on \|\cdot\| on VV. Then, there exist T0>0T_{0}>0, C>0C>0, and α>0\alpha>0 such that for all 0vV{0}0\neq v\in V\smallsetminus\{0\} the function Θv(t):=φ(t).v\Theta_{v}(t):=\|\varphi(t).v\| is (C,α)(C,\alpha)-good in [T0,)[T_{0},\infty).

1.5. Proof ideas and structure of the paper

Our main theorem will be proved by the following strategy which is common in many works on equidistribution in homogeneous spaces, see e.g. [DM93, MS95]:

  1. (1)

    Proving that there is no escape of mass – that is, any limiting measure μ\mu of the probability measures μT\mu_{T} is a probability measure.

  2. (2)

    Showing that any limiting measure μ\mu of the measures μT\mu_{T} is invariant under a group generated by unipotent one-parameter subgroups.

  3. (3)

    Studying the ergodic decomposition of a limiting measure μ\mu. Since μ\mu is invariant under a unipotent group, every ergodic component appearing in the ergodic decomposition of μ\mu is homogeneous by Ratner’s measure classification theorem. The theorem is proved once it is verified that μ\mu equals to exactly one uniquely deteremined ergodic component. This will step will be treated using the linearization technique of [DM93].

To establish unipotent invariance, we will use an idea similar to the one used for polynomial curves [Sha94]. This idea generalizes to the o-minimal setting via the notion of the Peterzil-Steinhorn group (see Section 5), which was also used in [PS18].

The (C,α)(C,\alpha)-good property will be used to prove the non-escape of mass (Section 4) and to apply the linearization technique for the ergodic decomposition of our limiting measure (Section 6).

1.6. Some conventions about notation

We will denote the Zariski closure of a subset AA of an affine space by zcl(A)\text{zcl}(A). For a linear Lie group HH, we denote by HuH_{u} the subgroup of HH generated by all one-parameter unipotent subgroups contained in HH. We will write G:=𝐆()G:={\bf G}(\mathbb{R}) to mean that GG is the set of \mathbb{R} points of an algebraic group 𝐆{\bf G} defined over \mathbb{R}.

Acknowledgement.

We thank Chris Miller for helpful discussions.

2. Consequences of the non-contracting property

2.1. The criterion for the non-contracting property

Our goal here is to prove Proposition 1.6, namely to reduce the verification of the non-contracting property for any vector in any representation to decomposable vectors in the exterior of the standard representation of SL(n,)\text{SL}(n,\mathbb{R}). Our proof follows from the observations in [SY24, Section 2], which uses Kempf’s numerical criteria on geometric invariant theory [Kem78].

We recall the following preliminaries. Let S denote the full diagonal subgroup of SL(n,)\text{SL}(n,\mathbb{R}). The group of strictly upper triangular matrices in SL(n,)\text{SL}(n,\mathbb{R}), denoted by NN, is a maximal unipotent subgroup of SL(n,)\text{SL}(n,\mathbb{R}) normalized by S. Let X(S)X^{\ast}(\textbf{S}) denote the group of algebraic homomorphisms from S to \mathbb{R}^{\ast} defined over \mathbb{R}. Then X(S)X^{\ast}(\textbf{S}) is a free abelian group on (n1)(n-1)-generators, and we treat it as an additive group. For each i{1,,n1}i\in\{1,\ldots,n-1\}, let αiX(S)\alpha_{i}\in X^{\ast}(\textbf{S}) be defined by αi(diag(t1,,tn))=ti/ti+1\alpha_{i}(\text{diag}(t_{1},\ldots,t_{n}))=t_{i}/t_{i+1}. Then Δ={αi:1in1}\Delta=\{\alpha_{i}:1\leq i\leq n-1\} is the set of simple roots on S corresponding to the choice of the maximal unipotent subgroup NN. For each 1in11\leq i\leq n-1, let μiX(S)\mu_{i}\in X^{\ast}(\textbf{S}) be defined by μi(diag(t1,,tn))=t1ti\mu_{i}(\text{diag}(t_{1},\ldots,t_{n}))=t_{1}\cdots t_{i}. Then {μi:1in1}\{\mu_{i}:1\leq i\leq n-1\} is the set of fundamental characters (weights) of 𝐒\mathbf{S} with respect to our choice of the simple roots. For each 1in11\leq i\leq n-1, let WiW_{i} denote the ii-th exterior of the standard representation of SL(n,)\text{SL}(n,\mathbb{R}) on n\mathbb{R}^{n}, called a fundamental representation of SL(n,)\text{SL}(n,\mathbb{R}). So Wi=inW_{i}=\wedge^{i}\mathbb{R}^{n}, and for any v1viWiv_{1}\wedge\cdots\wedge v_{i}\in W_{i} and gSL(n,)g\in\text{SL}(n,\mathbb{R}), g.(v1vi)=gv1gvig.(v_{1}\wedge\cdots\wedge v_{i})=gv_{1}\wedge\cdots\wedge gv_{i}. Let wi=e1eiw_{i}=e_{1}\wedge\cdots\wedge e_{i}. Then wiw_{i} is fixed by NN, and S acts on the line wi\mathbb{R}w_{i} via the fundamental character μi\mu_{i}.

A dominant integral character (weight) is a non-negative integral combination of fundamental characters. Suppose χ=m1μ1++mn1μn1\chi=m_{1}\mu_{1}+\cdots+m_{n-1}\mu_{n-1}, where mi0m_{i}\in\mathbb{Z}_{\geq 0}. For each ii, let WimiW_{i}^{\otimes m_{i}} denote the tensor product of mim_{i}-copies of WiW_{i}, and wimi:=wiwiWimiw_{i}^{m_{i}}:=w_{i}\otimes\cdots\otimes w_{i}\in W_{i}^{\otimes m_{i}}. Consider the tensor product representation of SL(n,)\text{SL}(n,\mathbb{R}) on Wχ=W1m1Wn1mn1W_{\chi}=W_{1}^{\otimes m_{1}}\otimes\cdots\otimes W_{n-1}^{\otimes m_{n-1}}, and let wχ=w1m1wn1mn1w_{\chi}=w_{1}^{m_{1}}\otimes\cdots\otimes w_{n-1}^{m_{n-1}}. Then wχw_{\chi} is fixed by NN, and S acts on the line wχ\mathbb{R}w_{\chi} via the character χ\chi.

Proof of Proposition 1.6.

Let VV be a representation of G:=SL(n,)G:=\text{SL}(n,\mathbb{R}). Suppose vVv\in V is a a nonzero such that φ(t)v0\varphi(t)v\to 0 as tt\to\infty. We call such a vv a GG-unstable vector in VV. So, by [SY24, Remark 2.3], there exist g0Gg_{0}\in G, a dominant integral character χX(S)\chi\in X^{\ast}(\textbf{S}), and constants β>0\beta>0 and C>0C>0 such that for any gGg\in G,

(2.1) gg0.wχCg.vβ.\lVert gg_{0}.w_{\chi}\rVert\leq C\lVert g.v\rVert^{\beta}.

Therefore, limtφ(t)g0.wχ=0\lim_{t\to\infty}\varphi(t)g_{0}.w_{\chi}=0. Let A=S0A=\textbf{S}^{0}, the component of the identity in S. Let K=SO(n)K=\text{SO}(n). Then by Iwasawa decomposition, G=KANG=KAN. For each tt, we express φ(t)g0=ktatnt\varphi(t)g_{0}=k_{t}a_{t}n_{t}, where ktKk_{t}\in K, atAa_{t}\in A and ntNn_{t}\in N. Without loss of generality, we may assume that the norm on WχW_{\chi} is KK-invariant. Since, NN fixes wχw_{\chi},

φ(t)g0.wχ=at.wχ=χ(at)wχ.\lVert\varphi(t)g_{0}.w_{\chi}\rVert=\lVert a_{t}.w_{\chi}\rVert=\chi(a_{t})\lVert w_{\chi}\rVert.

Hence, since limtφ(t)g0.wχ=0\lim_{t\to\infty}\varphi(t)g_{0}.w_{\chi}=0, we get limtχ(at)=0\lim_{t\to\infty}\chi(a_{t})=0. Now

χ(at)=μ1(at)m1μn1(at)mn1.\chi(a_{t})=\mu_{1}(a_{t})^{m_{1}}\cdots\mu_{n-1}(a_{t})^{m_{n-1}}.

Since mi0m_{i}\geq 0, after passing to a subsequence, we can pick i{1,,n1}i\in\{1,\ldots,n-1\} and a sequence tlt_{l}\to\infty such that limlμi(atl)=0\lim_{l\to\infty}\mu_{i}(a_{t_{l}})=0. Therefore, since NN fixes wiw_{i}, and the norm is KK-invariant,

φ(tl)g0.wi=atl.wi=μi(atl)wi0, as l.\lVert\varphi(t_{l})g_{0}.w_{i}\rVert=\lVert{a_{t_{l}}.w_{i}}\rVert=\mu_{i}(a_{t_{l}})\lVert w_{i}\rVert\to 0\text{, as $l\to\infty$.}

Since the coordinate functions of tφ(t)g0.wit\to\varphi(t)g_{0}.w_{i} are definable in an o-minimal structure, we get limtφ(t)g0.wi=0\lim_{t\to\infty}\varphi(t)g_{0}.w_{i}=0. Let vj=g0ejv_{j}=g_{0}e_{j} for each jj. Since wi=e1eiw_{i}=e_{1}\wedge\cdots\wedge e_{i}, φ(t).(g0wi)=φ(t).(v1vi)0\varphi(t).(g_{0}w_{i})=\varphi(t).(v_{1}\wedge\cdots\wedge v_{i})\to 0 as tt\to\infty. This contradicts our assumption 1.4. ∎

The following statement will be used in the proof of Lemma 7.1, which will needed for proving Propositions 1.10.

Lemma 2.1.

Suppose that φ:[0,)SL(n,)\varphi:[0,\infty)\to\text{SL}(n,\mathbb{R}) is a continuous curve such that

(2.2) limtφ(t).(v1vk)=,\lim_{t\to\infty}\varphi(t).(v_{1}\wedge\dots\wedge v_{k})=\infty,

for all 1kn11\leq k\leq n-1 and linearly independent vectors v1,,vknv_{1},...,v_{k}\in\mathbb{R}^{n}. Then for every finite dimensional rational representation VV and an vVv\in V such that the curve φ(t).v,t0\varphi(t).v,t\geq 0 is bounded in VV, it holds that SL(n,).v\text{SL}(n,\mathbb{R}).v is Zariski closed.

Proof.

We will use the notations above. Suppose that VV is a finite dimensional rational representation of SL(n,)\text{SL}(n,\mathbb{R}) and let vVv\in V such that SL(n,).v\text{SL}(n,\mathbb{R}).v is not Zariski closed. By [SY24, Corollary 2.5] and by [SY24, Remark 2.3], there exist a g0SL(n,)g_{0}\in\text{SL}(n,\mathbb{R}), a dominant integral character χX(S)\chi\in X^{\ast}(\textbf{S}), and a constant β>0\beta>0 such that for any R>0R>0 there is a constant C>0C>0 with the following property:

(2.3) if g.vR, then gg0.wχCg.vβ.\displaystyle\text{if $\|g.v\|\leq R$, then }\lVert gg_{0}.w_{\chi}\rVert\leq C\lVert g.v\rVert^{\beta}.

Now suppose that φ:[0,)SL(n,)\varphi:[0,\infty)\to\text{SL}(n,\mathbb{R}) is a continuous curve satisfying the diverging property (2.2) and assume for contradiction that φ(t).v,t0\varphi(t).v,~{}t\geq 0 is bounded in VV. Then, by (2.3), we get that φ(t)g0.wχ,t>0\lVert\varphi(t)g_{0}.w_{\chi}\rVert,~{}t>0 is bounded. As in the proof of Proposition 1.6, we use Iwasawa decomposition to express φ(t)g0=ktatnt,t>0\varphi(t)g_{0}=k_{t}a_{t}n_{t},t>0, where ktKk_{t}\in K, atAa_{t}\in A and ntNn_{t}\in N, and observe that

φ(t)g0.wχ=at.wχ=χ(at)wχ.\lVert\varphi(t)g_{0}.w_{\chi}\rVert=\lVert a_{t}.w_{\chi}\rVert=\chi(a_{t})\lVert w_{\chi}\rVert.

We claim that limtχ(at)=,\lim_{t\to\infty}\chi(a_{t})=\infty, which is a contradiction. Indeed, we have

χ(at)=μ1(at)m1μn1(at)mn1,\chi(a_{t})=\mu_{1}(a_{t})^{m_{1}}\cdots\mu_{n-1}(a_{t})^{m_{n-1}},

for mi0m_{i}\geq 0, and

φ(t)g0.wi=at.wi=μi(at)wi,\lVert\varphi(t)g_{0}.w_{i}\rVert=\lVert{a_{t}.w_{i}}\rVert=\mu_{i}(a_{t})\lVert w_{i}\rVert,

where wi=e1eiw_{i}=e_{1}\wedge\cdots\wedge e_{i} and 1in11\leq i\leq n-1. Finally, due to (2.2), we have limtφ(t)g0.wi=.\lim_{t\to\infty}\lVert\varphi(t)g_{0}.w_{i}\rVert=\infty.

2.2. On the hull and the correcting curve

In order to define the hull and to obtain its usesful properties we will need the following notion.

Definition 2.2.

Let G:=𝐆()G:=\mathbf{G}(\mathbb{R}). We say that a closed subgroup HH is observable in GG if there is a finite-dimensional rational representation VV and an vVv\in V such that H={gG:g.v=v}H=\{g\in G:g.v=v\}.

The notion of observable groups was introduced [BHM63]. More precisely, [BHM63] defines observable groups as the algebraic subgroups of algebraic groups for which every finite dimensional rational representation extends to a finite dimensional rational representation of the ambient group. The above definition turns out equivalent to the extension property [BHM63, Theorem 8] (see [TB05, Theorem 9] for this statement for non-algebraically closed fields).

We note:

  • [Gro97, Corollary 2.8]: If HH is an algebraic subgroup over \mathbb{R} whose radical is unipotent, then HH is observable.

Definition 2.3.

We say that a curve φG\varphi\subset G is bounded modulo a closed subgroup HGH\leq G if there exists a compact subset KGK\subseteq G such that φKH\varphi\subseteq KH. Alternatively, φ\varphi is bounded modulo HH if its image in G/HG/H is contained in a compact subset of G/HG/H.

Here is our main theorem of this section.

Theorem 2.4.

Let G:=𝐆()G:=\mathbf{G}(\mathbb{R}). Suppose that φ:[0,)G\varphi:[0,\infty)\to G is a continuous, non-contracting curve definable in an o-minimal structure 𝒮\mathscr{S}. Then, there exists a unique closed connected (with respect to the real topology) subgroup HφGH_{\varphi}\leq G of the smallest dimension such that HφH_{\varphi} is generated by one-parameter unipotent subgroup and such that φ\varphi is bounded modulo HφH_{\varphi}. Moreover, the following properties hold:

  1. (1)

    The subgroup HφH_{\varphi} is minimal in the following sense: if HH^{\prime} is an observable group such that φ\varphi is bounded modulo HH^{\prime}, then HφHH_{\varphi}\leq H^{\prime}. Moreover, if VV is a rational representation, and vVv\in V is such that {φ(t).v:t0}\{\varphi(t).v:t\geq 0\} is bounded in VV, then φ\varphi is bounded modulo the isotropy group of vv and Hφ.v={v}H_{\varphi}.v=\{v\}.

  2. (2)

    There is a bounded, continuous curve β:[0,)G\beta:[0,\infty)\to G definable in 𝒮\mathscr{S} such that β(t)φ(t)Hφ\beta(t)\varphi(t)\in H_{\varphi} for all t0t\geq 0.

  3. (3)

    If VV is a rational finite dimensional representation of GG over \mathbb{R}, then for vVv\in V either Hφ.v={v}H_{\varphi}.v=\{v\} or limtβ(t)φ(t).v=\lim_{t\to\infty}\beta(t)\varphi(t).v=\infty.

In view of the above statement, we can now define the hull.

Definition 2.5.

Let G=𝐆()G=\mathbf{G}(\mathbb{R}) and and let φ:[0,)G\varphi:[0,\infty)\to G be a continuous non-contracting curve definable in an o-minimal structure 𝒮\mathscr{S}. We define the hull HφH_{\varphi} to be the smallest connected closed subgroup generated by unipotent elements such that φ\varphi is bounded modulo HφH_{\varphi}, and we call a curve β:[0,)G\beta:[0,\infty)\to G a correcting curve if β(t)φ(t)Hφ\beta(t)\varphi(t)\in H_{\varphi} for all t0t\geq 0.

Remark 2.6.

We note that if φ\varphi is as in the above theorem and β\beta is a definable correcting curve, then the hull of ψ(t):=β(t)φ(t)\psi(t):=\beta(t)\varphi(t) is the same as the hull of φ\varphi, and the correcting curve of ψ\psi can be chosen to be the constant identity matrix. This follows from the uniqueness of the hull.

Remark 2.7.

Suppose that φ:[0,)SL(n,)\varphi:[0,\infty)\to\text{SL}(n,\mathbb{R}) is a curve whose entry functions φi,j\varphi_{i,j} generate an algebra which only consists of either constant functions or functions diverging to \infty, e.g., when tφi,j(t)t\mapsto\varphi_{i,j}(t) are polynomials for all i,ji,j. Then it follows by Theorem 2.4 that φ(0)1φ([0,))Hφ\varphi(0)^{-1}\varphi([0,\infty))\subset H_{\varphi}. A correcting curve is given by the constant curve β(t)=φ(0)1,t0\beta(t)=\varphi(0)^{-1},\forall t\geq 0.

Before proving Theorem 2.4, we establish that the hull is intrinsically defined.

Lemma 2.8.

Let G1,G2G_{1},G_{2} be the real points of algebraic groups over \mathbb{R}, and let f:G1G2f:G_{1}\to G_{2} be a rational homomorphism. Suppose that φ:[0,)G1\varphi:[0,\infty)\to G_{1} is a non-contracting curve definable in an o-minimal structure 𝒮\mathscr{S}. Then f(Hφ)=Hfφf(H_{\varphi})=H_{f\circ\varphi}.

Proof.

We start with showing that Hfφf(Hφ)H_{f\circ\varphi}\subseteq f(H_{\varphi}). Let β:[0,)G1\beta:[0,\infty)\to G_{1} be a correcting curve of φ\varphi. Since HφH_{\varphi} is generated by unipotents, we get that f(Hφ)f(H_{\varphi}) is generated by unipotents. Moreover, the curve fφf\circ\varphi is bounded modulo f(Hφ)f(H_{\varphi}). In fact, tf(β(t))t\mapsto f(\beta(t)) is bounded, and for all t0t\geq 0:

f(β(t)φ(t))=f(β(t))f(φ(t))f(Hφ).f(\beta(t)\varphi(t))=f(\beta(t))f(\varphi(t))\in f(H_{\varphi}).

Thus, by minimality of the hull, we conclude that Hfφf(Hφ)H_{f\circ\varphi}\subseteq f(H_{\varphi}).

We now show the other inclusion. Since zcl(Hfφ)\text{zcl}(H_{f\circ\varphi}) is observable, we may choose a rational representation ρ:G2GL(V)\rho:G_{2}\to\text{GL}(V) and a vVv\in V such that

zcl(Hfφ)={gG2:ρ(g)v=v}.\text{zcl}(H_{f\circ\varphi})=\{g\in G_{2}:\rho(g)v=v\}.

Since fφf\circ\varphi is bounded modulo HfφH_{f\circ\varphi}, we get that the trajectory

ρ((fφ)(t))v=(ρf)(φ(t))v,t0,\rho((f\circ\varphi)(t))v=(\rho\circ f)(\varphi(t))v,~{}t\geq 0,

is bounded in VV. Now, η:=ρf:G1GL(V)\eta:=\rho\circ f:G_{1}\to\text{GL}(V) is a representation of G1G_{1} such that tη(φ(t))v,t0t\mapsto\eta(\varphi(t))v,~{}t\geq 0 is bounded. By Theorem 2.4(1), we obtain that ρ(f(Hφ))v=η(Hφ)v={v}\rho(f(H_{\varphi}))v=\eta(H_{\varphi})v=\{v\}. That is, f(Hφ)f(H_{\varphi}) is contained in the isotropy group of vv which equals to zcl(Hfφ)\text{zcl}(H_{f\circ\varphi}). ∎

2.3. Proving Theorem 2.4

We begin with the following “curve correcting” lemma.

Lemma 2.9.

Consider an algebraic action of G:=𝐆()G:=\mathbf{G}(\mathbb{R}) on an affine or projective variety 𝒵\mathcal{Z}. Let φ:[0,)G\varphi:[0,\infty)\to G be a continuous curve definable in an o-minimal structure 𝒮\mathscr{S}. Suppose that x𝒵x\in\mathcal{Z} and g0Gg_{0}\in G are such that

limtφ(t).x=g0.x.\lim_{t\to\infty}\varphi(t).x=g_{0}.x.

Then there exists a continuous 𝒮\mathscr{S}-definable curve δ:[0,)G\delta:[0,\infty)\to G such that limtδ(t)=e\lim_{t\to\infty}\delta(t)=e, and

δ(t)φ(t).x=g0.x,t0.\delta(t)\varphi(t).x=g_{0}.x,\forall t\geq 0.
Proof.

Identify GG as a closed subgroup of SL(n,)\text{SL}(n,\mathbb{R}). Consider the following definable set:

𝒟\displaystyle\mathcal{D} :={(t,g):t>0,gG,gIn<1 and g01φ(t).x=g.x}.\displaystyle:=\{(t,g):t>0,g\in G,\|g-I_{n}\|<1\text{ and }g_{0}^{-1}\varphi(t).x=g.x\}.

Here \|\cdot\| is a definable norm (e.g. the sum of squares norm). Since the orbit map is an open map (maps open sets in GG to open sets in G.xG.x with respect to the induced topology from 𝒵\mathcal{Z}, see [PRR23, Corollary 3.9]), and since

limtg01φ(t).x=x,\lim_{t\to\infty}g_{0}^{-1}\varphi(t).x=x,

we get that for all tt large enough, say tT0t\geq T_{0}, there exists gGg\in G such that (t,g)𝒟(t,g)\in\mathcal{D}. Using the choice function theorem [DM96, Theorem 4.5], there is a definable curve δ~(t):[T0,)G\tilde{\delta}(t):[T_{0},\infty)\to G such that (t,δ~(t))𝒟,tT0(t,\tilde{\delta}(t))\in\mathcal{D},~{}\forall t\geq T_{0}. In particular,

δ~(t).x=g01φ(t).x,tT0,\tilde{\delta}(t).x=g_{0}^{-1}\varphi(t).x,\forall t\geq T_{0},

and δ~\tilde{\delta} is bounded. Since the curve δ~(t)\tilde{\delta}(t) is bounded, by the Monotonicity Theorem [DM96, Theorem 4.1] we have that limtδ~(t)=δ0\lim_{t\to\infty}\tilde{\delta}(t)=\delta_{0}. Note that δ0.x=x\delta_{0}.x=x. Consider, δ(t):=g0(δ~(t)δ01)1g01\delta(t):=g_{0}(\tilde{\delta}(t)\delta_{0}^{-1})^{-1}g_{0}^{-1} for tT0t\geq T_{0}. Then,

δ(t)φ(t).x=g0.x,tT0.\delta(t)\varphi(t).x=g_{0}.x,\forall t\geq T_{0}.

By cell decomposition [DM96, Theorem 4.2], δ(t)\delta(t) is continuous in [T0,)[T^{\prime}_{0},\infty). We now extend δ\delta to the interval [0,T0][0,T^{\prime}_{0}] by putting

δ(t):=δ(T0)φ(T0)φ(t)1,tT0.\delta(t):=\delta(T^{\prime}_{0})\varphi(T^{\prime}_{0})\varphi(t)^{-1},t\leq T^{\prime}_{0}.

Then δ(t)φ(t).x=g0.x,t0\delta(t)\varphi(t).x=g_{0}.x,\forall t\geq 0. Since φ\varphi is continuous and definable, δ:[0,)G\delta:[0,\infty)\to G is continuous, definable and bounded. ∎

We now observe the following.

Lemma 2.10.

Let G:=𝐆()G:=\mathbf{G}(\mathbb{R}). Suppose that φ:[0,)G\varphi:[0,\infty)\to G is a non-contracting, continuous curve, definable is an o-minimal structure 𝒮\mathscr{S}. Suppose that VV is a finite dimensional rational representation and assume that there is an vV{0}v\in V\smallsetminus\{0\} such that φ(t).v\varphi(t).v is bounded. Then, there exists a g0Gg_{0}\in G such that

limtφ(t).v=g0.v.\lim_{t\to\infty}\varphi(t).v=g_{0}.v.

In particular, there exists a bounded, continuous curve β:[0,)SL(n,)\beta:[0,\infty)\to\text{SL}(n,\mathbb{R}) definable in 𝒮\mathscr{S}, such that β(t)φ(t).v=v,t0\beta(t)\varphi(t).v=v,\forall t\geq 0.

Proof.

Since φ:[0,)G\varphi:[0,\infty)\to G is definable, and since VV is a rational representation, it follows that φ(t).v\varphi(t).v is a bounded definable curve. By the monotonicity theorem (see [DM96, Section 4]), it follows that there is an vVv^{\prime}\in V such that limtφ(t).v=v\lim_{t\to\infty}\varphi(t).v=v^{\prime}. We claim that vG.vv^{\prime}\in G.v. Denote S:=zcl(G.v)G.vS:=\text{zcl}(G.v)\smallsetminus G.v, and suppose for contradiction that vSv^{\prime}\in S. Then, by Lemma A.1 which is a slight modification of [Kem78, Lemma 1.1(b)], there exists a rational representation WW of GG and an GG-equivariant polynomial map P:VWP:V\to W such that P(v)0P(v)\neq 0 and P(S)=0P(S)=0. But then

limtφ(t).P(v)=P(φ(t).v)=P(v)=0,\lim_{t\to\infty}\varphi(t).P(v)=P(\varphi(t).v)=P(v^{\prime})=0,

which is a contradiction since φ\varphi is non-contracting. Thus, we conclude that there is an g0Gg_{0}\in G such that

limtφ(t).v=g0.v.\lim_{t\to\infty}\varphi(t).v=g_{0}.v.

The existence of the bounded curve β\beta is obtained by Lemma 2.9. ∎

Lemma 2.11.

Let G:=𝐆()G:=\mathbf{G}(\mathbb{R}). Suppose that φ:[0,)G\varphi:[0,\infty)\to G is a non-contracting, continuous curve, definable in an o-minimal structure. Let H1,H2GH_{1},H_{2}\leq G be two observable groups such that the image of φ\varphi is bounded in G/HiG/H_{i} for i{1,2}i\in\{1,2\}. Then, the image of φ\varphi is bounded in G/(H1H2)G/(H_{1}\cap H_{2}).

Proof.

For i{1,2}i\in\{1,2\}, let ViV_{i} be a finite dimensional rational representations such that Hi={gG:g.vi=vi}H_{i}=\{g\in G:g.v_{i}=v_{i}\} where viViv_{i}\in V_{i}. Consider the direct-sum representation V1V2V_{1}\oplus V_{2}, and notice that

H1H2={gG:g.(v1,v2)=(v1,v2)}.H_{1}\cap H_{2}=\{g\in G:g.(v_{1},v_{2})=(v_{1},v_{2})\}.

Since the image of φ\varphi is bounded in G/HiG/H_{i}, φ(t).vi\varphi(t).v_{i} is bounded for all i{1,2}i\in\{1,2\}. In particular, φ(t).(v1,v2)\varphi(t).(v_{1},v_{2}) is bounded in V1V2V_{1}\oplus V_{2}. Then, by Lemma 2.10 there is a bounded continuous curve δ:[0,)G\delta:[0,\infty)\to G such that δ(t)φ(t).(v1,v2)=(v1,v2)\delta(t)\varphi(t).(v_{1},v_{2})=(v_{1},v_{2}) for all t0t\geq 0. Namely, δ(t)φ(t)H1H2\delta(t)\varphi(t)\in H_{1}\cap H_{2} for all t0t\geq 0, which shows that claim. ∎

Corollary 2.12.

Let G:=𝐆()G:=\mathbf{G}(\mathbb{R}). Suppose that φ:[0,)G\varphi:[0,\infty)\to G is a non-contracting, continuous curve definable in an o-minimal structure. Then there exists a unique observable subgroup H~φG\tilde{H}_{\varphi}\leq G of the smallest dimension such that φ\varphi is bounded in G/H~φG/\tilde{H}_{\varphi}, and if HH is an observable subgroup such that φ\varphi is bounded in G/HG/H, then H~φH\tilde{H}_{\varphi}\leq H.

Proof.

Let H0H_{0} be a Zariski connected observable group of the smallest dimension such that φ\varphi is bounded in G/H0G/H_{0}. If HH is any connected observable group such that φ\varphi is bounded in G/HG/H, then by Lemma 2.11, we get that φ\varphi is bounded in G/(H0H)G/(H_{0}\cap H). If HH does not include H0H_{0}, then HH0H\cap H_{0} is a proper subgroup of dimension strictly smaller than dimH0\dim H_{0}, which is impossible. ∎

Corollary 2.13.

Let G:=𝐆()G:=\mathbf{G}(\mathbb{R}). Suppose that φ:[0,)G\varphi:[0,\infty)\to G is a non-contracting, continuous curve, definable in an o-minimal structure. Let H~φG\tilde{H}_{\varphi}\leq G be the minimal observable group such that φ\varphi is bounded modulo H~φ\tilde{H}_{\varphi}. Denote Hφ:=H~φ0H_{\varphi}:=\tilde{H}_{\varphi}^{0}, the identity component of H~φ\tilde{H}_{\varphi}.

Then, there exists a bounded, continuous curve β:[T0,)G\beta:[T_{0},\infty)\to G such that β(t)φ(t)Hφ\beta(t)\varphi(t)\in H_{\varphi} for all t0t\geq 0. Moreover, if VV is a finite-dimensional rational representation of GG and vVv\in V is such that φ(t).v\varphi(t).v is bounded, then β(t)φ(t).v=v,tT0\beta(t)\varphi(t).v=v,\forall t\geq T_{0}.

Proof.

By Lemma 2.10, we obtain a continuous definable curve β~:[0,)SL(n,)\tilde{\beta}:[0,\infty)\to\text{SL}(n,\mathbb{R}) such that β~(t)φ(t)H~φ\tilde{\beta}(t)\varphi(t)\in\tilde{H}_{\varphi} for all t0t\geq 0. Since φ\varphi is continuous, it is contained in a connected component of HφH_{\varphi}. By choosing a suitable h0Hφh_{0}\in H_{\varphi} we get that for β(t):=h0β~(t)\beta(t):=h_{0}\tilde{\beta}(t) it holds β(t)φ(t)Hφ\beta(t)\varphi(t)\in H_{\varphi} for all t0t\geq 0. Now, if VV is a finite-dimensional rational representation such that φ(t).v\varphi(t).v is bounded, then by Lemma 2.10, we conclude that φ\varphi is bounded modulo the observable group H:={gG:g.v=v}H:=\{g\in G:g.v=v\}. Hence, by minimality of H~φ\tilde{H}_{\varphi}, we get that β(t)φ(t)HφH\beta(t)\varphi(t)\in H_{\varphi}\leq H. ∎

In order to establish Theorem 2.4 it remains to prove that the identity component of the above minimal observable group is generated by one-parameter unipotent subgroups.

Lemma 2.14.

Let G:=𝐆()G:=\mathbf{G}(\mathbb{R}). Suppose that φ:[0,)G\varphi:[0,\infty)\to G be a non-contracting, continuous curve, definable in an o-minimal structure, and let H~φ\tilde{H}_{\varphi} be the smallest observable subgroup such that φ\varphi is bounded modulo H~φ\tilde{H}_{\varphi}. Then Hφ=H~φ0H_{\varphi}=\tilde{H}_{\varphi}^{0} is generated by one-parameter unipotent subgroups.

Proof.

Denote ψ(t):=β(t)φ(t),t0\psi(t):=\beta(t)\varphi(t),t\geq 0, where β\beta is a correcting curve of φ\varphi. Let LH~φL\leq\tilde{H}_{\varphi} be the Zariski closure of the subgroup generated by all one-parameter unipotent subgroups of H~φ\tilde{H}_{\varphi}. We will now show that the image of φ\varphi in H~φ/L\tilde{H}_{\varphi}/L is bounded, and since LL is observable [Gro97, Corollary 2.8], we will get that H~φ=L\tilde{H}_{\varphi}=L, which proves the claim.

Now, note that LL is a normal subgroup, so that H~φ/L\tilde{H}_{\varphi}/L is an affine algebraic group. Since LL contains the unipotent radical of H~φ\tilde{H}_{\varphi}, we get that H~φ/L\tilde{H}_{\varphi}/L is reductive. Then, there is an almost direct product decomposition H~φ/L=ZD\tilde{H}_{\varphi}/L=ZD, where ZZ is the center, which is a torus, and DD is the derived group, which is a normal semi-simple group. Notice that each simple factor of DD must be compact since a non-compact simple group is generated by unipotents. Namely, DD is compact. Let q:H~φH~φ/Lq:\tilde{H}_{\varphi}\to\tilde{H}_{\varphi}/L be the natural quotient map. To finish the proof, we show below that for every character χ:H~φ/L×\chi:\tilde{H}_{\varphi}/L\to\mathbb{R}^{\times} it holds that χq(ψ(t))=1,t0\chi\circ q(\psi(t))=1,\forall t\geq 0. This will complete the proof since then qψ([0,))ZaDq\circ\psi([0,\infty))\subseteq Z_{a}D, where ZaZZ_{a}\leq Z is the maximal anisotropic compact torus subgroup of ZZ. Let χ:H~φ/L×\chi:\tilde{H}_{\varphi}/L\to\mathbb{R}^{\times} be a character, and consider the representation of H~φ/L\tilde{H}_{\varphi}/L on 2\mathbb{R}^{2} defined by

h¯.(x,y):=(χ(h¯)x,χ(h¯)1y),(x,y)2,h¯H~φ/L.\overline{h}.(x,y):=\left(\chi(\overline{h})x,\chi(\overline{h})^{-1}y\right),~{}(x,y)\in\mathbb{R}^{2},~{}\overline{h}\in\tilde{H}_{\varphi}/L.

Then H~φ\tilde{H}_{\varphi} acts on 2\mathbb{R}^{2} be pre-composing with qq. Namely, h.(x,y):=q(h).(x,y)h.(x,y):=q(h).(x,y). Since H~φG\tilde{H}_{\varphi}\leq G is observable, the above representation comes from a representation of GG, see [TB05, Theorem 9]. Now, fix two non-zero real numbers x0,y0x_{0},y_{0}. Because ψ\psi is non-contracting, we get that ψ(t).(x0,0)\psi(t).(x_{0},0) and ψ(t).(0,y0)\psi(t).(0,y_{0}) are bounded away from 0. But this is only possible if ψ(t).(x0,y0)\psi(t).(x_{0},y_{0}) is bounded. So by Corollary 2.13 we obtain that ψ(t).(x0,y0)=(x0,y0)\psi(t).(x_{0},y_{0})=(x_{0},y_{0}). Thus, χq(ψ(t))=1,t0\chi\circ q(\psi(t))=1,~{}\forall t\geq 0. ∎

2.4. The non-contraction property under homomorphisms

In this section we show that the homomorphic image of a definable non-contracting curve is non-contracting, Lemma 2.15, and also we show that if a definable curve in a homomorphic image is non-contracting , then it has a definable non-contracting lift, Lemma 2.19. Finally, we establish that the non-contracting property is intrinsic for curves in groups whose radical is unipotent, Proposition 2.20.

Lemma 2.15.

Let G1,G2G_{1},G_{2} be the real points of two affine algebraic groups over \mathbb{R}, and let f:G1G2f:G_{1}\to G_{2} be a homomorphism of algebraic groups. Let φ:[0,)G1\varphi:[0,\infty)\to G_{1} be a continuous, non-contracting curve in G1G_{1} definable in 𝒮\mathscr{S}. Then fφf\circ\varphi is a continuous, non-contracting curve in G2G_{2} definable in 𝒮\mathscr{S}.

Proof.

Let ρ:G2GL(V)\rho:G_{2}\to\text{GL}(V) be a rational representation of G2G_{2}. Let vVv\in V. Then ρf:G1GL(V)\rho\circ f:G_{1}\to\text{GL}(V) is a rational representation of G1G_{1}. Let vVv\in V. If ρ(fφ(t))v0\rho(f\circ\varphi(t))v\to 0 as tt\to\infty, then (ρf)(φ(t))v0(\rho\circ f)(\varphi(t))v\to 0 as tt\to\infty. Since φ\varphi is non-contracting, we get v=0v=0. ∎

Lemma 2.16.

Let G:=𝐆()G:=\mathbf{G}(\mathbb{R}) and suppose that UGU\leq G is a normal, unipotent algebraic subgroup. Let q:GG/Uq:G\to G/U be the natural map. Then, a curve φ:[0,)G\varphi:[0,\infty)\to G is not contracting in GG if and only if qφq\circ\varphi is not contracting in G/UG/U.

Proof.

In view of Lemma 2.15, it suffices to show that if qφq\circ\varphi is non-contracting, then φ\varphi is non-contracting. Suppose for contradiction that qφq\circ\varphi is non-contracting, but φ\varphi is contracting. Let VV be a rational representation of the smallest dimension such that there exists a vV{0}v\in V\smallsetminus\{0\} with limtφ(t).v=0\lim_{t\to\infty}\varphi(t).v=0. Since UU is a unipotent group, the subspace of UU-fixed points VU:={xV:u.x=x,uU}V^{U}:=\{x\in V:u.x=x,\forall u\in U\} is of positive dimension. Since UU is a normal subgroup, VUV^{U} is GG-invariant. Then GG acts on V/VUV/V^{U}, and the natural quotient map π:VV/VU\pi:V\to V/V^{U} is GG-equivariant. Now,

limtφ(t).π(v)=limtπ(φ(t).v)=0.\lim_{t\to\infty}\varphi(t).\pi(v)=\lim_{t\to\infty}\pi(\varphi(t).v)=0.

Since dim(V/VU)<dim(V)\dim(V/V^{U})<\dim(V), we get that π(v)=0\pi(v)=0. Namely, vVUv\in V^{U}, and again, since VV is of smallest dimension with a φ(t)\varphi(t)-contracting nonzero vector, we get V=VUV=V^{U}. Since UU acts trivially on VV, the action of GG on VV factors through the action of G/UG/U. Therefore, φ(t)v=(qφ(t))v\varphi(t)v=(q\circ\varphi(t))v. Since φ(t)v0\varphi(t)v\to 0 as tt\to\infty, and qφq\circ\varphi is non-contracting, we get a contradiction. ∎

Lemma 2.17.

Let G1,G2G_{1},G_{2} be the real points of algebraic groups over \mathbb{R}. Suppose that f:G1G2f:G_{1}\to G_{2} is a homomorphism and φ:[0,)f(G1)\varphi:[0,\infty)\to f(G_{1}) is a curve definable in an o-minimal structure 𝒮\mathscr{S}. Then there exists a definable lift ψ:[0,)G1\psi:[0,\infty)\to G_{1}. Namely, ψ\psi is definable in 𝒮\mathscr{S}, and fψ=φf\circ\psi=\varphi.

Proof.

Consider the definable set

𝒟:={(t,g):gG1,t0,f(g)=φ(t)}.\mathcal{D}:=\{(t,g):g\in G_{1},~{}t\geq 0,~{}f(g)=\varphi(t)\}.

By the choice function theorem [DM96, Theorem 4.5], there exists a definable curve ψ:[0,)G1\psi:[0,\infty)\to G_{1} such that (t,ψ(t))𝒟(t,\psi(t))\in\mathcal{D} for all t0t\geq 0. ∎

Lemma 2.18.

Let G1G_{1}, G2G_{2} be the \mathbb{R}-points of irreducible semi-simple groups over \mathbb{R}, and let f:G1G2f:G_{1}\to G_{2} be homomorphism with finite kernel. Let ψ:[0,)G1\psi:[0,\infty)\to G_{1} be a definable curve. Suppose that fψf\circ\psi is non-contracting for G2G_{2}, then ψ\psi is non-contracting for G1G_{1}.

Proof.

Let VV be a finite dimensional irreducible rational representation of G1G_{1}. Then, by Schur’s lemma the center of G1G_{1} acts via a character on VV\otimes\mathbb{C}. Then, kerf\ker f, which is central in G1G_{1}, acts trivially on the nn-fold tensor product W=n(V)W=\otimes^{n}(V\otimes\mathbb{C}), where nn is the order of kerf\ker f. Therefore, the action of G1G_{1} on WW factors through an action of G1/(kerf)=G2G_{1}/(\ker f)=G_{2} on WW. Let vVv\in V. If ψ(t).v0\psi(t).v\to 0, then fψ(t).(vv)0f\circ\psi(t).(v\otimes\cdots\otimes v)\to 0 as tt\to\infty. But, since fψf\circ\psi is non-contracting, we get that v=0v=0. ∎

Lemma 2.19.

Let G1,G2G_{1},G_{2} be the real points of two affine algebraic groups over \mathbb{R}, and let f:G1G2f:G_{1}\to G_{2} be a surjective homomorphism of algebraic groups. Let φ:[0,)G2\varphi:[0,\infty)\to G_{2} be a non-contracting curve in G2G_{2} definable in 𝒮\mathscr{S}. Then, φ\varphi has a definable lift ψ:[0,)G1\psi:[0,\infty)\to G_{1} that is a non-contracting in G1G_{1}.

Proof.

In this proof, all subgroups considered are the open subgroups of \mathbb{R}-points of algebraic groups defined over \mathbb{R}.

Without loss of generality, since the hull is generated by one-parameter unipotent subgroups, by modifying φ\varphi with a bounded correcting curve and replacing G2G_{2} by (G2)u(G_{2})_{u}, the subgroup of G2G_{2} generated by unipotent one-parameter subgroups, we may assume that the radical of G2G_{2} is unipotent, call it U2U_{2}. Let G¯2=G2/U2\bar{G}_{2}=G_{2}/U_{2} and q2:G2G¯2q_{2}:G_{2}\to\bar{G}_{2} be the quotient homomorphism. Let U1U_{1} be the unipotent radical of G1G_{1}. Then, there exists a reductive subgroup M1M_{1} of G1G_{1} such that G1=M1U1G_{1}=M_{1}U_{1} and M1U1={e}M_{1}\cap U_{1}=\{e\}, see [Rag72, Page 11]. Since ff is surjective, f(U1)f(U_{1}) is a unipotent normal subgroup of G2G_{2}, and hence f(U1)U2f(U_{1})\subset U_{2}. In fact, f(U1)=U2f(U_{1})=U_{2}. This can be proved as follows. f1(U2)=(M1f1(U2))U1f^{-1}(U_{2})=(M_{1}\cap f^{-1}(U_{2}))U_{1}. Since f1(U2)f^{-1}(U_{2}) is a normal subgroup of G1G_{1}, M1f1(U2)M_{1}\cap f^{-1}(U_{2}) is reductive. So f(M1f1(U2))f(M_{1}\cap f^{-1}(U_{2})) being a reductive subgroup of U2U_{2}, it is trivial. Since ff is surjective, we get f(U1)=U2f(U_{1})=U_{2}. Therefore, q2f:M1G¯2q_{2}\circ f:M_{1}\to\bar{G}_{2} is surjective. Since G¯2\bar{G}_{2} is semisimple, q2f([M1,M1])=G¯2q_{2}\circ f([M_{1},M_{1}])=\bar{G}_{2}. Let L2=[M1,M1]ker(q2f)L_{2}=[M_{1},M_{1}]\cap\ker(q_{2}\circ f). Since L2L_{2} is a normal subgroup of the semisimple group [M1,M1][M_{1},M_{1}], there exists a semisimple normal subgroup L1L_{1} of [M1,M1][M_{1},M_{1}] such that [M1,M1]=L1L2[M_{1},M_{1}]=L_{1}L_{2}, where L1L2L_{1}\cap L_{2} is a finite central subgroup of L1L_{1}. Therefore, q2f:L1G¯2q_{2}\circ f:L_{1}\to\bar{G}_{2} is a surjective homomorphism whose kernel is L1L2L_{1}\cap L_{2}. Also, since G¯2\bar{G}_{2} is generated by unipotent one-parameter subgroups, it has no compact simple normal subgroup. Therefore, L1L_{1} also does not have a compact normal subgroup, and L1L_{1} is generated by unipotent one-parameter subgroups.

Let G~1=L1U1\tilde{G}_{1}=L_{1}U_{1}. Since q2f(L1)=G2/U2q_{2}\circ f(L_{1})=G_{2}/U_{2} and f(U1)=U2f(U_{1})=U_{2}, we have f:G~1G2f:\tilde{G}_{1}\to G_{2} is a surjective map. Therefore, by Lemma 2.17, there exists a definable ψ:[0,)G~1\psi:[0,\infty)\to\tilde{G}_{1} such that fψ=φf\circ\psi=\varphi. Let q1:G~1G~1/U1q_{1}:\tilde{G}_{1}\to\tilde{G}_{1}/U_{1} be the natural quotient map. Since q1:L1G~1/U1q_{1}:L_{1}\to\tilde{G}_{1}/U_{1} is surjective (it is an isomorphism), let ψ~:[0,)L1\tilde{\psi}:[0,\infty)\to L_{1} be a definable curve such that q1ψ~=q1ψq_{1}\circ\tilde{\psi}=q_{1}\circ\psi. Since, U1kerq2fU_{1}\subset\ker q_{2}\circ f, q2fq_{2}\circ f factors through q1q_{1}. Therefore, we have (q2f)ψ~=(q2f)ψ=q2φ(q_{2}\circ f)\circ\tilde{\psi}=(q_{2}\circ f)\circ\psi=q_{2}\circ\varphi. Now q2f:L1G¯2q_{2}\circ f:L_{1}\to\bar{G}_{2} is surjective with finite central kernel, and q2φq_{2}\circ\varphi is definable and non-contracting in G¯2\bar{G}_{2}. Therefore, by Lemma 2.18, ψ~\tilde{\psi} is non-contracting in L1L_{1}. Therefore, q1ψ=q1ψ~q_{1}\circ\psi=q_{1}\circ\tilde{\psi} is non-contracting in G~1/U1\tilde{G}_{1}/U_{1}. Therefore, by Lemma 2.16, ψ\psi is non-contracting in G~1\tilde{G}_{1}. Finally, by Lemma 2.15 we get that ψ\psi is non-contracting in G1G_{1}. ∎

The following observation extends Lemma 2.18. Notice that Proposition 1.7 from the introduction is a direct consequence of the proposition below.

Proposition 2.20.

Suppose that G1G_{1} is a real algebraic group generated by unipotent one-parameter subgroups. Let f:G1G2f:G_{1}\to G_{2} be a homomorphism of real algebraic groups with finite kernel. Then a definable curve φ:[0,)G1\varphi:[0,\infty)\to G_{1} is non-contracting in G1G_{1} if and only if fφf\circ\varphi is non-contracting in G2G_{2}.

Proof.

If φ\varphi is non-contracting in G1G_{1}, then by Lemma 2.15, it follows that fφf\circ\varphi is non-contracting in G2G_{2}.

Now suppose that fφf\circ\varphi is non-contracting in G2G_{2}. Since G1G_{1} is generated by unipotent one-parameter subgroups, f(G1)f(G_{1}) is a subgroup of G2G_{2} generated by unipotent one-parameter subgroups. Therefore, the solvable radical of f(G1)f(G_{1}) is unipotent. Hence, f(G1)f(G_{1}) is an observable subgroup of G2G_{2}, see [Gro97, Corollary 2.8]. Therefore, given any finite-dimensional rational representation of VV of f(G1)f(G_{1}), VV is a subrepresentation of a finite-dimensional representation WW of G2G_{2} restricted to f(G1)f(G_{1}). Let vVv\in V. Suppose that fφ(t)v0f\circ\varphi(t)v\to 0. Since vVWv\in V\subset W and fφf\circ\varphi is non-contracting in G2G_{2}, we conclude that v=0v=0. This proves that fφf\circ\varphi is non-contracting for f(G1)f(G_{1}). Therefore, replacing G2G_{2} by f(G1)f(G_{1}), we may assume that ff is surjective. Therefore, by Lemma 2.19 there exists a lift ψ:[0,)G1\psi:[0,\infty)\to G_{1} of fφf\circ\varphi which is definable and non-contracting in G1G_{1}. By the Monotonicty Theorem [DM96, Theorem 4.1], there is a T00T_{0}\geq 0 such that φ\varphi and ψ\psi are continuous on [T0,)[T_{0},\infty). Let z=φ(T0)ψ(T0)1kerfz=\varphi(T_{0})\psi(T_{0})^{-1}\in\ker f. So, tzψ(t)t\mapsto z\psi(t) and tφ(t)t\mapsto\varphi(t) are two lifts of fφf\circ\varphi from [T0,)[T_{0},\infty) to G1G_{1} from the point φ(T0)G1\varphi(T_{0})\in G_{1}. Since kerf\ker f is finite, ff is a covering map, we get φ(t)=zψ(t)\varphi(t)=z\psi(t) for all t[0,)t\in[0,\infty). Since ψ\psi is non-contracting, so is φ\varphi. ∎

3. (C,α)(C,\alpha)-good property

In our setting, a property more fundamental than the (C,α)(C,\alpha)-good property (Theorem 1.14) is the following.

Definition 3.1.

Let δ(0,1]\delta\in(0,1]. A family \mathscr{F} of real functions defined on [T0,)[T_{0},\infty) is called δ\delta-good if there exists a constant M(δ)M(\delta) depending only on δ\delta such that

(3.1) fIfIδM(δ),\frac{\|f\|_{I}}{\|f\|_{I_{\delta}}}\leq M(\delta),

for all ff\in\mathscr{F} and all bounded sub-intervals IδI[T0,]I_{\delta}\subset I\subseteq[T_{0},\infty] satisfying |Iδ|=δ|I||I_{\delta}|=\delta|I|.

Such inequalities are well-known in the literature as Remez-type inequalities; see [Rem36]. We have the following theorem, which will imply Theorem 1.14.

Theorem 3.2.

Let 𝒱\mathcal{V} be a finite-dimensional vector space of real functions defined on [0,)[0,\infty) which are definable in a polynomially bounded o-minimal structure. Suppose that 𝒱\mathscr{F}\subseteq\mathcal{V} is a closed definable cone such that for all f{0}f\in\mathscr{F}\smallsetminus\{0\},

limtf(t)0.\lim_{t\to\infty}f(t)\neq 0.

Then, there exists T0>0T_{0}>0 such that for all δ(0,1]\delta\in(0,1] and all f{0}f\in\mathscr{F}\smallsetminus\{0\} it holds that ff restricted to [T0,)[T_{0},\infty) is δ\delta-good.

We now proceed to prove Theorem 1.14 by assuming Theorem 3.2, and the rest of the section will be dedicated to proving Theorem 3.2.

3.1. Proving Theorem 1.14 via δ\delta-goodness

We first note the following Corollary from Theorem 3.2.

Corollary 3.3.

Let 𝒱\mathcal{V} be a finite dimensional vector space of real functions on [0,)[0,\infty) which are definable in a polynomially bounded o-minimal structure. Suppose that 𝒱\mathscr{F}\subseteq\mathcal{V} is a closed definable cone such that for all f{0}f\in\mathscr{F}\smallsetminus\{0\},

limtf(t)0.\lim_{t\to\infty}f(t)\neq 0.

Let T01T_{0}\geq 1 such that the outcome of Theorem 3.2 holds. Then, there exists m,r>0\mathrm{m},r>0 such that for all x1x\geq 1, f{0}f\in\mathscr{F}\smallsetminus\{0\} and II[T0,)I^{\prime}\subseteq I\subseteq[T_{0},\infty) such that |I||I|x\frac{|I|}{|I^{\prime}|}\leq x it holds that

(3.2) fImxrfI.\|f\|_{I}\leq\mathrm{m}x^{r}\|f\|_{I^{\prime}}.
Proof.

Consider the definable set:

S:={(x,y):\displaystyle S:=\{(x,y): x,y>0, such that fIfIy,0f,\displaystyle x,y>0\text{, such that }\frac{\|f\|_{I}}{\|f\|_{I^{\prime}}}\leq y,\,\forall 0\neq f\in\mathscr{F},
II[T0,),|I||I|x}.\displaystyle I^{\prime}\subseteq I\subseteq[T_{0},\infty),\,\frac{|I|}{|I^{\prime}|}\leq x\}.

By Theorem 3.2, the projection of SS to the first coordinate includes [1,).[1,\infty). Then, by the definable choice theorem (see [DM96, Section 4]), there exists a function ϕ:[1,)2\phi:[1,\infty)\to\mathbb{R}^{2} definable in the same polynomially bounded o-minimal structure, such that

(x,ϕ(x))S,x1.(x,\phi(x))\in S,~{}\forall x\geq 1.

Namely for all x>1x>1, f{0}f\in\mathscr{F}\smallsetminus\{0\} and II[T0,)I^{\prime}\subseteq I\subseteq[T_{0},\infty) such that |I||I|x\frac{|I|}{|I^{\prime}|}\leq x it holds that

fIϕ(x)fI.\|f\|_{I}\leq\phi(x)\|f\|_{I^{\prime}}.

Since ϕ\phi is polynomially bounded, there exists r>0r>0 and T11T_{1}\geq 1 such that ϕ(x)xrc0\frac{\phi(x)}{x^{r}}\leq c_{0} for all xT1x\geq T_{1}, for some c0>0c_{0}>0. Finally, since ϕ(x)xr\frac{\phi(x)}{x^{r}} is bounded in [1,T1][1,T_{1}], the result follows. ∎

We are now ready to prove the (C,α)(C,\alpha)-good property. Our proof is based on [KM98, proof of Proposition 3.2].

Proof of Theorem 1.14.

By o-minimality, the number of connected components of the family of definable (in parameters) sub-level sets

{tI:|f(t)|ϵ},ϵ>0,I[T0,),f{0},\{t\in I:|f(t)|\leq\epsilon\},~{}\epsilon>0,~{}I\subset[T_{0},\infty),~{}f\in\mathscr{F}\smallsetminus\{0\},

is bounded uniformly, say by KK. Now fix I[T0,),ϵ>0I\subset[T_{0},\infty),~{}\epsilon>0 and an f{0}f\in\mathscr{F}\smallsetminus\{0\}. Let

I{tI:|f(t)|ϵ}I^{\prime}\subseteq\{t\in I:|f(t)|\leq\epsilon\}

be an interval of maximum length. Then

LK|I|,L\leq K|I^{\prime}|,

where

L:=|{tI:|f(t)|ϵ}|.L:=|\{t\in I:|f(t)|\leq\epsilon\}|.

The latter inequality implies,

|I||I||I|L/K.\frac{|I|}{|I^{\prime}|}\leq\frac{|I|}{L/K}.

By Corollary 3.3, we get

fIm(|I|L/K)rfIm(|I|L/K)rϵ.\|f\|_{I}\leq\mathrm{m}\left(\frac{|I|}{L/K}\right)^{r}\|f\|_{I^{\prime}}\leq\mathrm{m}\left(\frac{|I|}{L/K}\right)^{r}\epsilon.

Reordering the latter inequality, we get that

|{tI:|f(t)|ϵ}|=Lm1rK(ϵfI)1r|I|.|\{t\in I:|f(t)|\leq\epsilon\}|=L\leq\mathrm{m}^{\frac{1}{r}}K\left(\frac{\epsilon}{\|f\|_{I}}\right)^{\frac{1}{r}}|I|.

3.2. Proving Theorem 3.2

The space of polynomials of bounded degrees is the basic example of δ\delta-good functions. We provide a proof for the sake of completeness.

Proposition 3.4 ([DM93, Lemma 4.1]).

Fix nn\in\mathbb{N}, and consider :=Span{1,x,,xn}{0}\mathscr{F}:=\text{Span}_{\mathbb{R}}\{1,x,...,x^{n}\}\smallsetminus\{0\}. Then \mathscr{F} is δ\delta-good on \mathbb{R}. More precisely:

(3.3) c0+c1x++cnxnIc0+c1x++cnxnIδ(n+1)nnδn,\frac{\|c_{0}+c_{1}x+\cdots+c_{n}x^{n}\|_{I}}{\|c_{0}+c_{1}x+\cdots+c_{n}x^{n}\|_{I_{\delta}}}\leq(n+1)\frac{n^{n}}{\delta^{n}},

for all c0,c1,,cnc_{0},c_{1},...,c_{n}\in\mathbb{R} not all zero, interval II\subset\mathbb{R} and sub-interval IδII_{\delta}\subset I satisfying |Iδ|=δ|I|>0.|I_{\delta}|=\delta|I|>0.

Proof.

Let f(x):=c0+c1x++cnxnf(x):=c_{0}+c_{1}x+\cdots+c_{n}x^{n} and I=[a,a+T]I=[a,a+T]. Let at0a+TδTa\leq t_{0}\leq a+T-\delta T, and denote ti:=t0+iδTnt_{i}:=t_{0}+\frac{i\delta T}{n} for i=0,1,,ni=0,1,...,n, at0<t1<<tna+Ta\leq t_{0}<t_{1}<\cdots<t_{n}\leq a+T, and let Iδ:=[t0,tn]I_{\delta}:=[t_{0},t_{n}]. Then by polynomial interpolation of ff at points t0,t1,,tnt_{0},t_{1},...,t_{n}, we have

(3.4) f(t)=j=0nij(tti)ij(tjti)f(tj).f(t)=\sum_{j=0}^{n}\frac{\prod_{i\neq j}(t-t_{i})}{\prod_{i\neq j}(t_{j}-t_{i})}f(t_{j}).

Now,

|ij(tti)ij(tjti)|nnδn,tI,\left|\frac{\prod_{i\neq j}(t-t_{i})}{\prod_{i\neq j}(t_{j}-t_{i})}\right|\leq\frac{n^{n}}{\delta^{n}},~{}\forall t\in I,

and thus, by triangle inequality,

|f(t)|(n+1)nnδnmax|f(ti)|(n+1)nnδnfIδ,tI.\left|f(t)\right|\leq(n+1)\frac{n^{n}}{\delta^{n}}\max{\left|f(t_{i})\right|}\leq(n+1)\frac{n^{n}}{\delta^{n}}\|f\|_{I_{\delta}},~{}\forall t\in I.

We will need the following facts:

  • For any definable function f:[0,)f:[0,\infty)\to\mathbb{R}, there exits t00t_{0}\geq 0 such that ff differentiable on [t0,)[t_{0},\infty) and ff^{\prime} is continuous and definable on [t0,)[t_{0},\infty) (see [DM96, Cell decompsition, Section 4]).

  • For a definable function f:[0,)f:[0,\infty)\to\mathbb{R} in a polynomially bounded o-minimal structure with f(t)trf(t)\sim t^{r} where r0r\neq 0, we have f(t)rtr1, as tf^{\prime}(t)\sim rt^{r-1},\text{ as }t\to\infty, see [Mil94b, Proposition 3.1]. In particular, if deg(f)0,\deg(f)\neq 0, then deg(f)=deg(f)1\deg(f^{\prime})=\deg(f)-1.

Lemma 3.5.

For i=0,1,,Ni=0,1,\ldots,N, let Fi:[0,)F_{i}:[0,\infty)\to\mathbb{R} be a function definable in a polynomially bounded o-minimal structure such that FiF_{i} is not eventually the zero function. Suppose that the degrees deg(F0),deg(F1),,deg(FN)\deg(F_{0}),\deg(F_{1}),\ldots,\deg(F_{N}) are all distinct. Then, there exists a T0>0T_{0}>0 such that for any interval I[T0,)I\subset[T_{0},\infty) it holds that

c0F0+c1F1++cNFNI>0\|c_{0}F_{0}+c_{1}F_{1}+...+c_{N}F_{N}\|_{I}>0

for all (c0,c1,,cN)0(c_{0},c_{1},...,c_{N})\neq 0.

Proof.

By o-minimality, the definable set F01(0)F_{0}^{-1}(0) has finitely many connected components. Since F0F_{0} is not eventually the zero function, for some T0>0T_{0}>0, F0(t)0F_{0}(t)\neq 0 for all tT0t\geq T_{0}. Thus for all I[T0,)I\subset[T_{0},\infty) we have

c0F0+c1F1++cNFNI>0c0+c1F1F0++cNFNF0I>0.\left\|c_{0}F_{0}+c_{1}F_{1}+...+c_{N}F_{N}\right\|_{I}>0\iff\left\|c_{0}+c_{1}\frac{F_{1}}{F_{0}}+...+c_{N}\frac{F_{N}}{F_{0}}\right\|_{I}>0.

Clearly, since the degrees of F0,F1,,FNF_{0},F_{1},...,F_{N} are distinct, the degrees of 1,F1/F0,F2/F0,FN/F01,F_{1}/F_{0},F_{2}/F_{0},\cdots F_{N}/F_{0} are distinct. Thus, it is sufficient to prove the statement under the assumption that F0(t)1F_{0}(t)\equiv 1.

We now prove the statement for N=1N=1, and then argue by induction. Since deg(F1)deg(F0)=0\deg(F_{1})\neq\deg(F_{0})=0, we let T0>0T_{0}>0 be large enough such that F1(t)0F_{1}^{\prime}(t)\neq 0 for all tT0t\geq T_{0}. Assume that for an interval I[T0,)I\subset[T_{0},\infty) we have c0+c1F1I=0\|c_{0}+c_{1}F_{1}\|_{I}=0. Since c0+c1F1(t)=0c_{0}+c_{1}F_{1}(t)=0 for all tIt\in I, we get that

ddt(c0+c1F1(t))=c1F1(t)=0,tI,\frac{d}{dt}\left(c_{0}+c_{1}F_{1}(t)\right)=c_{1}F_{1}^{\prime}(t)=0,~{}\forall t\in I,

and as F1(t)0F_{1}^{\prime}(t)\neq 0 for all t[T0,)t\in[T_{0},\infty), it follows that c1=0c_{1}=0. As a consequence, c0=0c_{0}=0.

Now let N2N\geq 2, and let T1T_{1} such that the functions F1,F2,,FnF_{1},F_{2},...,F_{n} are differentiable in [T1,)[T_{1},\infty). Note that since the degrees of the functions 1,F1,F2,1,F_{1},F_{2},... ,Fn,F_{n} are distinct, we have that deg(Fi)0\deg(F_{i})\neq 0 for all 1iN1\leq i\leq N. In particular, deg(Fi)=deg(Fi)1\deg(F_{i}^{\prime})=\deg(F_{i})-1, and so the degrees of F1,F2,,FNF_{1}^{\prime},F_{2}^{\prime},\cdots,F_{N}^{\prime} are all distinct.

Now, if

c0+c1F1++cNFNI=0,\|c_{0}+c_{1}F_{1}+...+c_{N}F_{N}\|_{I}=0,

then

c1F1++cNFNI=0.\|c_{1}F_{1}^{\prime}+...+c_{N}F_{N}^{\prime}\|_{I}=0.

The claim now follows by induction. ∎

Note that if F0(t),F1(t),,FN(t)F_{0}(t),F_{1}(t),...,F_{N}(t) are definable functions, then they are continuously differentiable (N+1)(N+1) times for all tTt\geq T for some TT. Hence, the Wronskian matrix W(F0,,FN)(t)W(F_{0},...,F_{N})(t) which is the (N+1)×(N+1)(N+1)\times(N+1) matrix whose kk-th row is (F0(k)(t),F1(k)(t),,FN(k)(t))(F^{(k)}_{0}(t),F_{1}^{(k)}(t),\ldots,F_{N}^{(k)}(t)) for 0kN0\leq k\leq N is well defined for all tTt\geq T.

Lemma 3.6.

Suppose that Fi:[0,),i=0,1,,NF_{i}:[0,\infty)\to\mathbb{R},~{}i=0,1,...,N, are functions definable in a polynomially bounded o-minimal structure such that the degrees deg(Fi)\deg(F_{i}), for 0iN0\leq i\leq N are distinct. Then, there exists a T0>0T_{0}>0 such that the functions F0,,FNF_{0},...,F_{N} are (N+1)(N+1)-time continuously differentiable and W(F0,,FN)(t)W(F_{0},...,F_{N})(t) is non-singular for all tT0t\geq T_{0}.

Proof.

Since det(W(F0,,FN)(t))\det(W(F_{0},...,F_{N})(t)) is a definable function, then either

|det(W(F0,,FN)(t))|>0,|\det(W(F_{0},...,F_{N})(t))|>0,

for all large tt or det(W(F0,,FN)(t))=0\det(W(F_{0},...,F_{N})(t))=0 for all large tt. Suppose for contradiction the latter case, say det(W(F0,,FN)(t))=0,tT1\det(W(F_{0},...,F_{N})(t))=0,\forall t\geq T_{1}. Then, by a classical result of Bôcher [Bôc01], we get that F0,F1,,FNF_{0},F_{1},...,F_{N} are linearly dependent on any sub-interval of [T1,)[T_{1},\infty). This contradicts Lemma 3.5. ∎

Lemma 3.7.

Suppose that 𝒱\mathcal{V} is a finite dimensional vector space of functions f:[0,)f:[0,\infty)\to\mathbb{R} definable in a polynomially bounded o-minimal structure. Suppose that 𝒱\mathcal{V} does not contain non-zero functions, which are constantly zero eventually. Then there exist a basis {F1,,FN}\{F_{1},...,F_{N}\} of 𝒱\mathcal{V}, where

(3.5) deg(F1)<<deg(FN).\deg(F_{1})<\cdots<\deg(F_{N}).
Proof.

We argue by induction. The claim is trivial for N=1N=1. Now choose an arbitrary basis {H1,,HN}\{H_{1},...,H_{N}\} of 𝒱\mathcal{V}, and suppose without loss of generality that deg(Hi)deg(HN)\deg(H_{i})\leq\deg(H_{N}) for all ii. For each ii such that deg(Hi)=deg(HN)\deg(H_{i})=\deg(H_{N}), we have deg(HiciHN)<deg(HN)\deg(H_{i}-c_{i}H_{N})<\deg(H_{N}) for some ci0c_{i}\neq 0, and we replace HiH_{i} by HiciHNH_{i}-c_{i}H_{N}. Now, the modified basis {H1,,HN}\{H_{1},...,H_{N}\} is such that deg(Hi)<deg(HN),i<N\deg(H_{i})<\deg(H_{N}),\forall i<N. The vector space 𝒰:=Span{H1,,HN1}\mathcal{U}:=\text{Span}_{\mathbb{R}}\{H_{1},...,H_{N-1}\} is (N1)(N-1)-dimensional, and deg(f)<deg(HN)\deg(f)<\deg(H_{N}) for all f𝒰f\in\mathcal{U}. By induction, there is a basis {F1,,FN1}\{F_{1},...,F_{N-1}\} for 𝒰\mathcal{U} such that deg(Fi)<deg(Fi+1)\deg(F_{i})<\deg(F_{i+1}) for all ii. In particular, {F1,,FN1,FN}\{F_{1},...,F_{N-1},F_{N}\} with FN:=HNF_{N}:=H_{N} is the required basis for 𝒱\mathcal{V}. ∎

Remark 3.8.

If 𝒱\mathcal{V} is a finite dimensional vector spanned by definable functions, then the subset of functions which are eventually constantly zero forms a finite dimensional subspace 𝒱0𝒱\mathcal{V}_{0}\leq\mathcal{V}. This implies that there exists a uniform T0>0T_{0}>0 such that for all f𝒱0f\in\mathcal{V}_{0}, it holds that f(t)=0,tT0f(t)=0,\forall t\geq T_{0}. Thus, upon restricting the functions in 𝒱\mathcal{V} to [T0,)[T_{0},\infty), the subspace 𝒱0\mathcal{V}_{0} is the trivial vector space.

We note the following fact:

  • A definable function f:[0,)f:[0,\infty)\to\mathbb{R} either converges as tt\to\infty or diverges to \infty or to -\infty as t,t\to\infty, see [DM96, Montonicity theorem, Section 4].

Proof of Theorem 3.2.

Let 𝒱\mathcal{V} be a finite dimensional vector space of functions f:[0,)f:[0,\infty)\to\mathbb{R} definable in a polynomially bounded o-minimal structure. Because of Remark 3.8, there is no loss in generality in assuming that the only eventually constantly zero function is the zero function. So we may choose a basis {F1,,Fn,F0+,F1+,,Fm+}\{F_{1}^{-},...,F_{n}^{-},F_{0}^{+},F_{1}^{+},...,F_{m}^{+}\} for 𝒱\mathcal{V} such that

deg(F1)<<deg(Fn)<0deg(F0+)<<deg(Fm+)\deg(F_{1}^{-})<\cdots<\deg(F_{n}^{-})<0\leq\deg(F^{+}_{0})<\cdots<\deg(F_{m}^{+})

, where m,n{0}m,n\in\mathbb{N}\cup\{0\}. Let N=m+n+1N=m+n+1.

We will denote

ri=deg(Fi+),κj:=deg(Fj)r_{i}=\deg(F^{+}_{i}),~{}-\kappa_{j}:=\deg(F^{-}_{j})

where ri,κj0r_{i},\kappa_{j}\geq 0. For

c¯:=(c1,,cn,c0+,,cm+)N\underline{c}:=(c_{1}^{-},\ldots,c_{n}^{-},c_{0}^{+},\ldots,c^{+}_{m})\in\mathbb{R}^{N}

we denote

(3.6) fc¯:=c1F1++cnFn+c0+F0++cm+Fm+,\displaystyle f_{\underline{c}}:=c^{-}_{1}F^{-}_{1}+\cdots+c^{-}_{n}F^{-}_{n}+c^{+}_{0}F_{0}^{+}\cdots+c^{+}_{m}F^{+}_{m},
(3.7) fc¯:=c1F1++cnFn, and \displaystyle f^{-}_{\underline{c}}:=c^{-}_{1}F^{-}_{1}+...+c^{-}_{n}F^{-}_{n},\text{ and }
(3.8) fc¯+:=c0+F0+++cm+Fm+.\displaystyle f^{+}_{\underline{c}}:=c^{+}_{0}F_{0}^{+}+...+c^{+}_{m}F^{+}_{m}.

Let 𝒱\mathscr{F}\subseteq\mathcal{V} be a closed definable cone such that for all 0f0\neq f\in\mathscr{F} we have

limtf(t)0.\lim_{t\to\infty}f(t)\neq 0.
  • Let ~={c¯n+m+1:fc¯}\tilde{\mathscr{F}}=\{\underline{c}\in\mathbb{R}^{n+m+1}:f_{\underline{c}}\in\mathscr{F}\}, which is a closed cone in m+n+1\mathbb{R}^{m+n+1}.

In addition, we assume (without loss of generality) that

limtFi+(t)tri=1,limtFj(t)tκj=1.\lim_{t\to\infty}\frac{F^{+}_{i}(t)}{t^{r_{i}}}=1,~{}\lim_{t\to\infty}\frac{F^{-}_{j}(t)}{t^{-\kappa_{j}}}=1.

We choose T01T_{0}\geq 1 such that the outcome of Lemma 3.6 holds. Fix δ(0,1)\delta\in(0,1), and assume for contradiction that {0}\mathscr{F}\smallsetminus\{0\} is not δ\delta-good on [T0,)[T_{0},\infty). Consider the following definable subset:

(3.9) 𝒜:={(s,c¯,a,l,α):s1,c¯~{0},a>T0,l>0,aαa+lδl,fc¯[a,a+l]fc¯[α,α+δl]>s}.\mathcal{A}:=\left\{(s,\underline{c},a,l,\alpha):\begin{array}[]{c}s\geq 1,~{}\underline{c}\in\tilde{\mathscr{F}}\smallsetminus\{0\},~{}a>T_{0},~{}l>0,\\ a\leq\alpha\leq a+l-\delta l,~{}\frac{\|f_{\underline{c}}\|_{[a,a+l]}}{\|f_{\underline{c}}\|_{[\alpha,\alpha+\delta l]}}>s\\ \end{array}\right\}.

By the assumption for contradiction, the projection of 𝒜\mathcal{A} to the first coordinate is 1\mathbb{R}_{\geq 1}. By the choice function theorem (cf. [DM96, Section 4]), there exists a polynomially bounded definable curve

ϕ(s):=(c¯(s),a(s),l(s),α(s)) for s1,\phi(s):=(\underline{c}(s),a(s),l(s),\alpha(s))\text{ for $s\geq 1$},

such that (s,ϕ(s))𝒜,s1.(s,\phi(s))\in\mathcal{A},~{}\forall s\geq 1. In particular, there exists an C00C_{0}\neq 0 and an κ\kappa\in\mathbb{R} such that

(3.10) a(s)l(s)C0sκ, as s.\frac{a(s)}{l(s)}\sim C_{0}s^{\kappa},\text{ as $s\to\infty$}.

Let 1\lVert\cdot\rVert_{1} denote the 1\ell_{1}-norm on N\mathbb{R}^{N}, which is the sum of absolute values of the coordinates. Let

(3.11) c¯^(s):=c¯(s)c¯(s)1,s1.\hat{\underline{c}}(s):=\frac{\underline{c}(s)}{\|\underline{c}(s)\|_{1}},\,\forall s\geq 1.

We observe that by o-minimality, since c¯^(s)1=1\lVert\hat{\underline{c}}(s)\rVert_{1}=1, c¯^(s)\hat{\underline{c}}(s) converges to some vector v with v1=1\|\textbf{v}\|_{1}=1. Since c¯(s)~{0}\underline{c}(s)\in\tilde{\mathscr{F}}\setminus\{0\} and ~\tilde{\mathscr{F}} is a closed cone, we get v~{0}v\in\tilde{\mathscr{F}}\setminus\{0\}.

Case 1: κ0\kappa\leq 0. In this case, we treat two sub-cases in which l(s)l(s) is bounded or not.

Case 1.1. l(s)l(s) is bounded in ss. Then a(s)a(s) is also bounded, and by o-minimality l(s)l(s) and a(s)a(s) converge as ss\to\infty. We first observe that limsl(s)0\lim_{s\to\infty}l(s)\neq 0. In fact, if otherwise limsl(s)=0\lim_{s\to\infty}l(s)=0, then limsa(s)l(s)=\lim_{s\to\infty}\frac{a(s)}{l(s)}=\infty since a(s)T01,s1a(s)\geq T_{0}\geq 1,\forall s\geq 1, which is a contradiction to the assumption that κ0\kappa\leq 0. Also, the end-points of the intervals [α(s),α(s)+δl(s)][a(s),a(s)+l(s)][\alpha(s),\alpha(s)+\delta l(s)]\subset[a(s),a(s)+l(s)] converge to the end-points of intervals IδI[T0,)I_{\delta}\subset I\subseteq[T_{0},\infty) of positive length, where |Iδ||I|=δ\frac{|I_{\delta}|}{|I|}=\delta. In particular, fc¯^(s)[a(s),a(s)+l(s)]\|f_{\hat{\underline{c}}(s)}\|_{[a(s),a(s)+l(s)]} is uniformly bounded in ss. Now

fc¯^(s)[a(s),a(s)+l(s)]fc¯^(s)[α(s),α(s)+δl(s)]\displaystyle\frac{\|f_{\hat{\underline{c}}(s)}\|_{[a(s),a(s)+l(s)]}}{\|f_{\hat{\underline{c}}(s)}\|_{[\alpha(s),\alpha(s)+\delta l(s)]}} =fc¯(s)/c¯(s)1[a(s),a(s)+l(s)]fc¯(s)/c¯(s)1[α(s),α(s)+δl(s)]\displaystyle=\frac{\|f_{\underline{c}(s)}/\lVert\underline{c}(s)\rVert_{1}\|_{[a(s),a(s)+l(s)]}}{\|f_{\underline{c}(s)}/\lVert\underline{c}(s)\rVert_{1}\|_{[\alpha(s),\alpha(s)+\delta l(s)]}}
=fc¯(s)[a(s),a(s)+l(s)]fc¯(s)[α(s),α(s)+δl(s)]>s.\displaystyle=\frac{\|f_{\underline{c}(s)}\|_{[a(s),a(s)+l(s)]}}{\|f_{\underline{c}(s)}\|_{[\alpha(s),\alpha(s)+\delta l(s)]}}>s.

Namely, we have

fc¯^(s)[a(s),a(s)+l(s)]sfc¯^(s)[α(s),α(s)+δl(s)].\frac{\|f_{\hat{\underline{c}}(s)}\|_{[a(s),a(s)+l(s)]}}{s}\geq\|f_{\hat{\underline{c}}(s)}\|_{[\alpha(s),\alpha(s)+\delta l(s)]}.

By taking ss\to\infty we get

0=fvIδ.0=\|f_{\textbf{v}}\|_{I_{\delta}}.

This is a contradiction to Lemma 3.5.


Case 1.2. l(s)l(s) is unbounded in ss. Then, by o-minimality l(s)l(s)\to\infty.

For s1s\geq 1, we define

(3.12) C+(s)=i=0m|ci+(s)|l(s)ri,C(s)=j=1n|cj(s)|, and C(s)=C+(s)+C(s).\displaystyle C^{+}(s)=\sum_{i=0}^{m}|c^{+}_{i}(s)|l(s)^{r_{i}},\,C^{-}(s)=\sum_{j=1}^{n}|c^{-}_{j}(s)|\text{, and }C(s)=C^{+}(s)+C^{-}(s).

Then, the by properties of bounded definable functions, the following limit exists in m+1\mathbb{R}^{m+1}.

(3.13) (w0+,,wm+)=lims1C(s)(c0+(s)l(s)r0,,cm+(s)l(s)rm).(w^{+}_{0},...,w^{+}_{m})=\lim_{s\to\infty}\frac{1}{C(s)}(c^{+}_{0}(s)l(s)^{r_{0}},...,c^{+}_{m}(s)l(s)^{r_{m}}).

We denote

φc¯(s)(t):=fc¯(s)(a(s)+tl(s))λ(s),t[0,1].\varphi_{{\underline{c}}(s)}(t):=\frac{f_{\underline{c}(s)}(a(s)+tl(s))}{\lambda(s)},\,\forall t\in[0,1].

We first want to verify the following claim to be used in further arguments.

Claim 1.
(3.14) supsS0φc¯(s)[0,1]<.\sup_{s\geq S_{0}}\lVert\varphi_{{\underline{c}}(s)}\rVert_{[0,1]}<\infty.

Moreover, for any t0(0,1)t_{0}\in(0,1).

(3.15) limssupt[t0,1]|φc¯(s)(t)w0+(x0+t)r0++wm+(x0+t)rm|=0.\lim_{s\to\infty}\sup_{t\in[t_{0},1]}\lvert\varphi_{\underline{c}(s)}(t)-w^{+}_{0}\left(x_{0}+t\right)^{r_{0}}+\cdots+w^{+}_{m}\left(x_{0}+t\right)^{r_{m}}\rvert=0.
Proof of Claim 1.

For each 0im0\leq i\leq m, there exists νi>0\nu_{i}>0 such that for all t[0,1]t\in[0,1], we have

Fi+(a+tl)lri\displaystyle\frac{F^{+}_{i}(a+tl)}{l^{r_{i}}} =(a+lt)rilri+O((a+lt)riνilri)\displaystyle=\frac{\left(a+lt\right)^{r_{i}}}{l^{r_{i}}}+O\left(\frac{\left(a+lt\right)^{r_{i}-\nu_{i}}}{l^{r_{i}}}\right)
(3.16) =(al+t)ri+O(lνi(al+t)riνi).\displaystyle=\left(\frac{a}{l}+t\right)^{r_{i}}+O\left(l^{-\nu_{i}}\left(\frac{a}{l}+t\right)^{r_{i}-\nu_{i}}\right).

Since a(s)l(s)\frac{a(s)}{l(s)} is bounded, we have limsa(s)l(s)=x0[0,)\lim_{s\to\infty}\frac{a(s)}{l(s)}=x_{0}\in[0,\infty). We now make the following observations:

  • We can pick S01S_{0}\geq 1 such that l(s)1l(s)\geq 1 for all sS0s\geq S_{0} and the following two statements hold.

    1. (1)

      supt[0,1](a(s)l(s)+t)ri\sup_{t\in[0,1]}\left(\frac{a(s)}{l(s)}+t\right)^{r_{i}} is bounded for all sS0s\geq S_{0} and for all 0im0\leq i\leq m.

      This is immediate since ri0r_{i}\geq 0, limsa(s)l(s)=x0\lim_{s\to\infty}\frac{a(s)}{l(s)}=x_{0} and t[0,1]t\in[0,1].

    2. (2)

      supt[0,1]l(s)νi(a(s)l(s)+t)riνi\sup_{t\in[0,1]}l(s)^{-\nu_{i}}\left(\frac{a(s)}{l(s)}+t\right)^{r_{i}-\nu_{i}} is bounded for all sS0s\geq S_{0} and for all 0im0\leq i\leq m.

      This is clear if riνi0r_{i}-\nu_{i}\geq 0. Now suppose that riνi<0r_{i}-\nu_{i}<0. Then, for all t[0,1]t\in[0,1], since a(s)T01a(s)\geq T_{0}\geq 1 and l(s)1l(s)\geq 1,

      l(s)νi(a(s)l(s)+t)riνil(s)νi(a(s)l(s))riνi=a(s)riνil(s)ri1.\displaystyle l(s)^{-\nu_{i}}\left(\frac{a(s)}{l(s)}+t\right)^{r_{i}-\nu_{i}}\leq l(s)^{-\nu_{i}}\left(\frac{a(s)}{l(s)}\right)^{r_{i}-\nu_{i}}=\frac{a(s)^{r_{i}-\nu_{i}}}{l(s)^{r_{i}}}\leq 1.
  • Since, by definability, |a(s)/l(s)x0|=O(l(s)ϵ)\lvert a(s)/l(s)-x_{0}\rvert=O(l(s)^{-\epsilon}) for some ϵ>0\epsilon>0, for each 0im0\leq i\leq m, there exists ϵi>0\epsilon_{i}>0 such that as ss\to\infty

    supt[0,1]|(a(s)l(s)+t)ri(x0+t)ri|=O(l(s)ϵi).\sup_{t\in[0,1]}\left|\left(\frac{a(s)}{l(s)}+t\right)^{r_{i}}-\left(x_{0}+t\right)^{r_{i}}\right|=O(l(s)^{\epsilon_{i}}).
  • For any 0im0\leq i\leq m, as ss\to\infty,

    supt[0,1]l(s)νi|a(s)l(s)+t|riνi=O(l(s)νi).\sup_{t\in[0,1]}l(s)^{-\nu_{i}}\left|\frac{a(s)}{l(s)}+t\right|^{r_{i}-\nu_{i}}=O(l(s)^{-\nu_{i}}).
  • Let ν:=min{ϵ0,,ϵm,ν0,,νn}>0\nu:=\min\{\epsilon_{0},\ldots,\epsilon_{m},\nu_{0},\ldots,\nu_{n}\}>0. Then, for all 0im0\leq i\leq m,

    supt[0,1]|Fi+(a(s)+tl(s))l(s)ri(x0+t)ri|=O(l(s)ν),\sup_{t\in[0,1]}\left|\frac{F^{+}_{i}(a(s)+tl(s))}{l(s)^{r_{i}}}-(x_{0}+t)^{r_{i}}\right|=O(l(s)^{-\nu}),

Thus, as ss\to\infty, uniformly for t[0,1]t\in[0,1],

(3.17) fc¯(s)+(a(s)+tl(s))=i=0m(ci+(s)l(s)ri)(a(s)l(s)+t)ri+O(C+(s)l(s)ν).\displaystyle f^{+}_{\underline{c}(s)}(a(s)+tl(s))=\sum_{i=0}^{m}(c_{i}^{+}(s)l(s)^{r_{i}})\left(\frac{a(s)}{l(s)}+t\right)^{r_{i}}+O(C^{+}(s)l(s)^{-\nu}).

We also observe that as ss\to\infty, we have

(3.18) supt[0,1]|fc¯(s)(a(s)+tl(s))|=O(C(s)),\sup_{t\in[0,1]}\left|f_{\underline{c}(s)}^{-}(a(s)+tl(s))\right|=O(C^{-}(s)),

and for any t0(0,1)t_{0}\in(0,1),

(3.19) supt[t0,1]|fc¯(s)(a(s)+tl(s))|=O(C(s)l(s)κn).\sup_{t\in[t_{0},1]}\left|f_{\underline{c}(s)}^{-}(a(s)+tl(s))\right|=O(C^{-}(s)l(s)^{-\kappa_{n}}).

Now (3.14) follows from (3.17) and (3.18), and (3.15) follows from (3.17) and (3.19). This completes the proof of the Claim 1.

By (3.9) and the definition of the choice function ϕ\phi,

(3.20) s<fc¯(s)[a(s),a(s)+l(s)]fc¯(s)[α(s),α(s)+δl(s)]=φc¯(s)[0,1]φc¯(s)Iδ(s)s<\frac{\|f_{\underline{c}(s)}\|_{[a(s),a(s)+l(s)]}}{\|f_{\underline{c}(s)}\|_{[\alpha(s),\alpha(s)+\delta l(s)]}}=\frac{\|\varphi_{{\underline{c}}(s)}\|_{[0,1]}}{\|\varphi_{{\underline{c}}(s)}\|_{I_{\delta}(s)}}

where Iδ(s):=[α(s)a(s)l(s),α(s)a(s)+δl(s)l(s)][0,1]I_{\delta}(s):=\left[\frac{\alpha(s)-a(s)}{l(s)},\frac{\alpha(s)-a(s)+\delta l(s)}{l(s)}\right]\subseteq[0,1] having length |Iδ(s)|=δ|I_{\delta}(s)|=\delta for all s1.s\geq 1. Therefore, by (3.14),

(3.21) limsφc¯(s)Iδ(s)limsφc¯(s)[0,1]s=0.\lim_{s\to\infty}\|\varphi_{{\underline{c}}(s)}\|_{I_{\delta}(s)}\leq\lim_{s\to\infty}\frac{\|\varphi_{{\underline{c}}(s)}\|_{[0,1]}}{s}=0.

And, since the end-points of Iδ(s)I_{\delta}(s) are bounded, they converge to endpoints of a sub-interval IδI_{\delta} of [0,1][0,1] of length δ\delta. Therefore, from (3.15) and (3.21), we deduce that

w0+(x0+t)r0++wm+(x0+t)rm=0,w^{+}_{0}\left(x_{0}+t\right)^{r_{0}}+\cdots+w^{+}_{m}\left(x_{0}+t\right)^{r_{m}}=0,

for all tIδ[t0,1]t\in I_{\delta}\cap[t_{0},1] for any t0(0,1)t_{0}\in(0,1); and hence for all tIδt\in I_{\delta}. Since IδI_{\delta} is an interval of length δ\delta, by Lemma 3.5, we obtain that w0+==wm+=0w^{+}_{0}=\cdots=w^{+}_{m}=0. Therefore, by (3.13), we get

lims(c1(s),,cn(s),c0+(s)l(s)r0,,cm+(s)l(s)rm)(c1(s),,cn(s),c0+(s)l(s)r0,,cm+(s)l(s)rm)1\displaystyle\lim_{s\to\infty}\frac{(c_{1}^{-}(s),...,c_{n}^{-}(s),c^{+}_{0}(s)l(s)^{r_{0}},...,c^{+}_{m}(s)l(s)^{r_{m}})}{\|(c^{-}_{1}(s),...,c^{-}_{n}(s),c^{+}_{0}(s)l(s)^{r_{0}},...,c_{m}^{+}(s)l(s)^{r_{m}})\|_{1}}
=(w1,,wn,w0+,,wm+)=(w1,,wn,0,,0).\displaystyle=(w^{-}_{1},...,w^{-}_{n},w^{+}_{0},...,w^{+}_{m})=(w^{-}_{1},...,w^{-}_{n},0,...,0).

Therefore, n1n\geq 1 and

(3.22) (w1,,wn)1=1.\|(w^{-}_{1},...,w^{-}_{n})\|_{1}=1.

In view of (3.11), we have

v=(v1,,vn,v0+,,vm+)\displaystyle\textbf{v}=(v^{-}_{1},...,v^{-}_{n},v^{+}_{0},...,v^{+}_{m}) :=lims^c¯(s)\displaystyle:=\lim_{s\to\infty}\hat{}\underline{c}(s)
=lims(c1(s),,cn(s),c0+(s),,cm+(s))(c1(s),,cn(s),c0+(s),,cm+(s))1.\displaystyle=\lim_{s\to\infty}\frac{(c_{1}^{-}(s),...,c_{n}^{-}(s),c^{+}_{0}(s),...,c^{+}_{m}(s))}{\|(c^{-}_{1}(s),...,c^{-}_{n}(s),c^{+}_{0}(s),...,c^{+}_{m}(s))\|_{1}}.

Then, v1=1\|\textbf{v}\|_{1}=1 and v~{0}\textbf{v}\in\tilde{\mathscr{F}}\smallsetminus\{0\}.

By (3.22), we have

1\displaystyle 1 =(w1,,wn)1\displaystyle=\|(w_{1}^{-},...,w_{n}^{-})\|_{1}
=lims(c1(s),,cn(s))1(c1(s),,cn(s),c0+(s)l(s)r0,,cm+(s)l(s)rm)1\displaystyle=\lim_{s\to\infty}\frac{\|(c_{1}^{-}(s),...,c_{n}^{-}(s))\|_{1}}{\|(c^{-}_{1}(s),...,c^{-}_{n}(s),c^{+}_{0}(s)l(s)^{r_{0}},...,c^{+}_{m}(s)l(s)^{r_{m}})\|_{1}}
lims(c1(s),,cn(s))1(c1(s),,cn(s),c0+(s),,cm+(s))1, as l(s) and ri0\displaystyle\leq\lim_{s\to\infty}\frac{\|(c_{1}^{-}(s),...,c_{n}^{-}(s))\|_{1}}{\|(c^{-}_{1}(s),...,c^{-}_{n}(s),c^{+}_{0}(s),...,c^{+}_{m}(s))\|_{1}}\text{, as $l(s)\to\infty$ and $r_{i}\geq 0$}
=(v1,,vn)11.\displaystyle=\|(v_{1}^{-},...,v_{n}^{-})\|_{1}\leq 1.

Therefore, (v1,,vn)1\|(v_{1}^{-},...,v_{n}^{-})\|_{1}. As a consequence, we get v0+==vm+=0v^{+}_{0}=\cdots=v^{+}_{m}=0. Therefore, fv(t)0f_{\textbf{v}}(t)\to 0 as tt\to\infty, where v~{0}\textbf{v}\in\tilde{\mathscr{F}}\smallsetminus\{0\}, and hence fv{0}f_{\textbf{v}}\in\mathscr{F}\smallsetminus\{0\}. This contradicts our assumption that no nonzero function in \mathscr{F} decays to zero.

Remark 3.9.

This is the only place in the proof of Theorem 1.14 where we use the non-decaying function assumption for the closed cone \mathscr{F}. And this is the main reason we need the non-contraction assumption on the curves in this article.

Case 2: κ>0\kappa>0.

Let λ(s)=c¯(s)1\lambda(s)=\|\underline{c}(s)\|_{1} and denote: c¯^:=c¯λ\hat{\underline{c}}:=\frac{\underline{c}}{\lambda}. As above, note that

fc¯^(s)[a(s),a(s)+l(s)]fc¯^(s)[α(s),α(s)+δl(s)]=fc¯(s)[a(s),a(s)+l(s)]fc¯(s)[α(s),α(s)+δl(s)]>s.\frac{\|f_{\hat{\underline{c}}(s)}\|_{[a(s),a(s)+l(s)]}}{\|f_{\hat{\underline{c}}(s)}\|_{[\alpha(s),\alpha(s)+\delta l(s)]}}=\frac{\|f_{\underline{c}(s)}\|_{[a(s),a(s)+l(s)]}}{\|f_{\underline{c}(s)}\|_{[\alpha(s),\alpha(s)+\delta l(s)]}}>s.

We now consider two cases: a(s)a(s) is bounded or a(s)a(s)\to\infty.


Case 2.1. a(s)a(s) is bounded. We observe that limsl(s)=0\lim_{s\to\infty}l(s)=0. This follows since a(s)T01a(s)\geq T_{0}\geq 1 and κ>0\kappa>0, we get a(s)l(s)C0sκ\frac{a(s)}{l(s)}\sim C_{0}s^{\kappa}\to\infty. We have that limsa(s)=x0T0\lim_{s\to\infty}a(s)=x_{0}\geq T_{0}, and limsc¯^(s)=v\lim_{s\to\infty}\hat{\underline{c}}(s)=\textbf{v} with v1=1.\|\textbf{v}\|_{1}=1.

By our choice of T0T_{0}, the basis functions Fi,Fj+F^{-}_{i},F^{+}_{j} are (N+1)(N+1)-times continuously differentiable where N:=(n+m+1)N:=(n+m+1). For any tT0t\geq T_{0}, let W(t)W(t) denote the transpose of the Wronskian matrix W(Fn(t),,Fm(t))W(F_{-n}(t),\ldots,F_{m}(t)), whose kk-th row is (kFn(t),,kFm(t))(\partial^{k}F_{-n}(t),\ldots,\partial^{k}F_{m}(t)) for k=0,,Nk=0,\dots,N. For hh\in\mathbb{R}, define

(3.23) QN,a(s)(h)=c¯^(s)W(a(s))(h0,,hN),Q_{N,a(s)}(h)=\hat{\underline{c}}(s)W(a(s))\cdot(h^{0},\ldots,h^{N}),

where \cdot denotes the dot product on N+1\mathbb{R}^{N+1}. By Taylor’s theorem (mean value form for the remainder), for all h[0,l(s)]h\in[0,l(s)], there exists ξa(s)+[0,l(s)]:=[a(s),a(s)+l(s)]\xi\in a(s)+[0,l(s)]:=[a(s),a(s)+l(s)] such that

fc¯^(s)(a(s)+h)QN,a(s)(h)=fc¯^(s)(N+1)(ξ)(n+1)!hN+1.\displaystyle f_{\hat{\underline{c}}(s)}(a(s)+h)-Q_{N,a(s)}(h)=\frac{f_{\hat{\underline{c}}(s)}^{(N+1)}(\xi)}{(n+1)!}h^{N+1}.

Since a(s)a(s), c¯^(s)\hat{\underline{c}}(s), and l(s)l(s) are bounded, we get that |fc¯^(s)(N+1)(ξ)||f_{\hat{\underline{c}}(s)}^{(N+1)}(\xi)| is uniformly bounded in the range ξ[a(s),a(s)+l(s)]\xi\in[a(s),a(s)+l(s)].

Denote I(s):=[0,l(s)]I(s):=[0,l(s)], and Iδ(s):=[0,δl(s)]I_{\delta}(s):=[0,\delta l(s)]. We have:

(By Taylor approximation) fc¯^(s)a(s)+I(s)=\displaystyle\|f_{\hat{\underline{c}}(s)}\|_{a(s)+I(s)}= QN,a(s)I(s)+O(l(s)N+1)\displaystyle\|Q_{N,a(s)}\|_{I(s)}+O(l(s)^{N+1})
(By Proposition 3.4) \displaystyle\leq MN(δ)QN,a(s)Iδ(s)+O(l(s)N+1)\displaystyle M_{N}(\delta)\|Q_{N,a(s)}\|_{I_{\delta}(s)}+O(l(s)^{N+1})
(By Taylor approximation) \displaystyle\leq MN(δ)fc¯^(s)a(s)+Iδ(s)+O(l(s)N+1)\displaystyle M_{N}(\delta)\|f_{\hat{\underline{c}}(s)}\|_{a(s)+I_{\delta}(s)}+O(l(s)^{N+1})
=\displaystyle= MN(δ)fc¯^(s)a(s)+Iδ(s)[1+O(l(s)N+1fc¯^(s)a(s)+Iδ(s))].\displaystyle M_{N}(\delta)\|f_{\hat{\underline{c}}(s)}\|_{a(s)+I_{\delta}(s)}\left[1+O\left(\frac{l(s)^{N+1}}{\|f_{\hat{\underline{c}}(s)}\|_{a(s)+I_{\delta}(s)}}\right)\right].

Thus,

fc¯(s)a(s)+I(s)fc¯(s)a(s)+Iδ(s)1+O(l(s)N+1fc¯^(s)a(s)+Iδ(s)).\frac{\|f_{{\underline{c}}(s)}\|_{a(s)+I(s)}}{\|f_{{\underline{c}}(s)}\|_{a(s)+I_{\delta}(s)}}\ll 1+O\left(\frac{l(s)^{N+1}}{\|f_{\hat{\underline{c}}(s)}\|_{a(s)+I_{\delta}(s)}}\right).

Since fc¯^(s)a(s)+Iδ(s)=QN,a(s)Iδ(s)+O(l(s)N+1)\|f_{\hat{\underline{c}}(s)}\|_{a(s)+I_{\delta}(s)}=\|Q_{N,a(s)}\|_{I_{\delta}(s)}+O(l(s)^{N+1}), we will show that

(3.24) limsl(s)N+1QN,a(s)Iδ(s)=0.\lim_{s\to\infty}\frac{l(s)^{N+1}}{\|Q_{N,a(s)}\|_{I_{\delta}(s)}}=0.

This outcome gives a contradiction to our assumption that sfc¯(s)I(s)fc¯(s)Iδ(s)s\leq\frac{\|f_{{\underline{c}}(s)}\|_{I(s)}}{\|f_{{\underline{c}}(s)}\|_{I_{\delta}(s)}} for all s1s\geq 1.

To prove (3.24), we observe the following if q(x)=c0+c1x++ckxkq(x)=c_{0}+c_{1}x+...+c_{k}x^{k} is a polynomial of degree bounded by kk and J[0,1]J\subset[0,1] is an interval of length |J|<1|J|<1, then by Proposition 3.4,

(3.25) qJ(k+1)1kkq[0,1]|J|k(c0,c1,,ck)1|J|k,\|q\|_{J}\geq(k+1)^{-1}k^{-k}\|q\|_{[0,1]}|J|^{k}\geq\|(c_{0},c_{1},...,c_{k})\|_{1}|J|^{k},

where the implied constant depends only on kk, because q[0,1]\|q\|_{[0,1]} and c¯1\|\underline{c}\|_{1} are two norms on the (k+1)(k+1)-dimensional space of polynomials of degree at most kk.

By (3.23), QN,a(s)Q_{N,a(s)} is a polynomial of degree bounded by NN whose coefficients are given by W(a(s))c¯^(s)W(a(s))\hat{\underline{c}}(s). We have W(a(s))W(x0)W(a(s))\to W(x_{0}) as ss\to\infty. And, by Lemma 3.6, W(x0)W(x_{0}) is nonsingular. Therefore,

W(a(s))c¯^(s)1W(a(s))111c¯^(s)1(1/2)W(x0)111>0\lVert W(a(s))\hat{\underline{c}}(s)\rVert_{1}\geq\lVert W(a(s))^{-1}\rVert_{1}^{-1}\lVert\hat{\underline{c}}(s)\rVert_{1}\geq(1/2)\lVert W(x_{0})^{-1}\rVert_{1}^{-1}>0

for all s1s\gg 1, as c¯^(s)1=1\lVert\hat{\underline{c}}(s)\rVert_{1}=1. Then, by (3.25) applied to J=[0,δl(s)]J=[0,\delta l(s)], we get

(3.26) QN,a(s)Iδ(s)W(x0)111δNl(s)N.\|Q_{N,a(s)}\|_{I_{\delta}(s)}\gg\lVert W(x_{0})^{-1}\rVert_{1}^{-1}\delta^{N}l(s)^{N}.

Since δ\delta is fixed, and l(s)0l(s)\to 0, we get (3.24).


Case 2.2. a(s)a(s) is unbounded. Here a(s)sθa(s)\asymp s^{\theta}, for θ>0\theta>0. We recall that for each n\mathrm{n}\in\mathbb{N} there is TnT_{\mathrm{n}} such that all basis functions Fi,Fj+F^{-}_{i},F_{j}^{+} will be continuously differentiable n\mathrm{n} times in the ray [Tn,)[T_{\mathrm{n}},\infty) for all i,ji,j, see [DM96]. By [Mil94b], it holds that

|dpdtpFi(x)|xκip, and |dqdtqFj+(x)|xrjq,\left|\frac{d^{p}}{dt^{p}}F^{-}_{i}(x)\right|\ll x^{-\kappa_{i}-p},\text{ and }\left|\frac{d^{q}}{dt^{q}}F^{+}_{j}(x)\right|\ll x^{r_{j}-q},

where p,qp,q are non-negative integers.

Let J(s)=[a(s),a(s)+l(s)]J(s)=[a(s),a(s)+l(s)]. Then we conclude that for an integer n\mathrm{n} such that ri(n+1)<0r_{i}-(\mathrm{n}+1)<0 for all 0im0\leq i\leq m, we have

suptJ(s)|dn+1dtn+1fc¯^(s)(t)|=O(a(s)rm(n+1)).\sup_{t\in J(s)}\left\lvert\frac{d^{\mathrm{n}+1}}{dt^{\mathrm{n}+1}}f_{\hat{\underline{c}}(s)}(t)\right\rvert=O(a(s)^{r_{m}-(\mathrm{n}+1)}).

By Taylor’s theorem, for the Taylor polynomial Pn,a(s)(t)P_{\mathrm{n},a(s)}(t) of fc¯^(s)(t)f_{\hat{\underline{c}}(s)}(t) of degree n\mathrm{n} centered around a(s)a(s), we have

suptJ(s)|fc¯^(s)(t)Pn,a(s)(t)|\displaystyle\sup_{t\in J(s)}|f_{\hat{\underline{c}}(s)}(t)-P_{\mathrm{n},a(s)}(t)| =O(a(s)rm(n+1)l(s)n+1)\displaystyle=O(a(s)^{r_{m}-(\mathrm{n}+1)}l(s)^{\mathrm{n}+1})
=O(a(s)rm(l(s)a(s))n+1)\displaystyle=O\left(a(s)^{r_{m}}\left(\frac{l(s)}{a(s)}\right)^{\mathrm{n}+1}\right)
(3.27) =O(srmθs(n+1)κ),\displaystyle=O\left(s^{r_{m}\theta}s^{-(\mathrm{n}+1)\kappa}\right),

where we use the fact that a(s)/l(s)C0sκa(s)/l(s)\sim C_{0}s^{\kappa} for some C00C_{0}\neq 0 and κ>0\kappa>0. Due to Lemma 3.5, the function Ψ(s):=fc¯^(s)[α(s),α(s)+δl(s)]\Psi(s):=\|f_{\hat{\underline{c}}(s)}\|_{[\alpha(s),\alpha(s)+\delta l(s)]} is positive. Since Ψ\Psi is also definable in a polynomially bounded o-minimal structure, we pick η>0\eta>0 such that

(3.28) fc¯^(s)J(s)sη,as s.\|f_{\hat{\underline{c}}(s)}\|_{J(s)}\gg s^{-\eta},~{}\text{as }s\to\infty.

We take n\mathrm{n} large enough such that

ν:=rmθ(n+1)κ<η.\nu:=r_{m}\theta-(\mathrm{n}+1)\kappa<-\eta.

Then, for Jδ(s)=[α(s),α(s)+δl(s)]J_{\delta}(s)=[\alpha(s),\alpha(s)+\delta l(s)],

(by (3.27)) fc¯^(s)J(s)\displaystyle\|f_{\hat{\underline{c}}(s)}\|_{J(s)}\leq Pn,a(s)J(s)+O(sν)\displaystyle\|P_{\mathrm{n},a(s)}\|_{J(s)}+O(s^{\nu})
(by Proposition 3.4) \displaystyle\leq Mn(δ)Pn,a(s)Jδ(s)+O(sν)\displaystyle M_{\mathrm{n}}(\delta)\|P_{\mathrm{n},a(s)}\|_{J_{\delta}(s)}+O(s^{\nu})
(by (3.27)) \displaystyle\leq Mn(δ)fc¯^(s)Jδ(s)+O(sν)\displaystyle M_{\mathrm{n}}(\delta)\|f_{\hat{\underline{c}}(s)}\|_{J_{\delta}(s)}+O(s^{\nu})
=\displaystyle= Mn(δ)(fc¯^(s)Jδ(s)[1+O(sνfc¯^(s)Jδ(s))]\displaystyle M_{\mathrm{n}}(\delta)(\|f_{\hat{\underline{c}}(s)}\|_{J_{\delta}(s)}\left[1+O\left(\frac{s^{\nu}}{\|f_{\hat{\underline{c}}(s)}\|_{J_{\delta}(s)}}\right)\right]
(by (3.28)) \displaystyle\ll Mn(δ)fc¯^(s)Jδ(s)[1+O(sνsη)]\displaystyle M_{\mathrm{n}}(\delta)\|f_{\hat{\underline{c}}(s)}\|_{J_{\delta}(s)}\left[1+O\left(\frac{s^{\nu}}{s^{-\eta}}\right)\right]

Namely, fc¯^(s)J(s)fc¯^(s)Jδ(s)\frac{\|f_{\hat{\underline{c}}(s)}\|_{J(s)}}{\|f_{\hat{\underline{c}}(s)}\|_{J_{\delta}(s)}} is bounded for all large ss, contradicting assumption s<fc¯^(s)J(s)fc¯^(s)Jδ(s)s<\frac{\|f_{\hat{\underline{c}}(s)}\|_{J(s)}}{\|f_{\hat{\underline{c}}(s)}\|_{J_{\delta}(s)}}. ∎

3.3. Relative time near varieties

Using the (C,α)(C,\alpha)-good property, we prove the following proposition. It is an analog of [DM93, Proposition 4.2], and it is key for the linearization technique.

Proposition 3.10.

Let ψ:[0,)GL(m,)\psi:[0,\infty)\to\text{GL}(m,\mathbb{R}) be a continuous, unbounded curve definable in a polynomially bounded o-minimal structure such that limtψ(t)v0,vm{0}\lim_{t\to\infty}\psi(t)v\neq 0,\forall v\in\mathbb{R}^{m}\smallsetminus\{0\}. Consider a non-zero polynomial Q:mQ:\mathbb{R}^{m}\to\mathbb{R}, and let

V:={xm:Q(x)=0}.\emph{\text{V}}:=\{x\in\mathbb{R}^{m}:Q(x)=0\}.

Then, there exists T0>0T_{0}>0 such that the following holds: for an ϵ>0\epsilon>0 and a compact subset K1V\emph{\text{K}}_{1}\subset\emph{\text{V}}, there exists a compact subset K2\emph{\text{K}}_{2} with

K1K2V\emph{\text{K}}_{1}\subset\emph{\text{K}}_{2}\subset\emph{\text{V}}

such that for every compact neighborhood Φ\Phi of K2K_{2} in m\mathbb{R}^{m}, there exists a compact neighborhood Ψ\Psi of K1K_{1} in m\mathbb{R}^{m}, with ΨΦ̊\Psi\subset\mathring{\Phi}, where Φ̊\mathring{\Phi} denotes the interior of Φ\Phi, such that:

(3.29) |{t[a,b]:ψ(t)vΨ}|ϵ(ba).|\{t\in[a,b]:\psi(t)v\in\Psi\}|\leq\epsilon(b-a).

for all vmv\in\mathbb{R}^{m} and [a,b][T0,)[a,b]\subseteq[T_{0},\infty), for which:

  • ψ(t)vΦ,t[a,b]\psi(t)v\in\Phi,~{}\forall t\in[a,b], and

  • ψ(b)vΦΦ̊\psi(b)v\in\Phi\smallsetminus\mathring{\Phi}.

Remark 3.11.

Compared to [DM93, Proposition 4.2], the statement of the above proposition is different in the assumption that the curve ψ(t)v\psi(t)v remains in Φ¯\overline{\Phi} throughout the whole interval [a,b][a,b], and exists Φ\Phi at t=bt=b. This is a stronger requirement than the one in [DM93, Proposition 4.2] which only asks for ψ(t0)vΦ\psi(t_{0})v\notin\Phi for some t0[a,b]t_{0}\in[a,b].

Remark 3.12.

Notice that for any affine variety Vm\text{V}\subset\mathbb{R}^{m} there exists a polynomial Q:mQ:\mathbb{R}^{m}\to\mathbb{R} such that V is the zero set of QQ. Indeed, by the Hilbert basis theorem, V is the zero set of finitely many polynomials f1,,frf_{1},...,f_{r}, and we put Q:=f12++fr2Q:=f_{1}^{2}+\cdots+f_{r}^{2}.

Our proof of Proposition 3.10 builds on the following versions of the modified (C,α)(C,\alpha)-good property.

Proposition 3.13.

Let ψ:[0,)GL(m,)\psi:[0,\infty)\to\text{GL}(m,\mathbb{R}) be a continuous curve definable in a polynomially bounded o-minimal structure, and let Q:mQ:\mathbb{R}^{m}\to\mathbb{R} be a polynomial. Then there exist constants ν,C,α,T0>0\nu,C,\alpha,T_{0}>0 such that for any vmv\in\mathbb{R}^{m} it holds that either:

  1. (1)

    Q(ψ(t)v)=0,tT0Q(\psi(t)v)=0,\forall t\geq T_{0}, or

  2. (2)

    Θv(t):=tνQ(ψ(t)v)\Theta_{v}(t):=t^{\nu}Q(\psi(t)v) is (C,α)(C,\alpha)-good on [T0,)[T_{0},\infty).

Proof.

Consider the vector space 𝒱\mathcal{V}, of functions from [0,)[0,\infty) to \mathbb{R}, defined by:

𝒱=Span{tQ(ψ(t)v):vm}.\mathcal{V}=\text{Span}_{\mathbb{R}}\{t\mapsto Q(\psi(t)v):v\in\mathbb{R}^{m}\}.

It is straightforward to verify that 𝒱\mathcal{V} is finite dimensional, and that all functions in 𝒱\mathcal{V} are definable. In view of Remark 3.8, there is a T00T_{0}\geq 0 such that if f𝒱f\in\mathcal{V} is eventually zero, then f(t)=0,t[T0,)f(t)=0,\forall t\in[T_{0},\infty). We consider \mathscr{F} to be the space of functions of 𝒱\mathcal{V} restricted to [T0,).[T_{0},\infty). By Lemma 3.7 that there is a basis {f0,f1,,fn}\{f_{0},f_{1},...,f_{n}\} for \mathscr{F} such that deg(f0)<deg(f1)<<deg(fn)\deg(f_{0})<\deg(f_{1})<...<\deg(f_{n}). Let ν>0\nu>0 such that deg(tνf0)0\deg(t^{\nu}f_{0})\geq 0. Then, by Theorem 1.14, we pick T1>0T_{1}>0 such that any non-zero function

ftν=Span{ttνQ(ψ(t)v):vm}f\in t^{\nu}\otimes\mathscr{F}=\text{Span}_{\mathbb{R}}\{t\mapsto t^{\nu}Q(\psi(t)v):v\in\mathbb{R}^{m}\}

is (C,α)(C,\alpha)-good in [T1,)[T_{1},\infty). ∎

We now note the following corollary.

Corollary 3.14.

Let ψ:[0,)GL(m,)\psi:[0,\infty)\to\text{GL}(m,\mathbb{R}) be a continuous curve, definable in a polynomially bounded o-minimal structure. Fix a polynomial Q:mQ:\mathbb{R}^{m}\to\mathbb{R}. Then there exist T0,C,α>0T_{0},C,\alpha>0 such that for any vmv\in\mathbb{R}^{m} and any interval [x,y][T0,)[x,y]\subseteq[T_{0},\infty), if supt[x,y]|Q(ψ(t)v)|=|Q(ψ(y)v)|\sup_{t\in[x,y]}|Q(\psi(t)v)|=|Q(\psi(y)v)|, then

(3.30) |{t[x,y]:|Q(ψ(t)v)|η}|C(η|Q(ψ(y)v)|)α(yx),\left|\{t\in[x,y]:|Q(\psi(t)v)|\leq\eta\}\right|\leq C\left(\frac{\eta}{|Q(\psi(y)v)|}\right)^{\alpha}(y-x),

for all η>0\eta>0; we may say that t|Q(ψ(t)v)|t\mapsto\lvert Q(\psi(t)v)\rvert is right-max-(C,α)(C,\alpha)-good in [T0,)[T_{0},\infty).

Proof.

Let ν,C,α,T0>0\nu,C,\alpha,T_{0}>0 be as in Proposition 3.13. Let [x,y][T0,)[x,y]\subseteq[T_{0},\infty) and consider a vector vmv\in\mathbb{R}^{m} with |Q(ψ(y)v)|=δ>0|Q(\psi(y)v)|=\delta>0 and |Q(ψ(t)v)|δ|Q(\psi(t)v)|\leq\delta for all t[x,y]t\in[x,y]. Then it follows that

supt[x,y]|tνQ(ψ(t)v)|=yνδ.\sup_{t\in[x,y]}|t^{\nu}Q(\psi(t)v)|=y^{\nu}\delta.

Now, by the (C,α)(C,\alpha)-good property of |tνQ(ψ(t)v)||t^{\nu}Q(\psi(t)v)|, we get:

|{t[x,y]:|Q(ψ(t)v)|η}|\displaystyle\left|\{t\in[x,y]:|Q(\psi(t)v)|\leq\eta\}\right| =|{t[x,y]:|tνQ(ψ(t)v)|tνη}|\displaystyle=\left|\{t\in[x,y]:|t^{\nu}Q(\psi(t)v)|\leq t^{\nu}\eta\}\right|
|{t[x,y]:|tνQ(ψ(t)v)|yνη}|\displaystyle\leq\left|\{t\in[x,y]:|t^{\nu}Q(\psi(t)v)|\leq y^{\nu}\eta\}\right|
(3.31) C(yνηyνδ)α(yx)=C(ηδ)α(yx).\displaystyle\leq C\left(\frac{y^{\nu}\eta}{y^{\nu}\delta}\right)^{\alpha}(y-x)=C\left(\frac{\eta}{\delta}\right)^{\alpha}(y-x).

Proof for Proposition 3.10.

Let ψ:[0,)GL(m,)\psi:[0,\infty)\to\text{GL}(m,\mathbb{R}) be a continuous unbounded curve definable in a polynomially bounded o-minimal structure such that limtψ(t)v0,v0\lim_{t\to\infty}\psi(t)v\neq 0,\forall v\neq 0. Pick T0,C,α>0T_{0},C,\alpha>0 such that for all vmv\in\mathbb{R}^{m}, the (C,α)(C,\alpha)-good property holds for the norm-map tψ(t)vt\mapsto\lVert\psi(t)v\rVert as in Proposition 1.15, and the right-max-(C,α)(C,\alpha)-property holds for the map t|Q(ψ(t)v)|t\mapsto\lvert Q(\psi(t)v)\rvert as in Corollary 3.14. Let ϵ>0\epsilon>0 be given.

Given a compact set K1V\text{K}_{1}\subset\text{V}, let R1>0R_{1}>0 be such that K1BR1(0)\text{K}_{1}\subset B_{R_{1}}(0). Let K2:=BR2(0)¯V\text{K}_{2}:=\overline{\text{B}_{R_{2}}(0)}\cap\text{V}, where R2>R1R_{2}>R_{1} is such that C(R1/R2)αϵC(R_{1}/R_{2})^{\alpha}\leq\epsilon, see (3.34).

For any ϵ>0\epsilon^{\prime}>0, define

Vϵ:={xm:|Q(x)|<ϵ}.\text{V}^{\epsilon^{\prime}}:=\{x\in\mathbb{R}^{m}:|Q(x)|<\epsilon^{\prime}\}.

Let Φ\Phi be a compact neighborhood of K2\text{K}_{2}. Then, we can pick an ϵ1>0\epsilon_{1}>0 such that

(3.32) BR2(0)Vϵ1Φ̊.B_{R_{2}}(0)\cap\text{V}^{\epsilon_{1}}\subset\mathring{\Phi}.

We pick ϵ2(0,ϵ1)\epsilon_{2}\in(0,\epsilon_{1}) such that K1BR1(0)Vϵ2K_{1}\subset B_{R_{1}}(0)\cap\text{V}^{\epsilon_{2}} and C(ϵ2/ϵ1)αϵC(\epsilon_{2}/\epsilon_{1})^{\alpha}\leq\epsilon, see (3.35). Let

(3.33) Ψ:=BR1(0)Vϵ2¯.\Psi:=\overline{B_{R_{1}}(0)\cap\text{V}^{\epsilon_{2}}}.

Let vmv\in\mathbb{R}^{m} and an interval [a,b][T0,)[a,b]\subset[T_{0},\infty) such that:

  • ψ(t)vΦ,t[a,b]\psi(t)v\in\Phi,~{}\forall t\in[a,b], and

  • ψ(b)vΦΦ̊\psi(b)v\in\Phi\smallsetminus\mathring{\Phi}.

Then by (3.32), ψ(b)vR2\|\psi(b)v\|\geq R_{2} or |Q(ψ(b)v)|ϵ1.|Q(\psi(b)v)|\geq\epsilon_{1}. Now, if ψ(b)vR2\|\psi(b)v\|\geq R_{2}, then by (3.33) and Proposition 1.15,

|{t[a,b]|ψ(t)vΨ}|\displaystyle|\{t\in[a,b]|\psi(t)v\in\Psi\}|\leq |{t[a,b]|ψ(t)vR1}|\displaystyle|\{t\in[a,b]|\|\psi(t)v\|\leq R_{1}\}|
(3.34) \displaystyle\leq C(R1R2)α(ba)ϵ(ba).\displaystyle C\left(\frac{R_{1}}{R_{2}}\right)^{\alpha}(b-a)\leq\epsilon(b-a).

Suppose now that |Q(ψ(b)v)|ϵ1|Q(\psi(b)v)|\geq\epsilon_{1}. Since t|Q(ψ(t)v)|t\mapsto|Q(\psi(t)v)| is continuous and definable, and since a definable map takes a given value only finitely many times, there is a decomposition

{t[a,b]:|Q(ψ(t)v)|ϵ1}=[a1,b1][ak,bk],\{t\in[a,b]:|Q(\psi(t)v)|\leq\epsilon_{1}\}=[a_{1},b_{1}]\sqcup\cdots\sqcup[a_{k},b_{k}],

for some kk\in\mathbb{N}, where for all 1ik1\leq i\leq k it holds

|Q(ψ(bi)v)|=ϵ1 and |Q(ψ(t)v)|ϵ1,t[ai,bi].|Q(\psi(b_{i})v)|=\epsilon_{1}\text{ and }|Q(\psi(t)v)|\leq\epsilon_{1},\forall t\in[a_{i},b_{i}].

Now, since Ψ=BR1(0)Vϵ2\Psi=B_{R_{1}}(0)\cap\text{V}^{\epsilon_{2}}, we have

{t[a,b]|ψ(t)vΨ}{t[a,b]:|Q(ψ(t)v)|ϵ2},\displaystyle\{t\in[a,b]|\psi(t)v\in\Psi\}\subseteq\{t\in[a,b]:|Q(\psi(t)v)|\leq\epsilon_{2}\},

and since ϵ2<ϵ1\epsilon_{2}<\epsilon_{1}, we have

{t[a,b]:|Q(ψ(t)v)|ϵ2}=i=1k{t[ai,bi]:|Q(ψ(t)v)|ϵ2}.\displaystyle\{t\in[a,b]:|Q(\psi(t)v)|\leq\epsilon_{2}\}=\bigsqcup_{i=1}^{k}\{t\in[a_{i},b_{i}]:|Q(\psi(t)v)|\leq\epsilon_{2}\}.

By (3.30) of Corollary 3.14,

(3.35) |{t[ai,bi]:|Q(ψ(t)v)|ϵ2}|C(ϵ2ϵ1)α(biai)ϵ(biai),\displaystyle|\{t\in[a_{i},b_{i}]:|Q(\psi(t)v)|\leq\epsilon_{2}\}|\leq C\left(\frac{\epsilon_{2}}{\epsilon_{1}}\right)^{\alpha}(b_{i}-a_{i})\leq\epsilon(b_{i}-a_{i}),

for each 1ik1\leq i\leq k. Thus we may conclude that

|{t[a,b]|ψ(t)vΨ}|ϵ(ba).|\{t\in[a,b]|\psi(t)v\in\Psi\}|\leq\epsilon(b-a).

Thus, (3.29) holds in all cases. ∎

4. Non-escape of mass

Let X{}X\cup\{\infty\} be the one-point compactification of

X:=SL(n,)/SL(n,).X:=\text{SL}(n,\mathbb{R})/\text{SL}(n,\mathbb{Z}).

Let φ:[0,)SL(n,)\varphi:[0,\infty)\to\text{SL}(n,\mathbb{R}) be a continuous unbounded curve definable in a polynomially bounded o-minimal structure, and for xXx\in X recall the measures μT,φ,x\mu_{T,\varphi,x} defined in (1.1).

By the Banach-Alaoglu theorem, any weak-star limit μ\mu as TT\to\infty of the measures μT,φ,x\mu_{T,\varphi,x} is a probability measure on X{}X\cup\{\infty\}. First will show the following. Then we will extend the result to the case of finite volume quotient spaces of Lie groups.

Proposition 4.1.

If φ:[0,)SL(n,)\varphi:[0,\infty)\to\text{SL}(n,\mathbb{R}) is a continuous unbounded non-contracting curve definable in a polynomially bounded o-minimal structure. Then, there exists T0>0T_{0}>0 such that given a compact set FSL(n,)F\subset\text{SL}(n,\mathbb{R}) and ϵ>0\epsilon>0, there exists a compact set KSL(n,)/SL(n,)K\subset\text{SL}(n,\mathbb{R})/\text{SL}(n,\mathbb{Z}) such that for any gFg\in F and TT0T\geq T_{0}, we have

|{t[T0,T]:φ(t)gSL(n,)/SL(n,)K}|(1ϵ)(TT0).\lvert\{t\in[T_{0},T]:\varphi(t)g\text{SL}(n,\mathbb{Z})/\text{SL}(n,\mathbb{Z})\in K\}\rvert\geq(1-\epsilon)(T-T_{0}).

We give some preliminaries before the proof. In the discussion that follows, we will refer to rank-kk discrete subgroups of n\mathbb{R}^{n} as kk-lattices. For a kk-lattice Λn\Lambda\subset\mathbb{Z}^{n}, let Λ\|\Lambda\| be the volume of a fundamental domain FSpan{Λ}F\leq\text{Span}_{\mathbb{R}}\{\Lambda\} of Λ\Lambda with respect to the usual measure on the subspace Span{Λ}n\text{Span}_{\mathbb{R}}\{\Lambda\}\leq\mathbb{R}^{n} (which is obtained by restricting the usual Euclidean inner product). Recall that if {v1,,vk}\{\textbf{v}_{1},...,\textbf{v}_{k}\} forms a \mathbb{Z}-basis for Λ\Lambda, then

Λ=v1vk,\|\Lambda\|=\|\textbf{v}_{1}\wedge\cdots\wedge\textbf{v}_{k}\|,

where the norm is the standard Euclidean norm defined through the inner product for which the pure wedges of kk-positively oriented tuples of the canonical basis vectors ei,i=1,,ne_{i},i=1,...,n are orthogonal and have norm equal to one. Consider the representation of SL(n,)\text{SL}(n,\mathbb{R}) on kn\bigwedge^{k}\mathbb{R}^{n} defined by

(4.1) g.(v1v2vk):=gv1gv2gvk,gSL(n,).g.(\textbf{v}_{1}\wedge\textbf{v}_{2}...\wedge\textbf{v}_{k}):=g\textbf{v}_{1}\wedge g\textbf{v}_{2}\wedge\cdots\wedge g\textbf{v}_{k},~{}g\in\text{SL}(n,\mathbb{R}).

For a fixed k<nk<n, the action of SL(n,)\text{SL}(n,\mathbb{R}) is transitive on the space of kk-lattices of rank k<nk<n. We denote k:=Span{e1,,ek}\mathbb{Z}^{k}:=\text{Span}_{\mathbb{Z}}\{e_{1},...,e_{k}\}, and observe that for a unimodular nn-lattice L=gnnL=g\mathbb{Z}^{n}\leq\mathbb{R}^{n}, where gSL(n,)g\in\text{SL}(n,\mathbb{R}), we have that

gγk,γSL(n,),g\gamma\mathbb{Z}^{k},\gamma\in\text{SL}(n,\mathbb{Z}),

is the collection of primitive kk-sublattices of LL. The following powerful theorem on quantitative non-divergence due to Kleinbock [Kle07] will be needed.

Theorem 4.2.

Suppose an interval BB\subset\mathbb{R}, C,α>0C,\alpha>0, 0<ρ<10<\rho<1 and a continuous map h:BSL(n,)h:B\to\text{SL}(n,\mathbb{R}) are given. Assume that for any γSL(n,)\gamma\in\text{SL}(n,\mathbb{Z}), and 1k<n1\leq k<n we have

  1. (1)

    the function xh(x)γ.e1ekx\to\|h(x)\gamma.e_{1}\wedge\cdots\wedge e_{k}\| is (C,α)(C,\alpha)-good on BB, and

  2. (2)

    supxBh(x)γ.e1ekρk\sup_{x\in B}\|h(x)\gamma.e_{1}\wedge\cdots\wedge e_{k}\|\geq\rho^{k}.

Then, for any ϵ<ρ\epsilon<\rho,

(4.2) |{xB:λ1(h(x)n)ϵ}|Cn2n(ϵρ)α|B|,|\{x\in B:\lambda_{1}(h(x)\mathbb{Z}^{n})\leq\epsilon\}|\leq Cn2^{n}\left(\frac{\epsilon}{\rho}\right)^{\alpha}|B|,

where λ1()\lambda_{1}(\cdot) is the function that outputs the length of the shortest nonzero vector of an Euclidean lattice.

Proof of Proposition 4.1.

We identify SL(n,)/SL(n,)\text{SL}(n,\mathbb{R})/\text{SL}(n,\mathbb{Z}) with the space of unimodular lattices

n:={L:=Span{v1,,vn}:det(vi,j)=1}.\mathcal{L}_{n}:=\{L:=\text{Span}_{\mathbb{Z}}\{v_{1},...,v_{n}\}:\det(v_{i,j})=1\}.

Consider

δ:={Ln:(L0)Bδ(0)}.\mathcal{B}_{\delta}:=\{L\in\mathcal{L}_{n}:(L\smallsetminus 0)\cap B_{\delta}(0)\neq\emptyset\}.

Here Bδ(0)nB_{\delta}(0)\subseteq\mathbb{R}^{n} is the ball of radius δ\delta centered at the origin. By Mahler’s Criterion ([BM00], Theorem 3,2), nδ\mathcal{L}_{n}\smallsetminus\mathcal{B}_{\delta} is compact for all δ>0\delta>0 and thus δ\mathcal{B}_{\delta} is a neighborhood of \infty. Fix gnng\mathbb{Z}^{n}\in\mathcal{L}_{n} for gSL(n,)g\in\text{SL}(n,\mathbb{R}), and consider the measures

(4.3) μT(f)=1T0Tf(φ(t)gn)𝑑t,fCc(n)\mu_{T}(f)=\frac{1}{T}\int_{0}^{T}f(\varphi(t)g\mathbb{Z}^{n})dt,~{}f\in C_{c}(\mathcal{L}_{n})

Let μ\mu be a weak-* limit of the measures μT\mu_{T} as TT\to\infty. In order to show that μ()=0\mu(\infty)=0, it is enough to prove that for every ϵ>0\epsilon>0, there is δ\delta such that

(4.4) lim supT|{t[0,T]:λ1(φ(t)gn)δ}|Tϵ.\displaystyle\limsup_{T\to\infty}\frac{|\{t\in[0,T]:\lambda_{1}(\varphi(t)g\mathbb{Z}^{n})\leq\delta\}|}{T}\leq\epsilon.

This will be concluded by Theorem 4.2 as follows. We first verify the conditions. For k{1,2,..,n1}k\in\{1,2,..,n-1\}, we denote by θk:SL(n,)kn\theta_{k}:\text{SL}(n,\mathbb{R})\to\bigwedge^{k}\mathbb{R}^{n} the representation (4.1). By Proposition 1.15, there exists T0,C,α>0T_{0},C,\alpha>0 such that for any fixed k{1,2,..,n1}k\in\{1,2,..,n-1\}, and γSL(n,)\gamma\in\text{SL}(n,\mathbb{Z}) it holds that

(4.5) Θ(t)=θk(φ(t)gγ)(e1ek)=φ(t)gγ.(e1ek)\displaystyle\Theta(t)=\|\theta_{k}(\varphi(t)g\gamma)(e_{1}\wedge\cdots\wedge e_{k})\|=\|\varphi(t)g\gamma.(e_{1}\wedge\cdots\wedge e_{k})\|

is (C,α)(C,\alpha)-good in [T0,)[T_{0},\infty). Let FSL(n,)F\subset\text{SL}(n,\mathbb{R}) be a given compact subset. Then

ρ:=(inf{φ(T0)gγ.(e1ek):gF,γSL(n,)})1k>0.\rho:=(\inf\{\|\varphi(T_{0})g\gamma.(e_{1}\wedge\cdots\wedge e_{k})\|:g\in F,\,\gamma\in\text{SL}(n,\mathbb{Z})\})^{\frac{1}{k}}>0.

Thus, by Theorem 4.2, for any gFg\in F,

(4.6) |{t[T0,T]:λ1(φ(t)gn)δ}|C(δρ)α(TT0),\displaystyle|\{t\in[T_{0},T]:\lambda_{1}(\varphi(t)g\mathbb{Z}^{n})\leq\delta\}|\leq C\left(\frac{\delta}{\rho}\right)^{\alpha}(T-T_{0}),

4.1. Non-divergence for homogeneous space of Lie groups

In the setting of Lie groups, one obtains the following result.

Proposition 4.3.

Let GG be a Lie subgroup of SL(n,)\text{SL}(n,\mathbb{R}) and let Γ\Gamma be a lattice in GG. Suppose that φ:[0,)SL(n,)\varphi:[0,\infty)\to\text{SL}(n,\mathbb{R}) is a non-contracting curve definable in a polynomially bounded o-minimal structure 𝒮\mathscr{S}. Assume further that φ([0,))Gu\varphi([0,\infty))\subset G_{u}. Then, there exists a T0>0T_{0}>0 such that given ϵ>0\epsilon>0 and a compact set FGF\subset G, there exists a compact set KG/ΓK\subset G/\Gamma such that for all gFg\in F and all T>T0T>T_{0},

(4.7) |{t[T0,T]:φ(t)gΓK}|(1ϵ)(TT0).\lvert\{t\in[T_{0},T]:\varphi(t)g\Gamma\in K\}\rvert\geq(1-\epsilon)(T-T_{0}).
Remark 4.4.

We note that the assumption that φ([0,))Gu\varphi([0,\infty))\subset G_{u} is non restrictive for our purposes in view of the following. In Proposition 5.13, we show that for a curve contained in a Lie subgroup GSL(n,)G\subseteq\text{SL}(n,\mathbb{R}) which is non-contracing and definable in a polynomially bounded o-minimal structure, it holds that its hull is contained in GuG_{u}. Thus, by multiplying the original curve with a correcting curve, the assumption φ([0,))Gu\varphi([0,\infty))\subset G_{u} is satisfied.

Proof.

We follow the arguments as in [DM93, Theorem 6.1] to reduce the problem to the case when GG is a semisimple group and Γ\Gamma is an irreducible lattice in GG.

We note that if G/ΓG/\Gamma is compact, there is nothing to prove. So, we now assume that G/ΓG/\Gamma is not compact. Let MM be the smallest closed normal subgroup of GG such that G¯=G/M\bar{G}=G/M is semisimple with the trivial center and no compact factors. Then M/(MΓ)M/(M\cap\Gamma) is compact, see [Rag72, Theorem 8.24] and [Sha96, Proof of Theorem 2.2]. Let q:GG¯q:G\to\bar{G} denote the quotient homomorphism. Then q(Γ)q(\Gamma) is a lattice in G¯\bar{G}, and the natural quotient map q¯:G/ΓG¯/q(Γ)\bar{q}:G/\Gamma\to\bar{G}/q(\Gamma) is proper. We note that if G~\tilde{G} and M~\tilde{M} denote the Zariski closures of GG and MM in SL(n,)\text{SL}(n,\mathbb{R}), respectively, then G¯=G/M=G~0/M~0\bar{G}=G/M=\tilde{G}^{0}/\tilde{M}^{0}. Therefore, q|Gu:GuG¯q|_{G_{u}}:G_{u}\to\bar{G} is a rational homomorphism of real algebraic groups. Thus qφq\circ\varphi is a curve definable in 𝒮\mathscr{S}. By Lemma 2.15, qφ:[0,)G¯q\circ\varphi:[0,\infty)\to\bar{G} is a non-contracting. Since q¯\bar{q} is proper, proving the theorem for G¯\bar{G} in place of GG is sufficient.

Therefore, we assume that GG is semisimple with the trivial center and no compact factors. Then, by [Rag72, Theorem 5.22], there exist closed normal subgroups G1,,GkG_{1},\ldots,G_{k} of GG for some kk such that GG equals the direct product G1××GkG_{1}\times\cdots\times G_{k}, such that for Γi=GiΓ\Gamma_{i}=G_{i}\cap\Gamma is an irreducible lattice in GiG_{i}. Moreover, Γ1××Γk\Gamma_{1}\times\cdots\times\Gamma_{k} is a normal subgroup of finite index in Γ\Gamma. Hence, G/ΓG/\Gamma is finitely covered by G1/Γ1××Gk/ΓkG_{1}/\Gamma_{1}\times\cdots\times G_{k}/\Gamma_{k}. Let qi:GGiq_{i}:G\to G_{i} denote the natural factor map, and observe that qiφq_{i}\circ\varphi is a non-contracting curve in GiG_{i} definable in 𝒮\mathscr{S} by Lemma 2.15. So, proving the theorem for Gi/ΓiG_{i}/\Gamma_{i} separately for each ii will imply the theorem for G/ΓG/\Gamma. Therefore, without loss of generality, we may assume that Γ\Gamma is an irreducible lattice in GG, which is semisimple with the trivial center and no compact factors.

First, suppose that the real rank of GG is at least 22. Then, by the arithmeticity theorem due to G. A. Margulis, Γ\Gamma is arithmetic; that is, see [Zim84, page 3]: There exists an mm\in\mathbb{N}, an algebraic semisimple group HSL(m,)H\subset\text{SL}(m,\mathbb{R}) defined over \mathbb{Q} and a surjective algebraic homomorphism

ρ:H0G,\rho:H^{0}\to G,

such that kerρ\ker\rho is compact and ρ(H0())\rho(H^{0}(\mathbb{Z})) is commensurable with Γ\Gamma, where H0()=H0SL(m,)H^{0}(\mathbb{Z})=H^{0}\cap\text{SL}(m,\mathbb{Z}). Since GG has no compact factors, the map ρ\rho restricted to HuH_{u} is a finite cover of GG. By Lemma 2.19, we can lift φ\varphi to a 𝒮\mathscr{S}-definable curve ψ:[0,)HuH0SL(m,)\psi:[0,\infty)\to H_{u}\subset H^{0}\subset\text{SL}(m,\mathbb{R}) which is non contracting in HuH_{u}, such that ρψ=φ\rho\circ\psi=\varphi. By Proposition 2.20, ψ\psi in non-contracting in SL(m,)\text{SL}(m,\mathbb{R}). As in Proposition 4.1, let T0>0T_{0}>0 be such that for any compact set F1SL(m,)F_{1}\subset\text{SL}(m,\mathbb{R}), there exists a compact set K1YK_{1}\subset Y such that for any TT0T\geq T_{0} and hF1h\in F_{1}

(4.8) |{t[T0,T]:ψ(t)hSL(m,)K1}|(1ϵ)(TT0).\lvert\{t\in[T_{0},T]:\psi(t)h\text{SL}(m,\mathbb{Z})\in K_{1}\}\rvert\geq(1-\epsilon)(T-T_{0}).

Choose, F1=ρ1(F)H0F_{1}=\rho^{-1}(F)\subset H^{0}. Since H0/H0()Y:=SL(m,)/SL(m,)H^{0}/H^{0}(\mathbb{Z})\hookrightarrow Y:=\text{SL}(m,\mathbb{R})/\text{SL}(m,\mathbb{Z}) is a proper H0H^{0}-equivariant continuous injection, we can pick a compact set F2H0F_{2}\subset H^{0} such that F2SL(m,)K1H0SL(m,)F_{2}\text{SL}(m,\mathbb{Z})\supset K_{1}\cap H^{0}\text{SL}(m,\mathbb{Z}). Since Γρ(H0())\Gamma\cap\rho(H^{0}(\mathbb{Z})) contains a subgroup of finite index in ρ(H0())\rho(H^{0}(\mathbb{Z})), we can pick a finite set SH0()S\subset H^{0}(\mathbb{Z}) such that Sρ1(Γ)H0()S\rho^{-1}(\Gamma)\supset H^{0}(\mathbb{Z}). Now K:=π(ρ(F2S))K:=\pi(\rho(F_{2}S)) is a compact subset of G/ΓG/\Gamma. Then, for any tT0t\geq T_{0} and gFg\in F, let hρ1(g)F1h\in\rho^{-1}(g)\subset F_{1}. Let t[T0,T]t\in[T_{0},T] such that ψ(t)hSL(m,)K1\psi(t)h\text{SL}(m,\mathbb{Z})\in K_{1}. Then ψ(t)hH0()F2H0()\psi(t)hH^{0}(\mathbb{Z})\in F_{2}H^{0}(\mathbb{Z}), and hence π(φ(t)g)K\pi(\varphi(t)g)\in K. Therefore, for any gFg\in F, and hρ1(g)F1h\in\rho^{-1}(g)\subset F_{1}, we have

{t[T0,T]:φ(t)gΓK}{t[T0,T]:ψ(t)hSL(m,)K1}.\{t\in[T_{0},T]:\varphi(t)g\Gamma\in K\}\supset\{t\in[T_{0},T]:\psi(t)h\text{SL}(m,\mathbb{Z})\in K_{1}\}.

From this and by (4.8), we obtain (4.7). This completes the proof when the real rank of GG is at least 22.

Now suppose that the real rank of GG is 11. It is straightforward to adapt the proof of [Dan84, Proposition 1.2] to conclude (4.7). For this purpose, it is enough to use the following property of the map φ\varphi in place of [Dan84, Lemmas 2.5 and 2.7]: Let V=d𝔤V=\wedge^{d}\mathfrak{g}, where dd is the dimension of the maximal unipotent subgroup of GG and 𝔤\mathfrak{g} the Lie algebra of GG. Now, GG acts on VV via dAd\wedge^{d}\text{Ad}. Then there exists T>0T>0, C>0C>0 and α>0\alpha>0 such that for any gGg\in G, the map tφ(t)g.pt\mapsto\lVert\varphi(t)g.p\rVert has the (C,α)(C,\alpha)-good property on [T0,)[T_{0},\infty) by Proposition 1.15.

As an immediate consequence of Proposition 4.3 we obtain the following.

Theorem 4.5.

If φ:[0,)SL(n,)\varphi:[0,\infty)\to\text{SL}(n,\mathbb{R}) be a continuous unbounded non-contracting curve definable in a polynomially bounded o-minimal structure. Let GSL(n,)G\leq\text{SL}(n,\mathbb{R}) be a Lie subgroup, Γ\Gamma be a lattice in GG and suppose that φ([0,))Gu\varphi([0,\infty))\subset G_{u}. Then any weak-* limit μ\mu of μT,φ,x\mu_{T,\varphi,x} has μ()=0\mu(\infty)=0 for all xG/Γx\in G/\Gamma.

5. Unipotent invariance

We begin with the following general statement.

Proposition 5.1.

Let GGL(n,)G\leq\text{GL}(n,\mathbb{R}) be a closed subgroup. Consider a continuous curve φ:[0,)G\varphi:[0,\infty)\to G which is unbounded and definable in a polynomially bounded o-minimal structure such that det(φ(t))\det(\varphi(t)) is constant in tt. Suppose further that for all vn{0}v\in\mathbb{R}^{n}\smallsetminus\{0\}, limtφ(t)v0\lim_{t\to\infty}\varphi(t)v\neq 0. Then there exists a nontrivial one-parameter unipotent group PGP\leq G such that if ΓG\Gamma\leq G is a discrete subgroup, gGg\in G and ν\nu is weak-\ast limit of the measures

νT,φ,gΓ(f):=1T0Tf(φ(t)gΓ)𝑑t,fCc(G/Γ),\nu_{T,\varphi,g\Gamma}(f):=\frac{1}{T}\int_{0}^{T}f(\varphi(t)g\Gamma)dt,f\in C_{c}(G/\Gamma),

then ν\nu is invariant under PP.

The main idea behind the above statement is the fact that o-minimal curves exhibit “tangency at infinity”. More precisely, there are “change of speed” maps h(t)h(t) with limth(t)=\lim_{t\to\infty}h(t)=\infty, such that the limit of the differences limtφ(h(t))φ(t)1=ρ\lim_{t\to\infty}\varphi(h(t))\varphi(t)^{-1}=\rho exists. Importantly, under our polynomially boundedness and the non-contraction assumption, those matrices ρ\rho will generate a one-parameter unipotent group, and any limiting measure of averaging along φ\varphi will be invariant under ρ\rho.

Such an idea was used in [Sha94] for averaging along polynomial curves. In the polynomial case, it is possible to use Taylor expansion in order to conclude such change of speed maps. For the o-minimal case, Peterzil and Steinhorn in [PS99] showed that the collection of all possible limits of the form limtφ(h(t))φ(t)1\lim_{t\to\infty}\varphi(h(t))\varphi(t)^{-1}, where hh is a definable function converging to \infty forms a one-dimensional torsion free group. We will refer to this group as the Peterzil-Steinhorn subgroup, which we define below in Definition 5.3 in a simplified manner which suits our needs. Poulios in his thesis [Pou13] studied further the Peterzil-Steinhorn groups, and he proves that the P.S. group for an unbounded curve definable in polynomially bounded structure is either unipotent or \mathbb{R}-diagonalizable. Importantly, Poulios establishes a convenient condition for the Peterzil-Steinhorn group to be unipotent.

We now recall the required notions and results from [Pou13]. For the following, for r<1r<1, consider

(5.1) hr,s(t):=(t1r+(1r)s)11r=t+str+o(tr), where s,h_{r,s}(t):=\left(t^{1-r}+(1-r)s\right)^{\frac{1}{1-r}}=t+st^{r}+o(t^{r}),\text{ where }s\in\mathbb{R},

and let

(5.2) h1,s(t):=st, where s>0.h_{1,s}(t):=st,\text{ where }s>0.
Proposition 5.2.

[Pou13, Chapter 3] Let GGL(n,)G\leq\text{GL}(n,\mathbb{R}) be a closed subgroup, and consider an unbounded curve φ:[0,)G\varphi:[0,\infty)\to G definable in a polynomially bounded o-minimal structure. Then, there exists a unique r1r\leq 1 such that the limit

(5.3) Mφ:=limttrφ(t)φ(t)1M_{\varphi}:=\lim_{t\to\infty}t^{r}\cdot\varphi^{\prime}(t)\cdot\varphi(t)^{-1}

exists and non-zero. Notice that MφM_{\varphi} is naturally identified as a tangent vector in the Lie algebra 𝔤\mathfrak{g} of GG. We have that MφM_{\varphi} is nilpotent \iff r<1r<1, and MφM_{\varphi} is \mathbb{R}-diagonalizable r=1\iff r=1. We let 𝔭=Span{Mφ}𝔤\mathfrak{p}=\text{Span}_{\mathbb{R}}\{M_{\varphi}\}\leq\mathfrak{g}, and denote as PGP\leq G the connected subgroup with Lie-algebra 𝔭\mathfrak{p}. Then:

  1. (1)

    r<1r<1 \iff for each ss\in\mathbb{R}, the limit

    (5.4) ρ(s):=limtφ(hr,s(t))φ(t)1,\rho(s):=\lim_{t\to\infty}\varphi(h_{r,s}(t))\varphi(t)^{-1},

    exists. In this case ρ\rho defines an isomorphism P\mathbb{R}\to P. Here \mathbb{R} denotes the additive group of real numbers.

  2. (2)

    r=1r=1\iff for each s>0s>0, the limit

    (5.5) ρ(s):=limtφ(h1,s(t))φ(t)1,\rho(s):=\lim_{t\to\infty}\varphi(h_{1,s}(t))\varphi(t)^{-1},

    exists. In this case ρ\rho defines an isomorphism ρ:>0P\rho:\mathbb{R}_{>0}\to P. Here >0\mathbb{R}_{>0} denotes the multiplicative group of positive real numbers.

Definition 5.3.

We say that r1r\leq 1 is the P.S. order of an unbounded curve φ:[0,)GL(n,)\varphi:[0,\infty)\to\text{GL}(n,\mathbb{R}) definable in a polynomially bounded o-minimal structure, if rr satisfies (5.3), and we call the one-parameter subgroup generated by exp(Mφ)\exp(M_{\varphi}) the P.S. group of φ\varphi.

Remark 5.4.

We note that if φ\varphi as in Proposition 5.2 is a restriction of one-parameter group to [0,)[0,\infty), then r=0r=0, hr,s(t)=t+sh_{r,s}(t)=t+s, and ρ(s)=φ(s)\rho(s)=\varphi(s) for all ss\in\mathbb{R}; that is, φ\varphi is a restriction of a unipotent one-parameter subgroup.

5.1. Unipotent P.S. groups and limiting unipotent invariance

We now show that when the P.S. group is unipotent, any limiting measure of measures of the form (1.1) will be invariant under the P.S. group. This proof is found in [Sha94] and in [PS18], and we give it here also for completeness. We note the following elementary calculus lemma.

Lemma 5.5.

Let h:(0,)h:(0,\infty)\to\mathbb{R} be a differentiable function such that limth(t)=1\lim_{t\to\infty}h^{\prime}(t)=1. Then, for any bounded continuous function F:[0,)F:[0,\infty)\to\mathbb{R}, we have

(5.6) 1T0T[F(h(t))F(t)]𝑑t=0\frac{1}{T}\int_{0}^{T}[F(h(t))-F(t)]dt=0

Notice that for all r<1r<1, we have hs,r(t)=1+O(tr1)h^{\prime}_{s,r}(t)=1+O(t^{r-1}).

Lemma 5.6.

Let GGL(n,)G\leq\text{GL}(n,\mathbb{R}) be a closed subgroup and let ΓG\Gamma\leq G be a discrete subgroup. Suppose that φ:[0,)G\varphi:[0,\infty)\to G is a continuous, unbounded curve, definable in a polynomially bounded o-minimal structure such that the P.S. group is unipotent (namely, (5.3) holds for r<1r<1). Let ρ\rho as in (5.4). Then

(5.7) limT1T0T{f(ρ(s)φ(t)x)f(φ(t)x)}𝑑t=0,\lim_{T\to\infty}\frac{1}{T}\int_{0}^{T}\{f(\rho(s)\varphi(t)x)-f(\varphi(t)x)\}dt=0,

for all fCc(G/Γ)f\in C_{c}(G/\Gamma), xG/Γx\in G/\Gamma and ss\in\mathbb{R}.

Proof.

For fCc(G/Γ)f\in C_{c}(G/\Gamma) and an ϵ>0\epsilon>0, by uniform continuity of ff, there is an TϵT_{\epsilon} such that

|f(ρ(s)φ(t)x)f(φ(hr,s(t)x)|\displaystyle|f(\rho(s)\varphi(t)x)-f(\varphi(h_{r,s}(t)x)| =|f(ρ(s)φ(t)x)f([φ(hr,s(t))φ(t)1]φ(t)x)|\displaystyle=|f(\rho(s)\varphi(t)x)-f([\varphi(h_{r,s}(t))\varphi(t)^{-1}]\varphi(t)x)|
ϵ/2,\displaystyle\leq\epsilon/2,

for all tTϵt\geq T_{\epsilon}. Now

1T0T|f(ρ(s)φ(t)x)f(φ(t)x)|𝑑t1T0Tϵ|f(ρ(s)φ(t)x)f(φ(t)x)|𝑑t\displaystyle\frac{1}{T}\int_{0}^{T}|f(\rho(s)\varphi(t)x)-f(\varphi(t)x)|dt\leq\frac{1}{T}\int_{0}^{T_{\epsilon}}|f(\rho(s)\varphi(t)x)-f(\varphi(t)x)|dt
+\displaystyle+ 1TTϵT|f(ρ(s)φ(t)x)f(φ(t)x)|𝑑t1T0Tϵ|f(ρ(s)φ(t)x)f(φ(t)x)|𝑑t\displaystyle\frac{1}{T}\int_{T_{\epsilon}}^{T}|f(\rho(s)\varphi(t)x)-f(\varphi(t)x)|dt\leq\frac{1}{T}\int_{0}^{T_{\epsilon}}|f(\rho(s)\varphi(t)x)-f(\varphi(t)x)|dt
+\displaystyle+ 1TTϵT|f(φ(hr,s(t)x)f(φ(t)x)|dt+ϵ2.\displaystyle\frac{1}{T}\int_{T_{\epsilon}}^{T}|f(\varphi(h_{r,s}(t)x)-f(\varphi(t)x)|dt+\frac{\epsilon}{2}.

Now take TT\to\infty. The first term goes to zero by boundedness of the function ff, and the second term goes to zero by Lemma 5.5. ∎

Poulios’ observations were used to establish invariance for polynomially bounded o-minimal curves in nilmanifolds by Peterzil and Starchenko, see [PS18]. If a curve is contained in a unipotent group, as in [PS18], then the P.S. group is automatically unipotent. Here, in order to verify that the P.S. group is unipotent when the curve is non-contracting we give the following new observation.

Definition 5.7.

We call a curve {φ(t)}GL(n,)\{\varphi(t)\}\subset\text{GL}(n,\mathbb{R}) essentially diagonal if there exists a decomposition

(5.8) φ(t)=σ(t)b(t)C,\varphi(t)=\sigma(t)b(t)C,

where b(t)GL(n,)b(t)\in\text{GL}(n,\mathbb{R}) is a diagonal matrix for all large tt, CGL(n,)C\in\text{GL}(n,\mathbb{R}) is a constant matrix, and σ(t)SL(n,)\sigma(t)\in\text{SL}(n,\mathbb{R}) is convergent as tt\to\infty.

The following is our key observation.

Proposition 5.8.

Let φ:[0,)GL(n,)\varphi:[0,\infty)\to\text{GL}(n,\mathbb{R}) be an unbounded continuous curve definable in a polynomially bounded o-minimal structure. Suppose that det(φ(t))\det(\varphi(t)) is constant for all large enough tt. Then, the P.S. group of φ\varphi is unipotent if and only if the curve φ\varphi is not essentially diagonal.

We will prove the above Proposition 5.8 in the subsection below. Before we proceed, we consider the following lemma which proves Proposition 5.1 by assuming Proposition 5.8 in combination with Lemma 5.6.

Lemma 5.9.

Let φ:[0,)GL(n,)\varphi:[0,\infty)\to\text{GL}(n,\mathbb{R}) be an unbounded continuous curve definable in a polynomially bounded o-minimal structure. Suppose that the determinant det(φ(t))\det(\varphi(t)) is constant in tt and suppose that limtφ(t)v0,\lim_{t\to\infty}\varphi(t)v\neq 0, for all vn{0}v\in\mathbb{R}^{n}\smallsetminus\{0\}. Then, φ\varphi is not essentially diagonal and in particular, the P.S. group of φ\varphi is unipotent.

Proof.

Suppose for contradiction that the P.S. group is not unipotent. Then, by Proposition 5.8, φ\varphi is essentially diagonal. Namely φ(t)=σ(t)b(t)C\varphi(t)=\sigma(t)b(t)C, where CC, σ(t)SL(n,)\sigma(t)\in\text{SL}(n,\mathbb{R}) are bounded and b(t)b(t) is a diagonal matrix with det(b(t))=c0\det(b(t))=c_{0} for all tt. Since we assume that φ\varphi is unbounded, it follows that b(t)b(t) is unbounded, which in turn implies that there is an index 1in1\leq i\leq n such that the corresponding entry function bii(t)b_{ii}(t) on the diagonal of b(t)b(t) decays to zero. In particular, for v:=C1eiv:=C^{-1}e_{i} it holds that limtφ(t)v=0\lim_{t\to\infty}\varphi(t)v=0, which is a contradiction. ∎

5.1.1. Proving Proposition 5.8

We will call the (i,j)(i,j) index of an upper-triangular matrix bSL(n,)b\in\text{SL}(n,\mathbb{R}) of the form

(5.9) b=[f1,100000000000000fi,i00𝒇𝒊,𝒋fi,n0fj,jfj,n000fn,n],b=\begin{bmatrix}f_{1,1}&0&\cdots&0&0&\cdots&0&0&\cdots&0\\ 0&\ddots&\ddots&\vdots&\vdots&&\vdots&\vdots&&\vdots\\ \vdots&\ddots&\ddots&0&0&\cdots&0&0&\cdots&0\\ 0&\cdots&0&f_{i,i}&0&\cdots&0&\boldsymbol{f_{i,j}}&\cdots&f_{i,n}\\ \vdots&&\vdots&\ddots&\ddots&&&\vdots&&\vdots\\ \vdots&&\vdots&&\ddots&\ddots&&\vdots&&\vdots\\ \vdots&&\vdots&&&\ddots&\ddots&\vdots&&\vdots\\ \vdots&&\vdots&&&&0&f_{j,j}&\cdots&f_{j,n}\\ \vdots&&\vdots&&&&&\ddots&\ddots&\vdots\\ 0&\cdots&0&\cdots&\cdots&\cdots&\cdots&\cdots&0&f_{n,n}\end{bmatrix},

where fi,j0f_{i,j}\neq 0, the first non-zero off-diagonal entry. More precisely, (i,j)(i,j) is the first non-zero entry among the off-diagonal entries according to the following lexicographic order on 2\mathbb{N}^{2}:

(5.10) (i,j)(k,l)i<k or (i=k and j<l).\displaystyle(i,j)\prec(k,l)\iff i<k\text{ or }(i=k\text{ and }j<l).

Let BSL(n,)B\leq\text{SL}(n,\mathbb{R}) be the subgroup of upper-triangular matrices. Let b:[0,)Bb:[0,\infty)\to B be a definable curve. Recall that each definable function f(t)f(t) is either zero for all large enough tt or |f(t)|>0|f(t)|>0 for all large tt. Thus, for all large enough tt, there exists a unique first non-zero off-diagonal entry in b(t)b(t), or b(t)b(t) is diagonal for all large tt. We will refer to this entry as the first non-zero off-diagonal entry of the curve b(t)b(t).

Lemma 5.10.

Let φ:[0,)SL(n,)\varphi:[0,\infty)\to\text{SL}(n,\mathbb{R}) be a curve definable in an polynomially bounded o-minimal structure. Then there is a definable curve b:[0,)Bb:[0,\infty)\to B, such that b(t)b(t) is either diagonal for all large tt, or the first non-zero off-diagonal entry fi,jf_{i,j} satisfies

  1. (1)

    degfi,idegfi,j\deg f_{i,i}\neq\deg f_{i,j}, and

  2. (2)

    degfi,j>degfj,j\deg f_{i,j}>\deg f_{j,j},

and importantly,

(5.11) φ(t)=σ(t)b(t)C,\varphi(t)=\sigma(t)b(t)C,

where CSL(n,)C\in\text{SL}(n,\mathbb{R}) is constant, and σ(t)SL(n,)\sigma(t)\in\text{SL}(n,\mathbb{R}) is convergent as tt\to\infty.

Proof.

Using the KAN decomposition, we first write φ(t)SL(n,)\varphi(t)\in\text{SL}(n,\mathbb{R}), as φ(t)=k(t)p(t)\varphi(t)=k(t)p(t), where k(t)SO(n,)k(t)\in\text{SO}(n,\mathbb{R}) and p(t)Bp(t)\in B an upper-triangular matrix. Since k(t)k(t) is obtained by performing the Gram-Schimdt process on the columns of φ\varphi, we conclude that k(t)k(t) is definable. As a consequence, p(t)p(t) is definable. Since k(t)k(t) is bounded and definable, limtk(t)\lim_{t\to\infty}k(t) exists. For all tt large, p(t)p(t) is either diagonal or p(t)p(t) takes the form of (5.9).

We employ the following algorithm to achieve the outcome described in the lemma. At each step, we perform either a column or a row operation on p(t)p(t):

  1. (1)

    All off-diagonal entries are eventually 0. Then we are done.

  2. (2)

    We pick the first index (i,j)(i,j) (with i<ji<j) such that the (i,j)(i,j)-th entry is not eventually 0, and proceed to the next step.

  3. (3)

    Suppose that deg(fi,j)=deg(fi,i)\deg(f_{i,j})=\deg(f_{i,i}). Then there is an cc\in\mathbb{R} such that deg(fi,jcfi,i)<deg(fi,i)\deg(f_{i,j}-cf_{i,i})<\deg(f_{i,i}). For this step, subtract from the jj-th column the ii-th column multiplied by cc. This amounts to multiplying p(t)p(t) by a constant unipotent matrix from the right. The obtained matrix is the same besides the i,ji,j-th entry which is replaced with fi,jcfi,if_{i,j}-cf_{i,i}. There are two possibilities now:

    1. (a)

      fi,icfi,jf_{i,i}-cf_{i,j} is eventually zero. If all off-diagonal entries are eventually 0, then we are done. Otherwise, the first eventually non-zero off-diagonal entry in the resulting matrix has a strictly larger index (according to the lexicographic order), and we go back to step 2.

    2. (b)

      fi,icfi,jf_{i,i}-cf_{i,j} is not eventually zero: Then the first requirement of Lemma 5.10 is satisfied. We continue then with the following step.

  4. (4)

    If deg(fi,j)>deg(fj,j)\deg(f_{i,j})>\deg(f_{j,j}), then the second requirement of Lemma 5.10 is satisfied, and we are done, otherwise we proceed to the next step.

  5. (5)

    Now suppose deg(fi,j)deg(fj,j)\deg(f_{i,j})\leq\deg(f_{j,j}). Then subtract from the ii-th row the jj-th row multiplied by fi,jfj,j\frac{f_{i,j}}{f_{j,j}}. This amounts to multiplying from the left by a unipotent matrix, which converges as tt\to\infty. This is so because

    deg(fi,jfj,j)=deg(fi,j)deg(fj,j)0.\deg\left(\frac{f_{i,j}}{f_{j,j}}\right)=\deg(f_{i,j})-\deg(f_{j,j})\leq 0.

    Then, either all the off-diagonal entries in the resulting matrix are eventually 0, and we are done; or the first eventually non-zero off-diagonal entry in the resulting matrix has a strictly larger index (according to the lexicographic order), and we go back to step 2.

The algorithm ends with finitely many steps with either a diagonal matrix or a definable curve b(t)b(t) satisfying the requirements of the lemma. ∎

Proof for Proposition 5.8.

First, note that it is enough to prove the statement for φSL(n,)\varphi\subset\text{SL}(n,\mathbb{R}) since the P.S. group is unchanged if we multiply the curve by a constant invertible matrix from the right. In the notations of Lemma 5.10:

φ(t)=σ(t)b(t)C,\varphi(t)=\sigma(t)b(t)C,

where limtσ(t)=gSL(n,).\lim_{t\to\infty}\sigma(t)=g\in\text{SL}(n,\mathbb{R}). Let r1r\leq 1 be such that (5.3) holds for {b(t)}\{b(t)\}. Let hs,r(t)h_{s,r}(t) be is as in (5.1)–(5.2). We note that

(5.12) limtφ(hs,r(t))φ(t)1=g[limtb(hs,r(t))b(t)1]g1.\lim_{t\to\infty}\varphi(h_{s,r}(t))\varphi(t)^{-1}=g\left[\lim_{t\to\infty}b(h_{s,r}(t))b(t)^{-1}\right]g^{-1}.

Therefore, by Proposition 5.2, the rr corresponding to (5.3) for φ(t)\varphi(t) and b(t)b(t) are the same.

Now, b(t)b(t) is either eventually diagonal or upper-triangular and satisfies the conditions of the Lemma 5.10 for the first (eventually) non-zero off-diagonal entry.

If b(t)b(t) is eventually diagonal, then for all s>0s>0, limtb(h1,s(t))b1(t)\lim_{t\to\infty}b(h_{1,s}(t))b^{-1}(t) converges to a diagonal matrix. Namely, the P.S. group is diagonal in this case.

Otherwise, let (i,j)(i,j) be the first non-zero off-diagonal entry in b(t)b(t) (for all large enough tt). We observe that (i,j)(i,j)-th entry in the matrix b(t)b(t)1b^{\prime}(t)b(t)^{-1} is

(5.13) fi,ifi,j+fi,jfi,ifi,ifj,j=(fi,jfi,i)fi,ifj,j.\displaystyle\frac{-f_{i,i}^{\prime}f_{i,j}+f_{i,j}^{\prime}f_{i,i}}{f_{i,i}f_{j,j}}=\left(\frac{f_{i,j}}{f_{i,i}}\right)^{\prime}\cdot\frac{f_{i,i}}{f_{j,j}}.

Since ri,i:=degfi,iri,j:=degfi,jr_{i,i}:=\deg f_{i,i}\neq r_{i,j}:=\deg f_{i,j}, we have that

deg(fi,jfi,i)=ri,jri,i1.\deg\left(\frac{f_{i,j}}{f_{i,i}}\right)^{\prime}={r_{i,j}-r_{i,i}-1}.

Thus,

deg((fi,jfi,i)fi,ifj,j)=ri,jrj,j1.\deg\left(\left(\frac{f_{i,j}}{f_{i,i}}\right)^{\prime}\cdot\frac{f_{i,i}}{f_{j,j}}\right)=r_{i,j}-r_{j,j}-1.

Recall by Lemma 5.10 that ri,jrj,j>0r_{i,j}-r_{j,j}>0. Thus the (i,j)(i,j)-th entry in tb(t)b(t)1tb^{\prime}(t)b(t)^{-1} is unbounded. So (5.5) is not satisfied. Hence r1r\neq 1. ∎

5.2. Unipotent invariance in quotients

For our purposes, it will not suffice only to establish unipotent invariance for our limiting measures, but we will also require unipotent invariance in certain quotients. We will now describe more precisely our motivation for the results in this section and how they play a role in the proof of Theorem 1.8 in Section 6.

Consider a non-contracting curve φ:[0,)G:=𝐆()\varphi:[0,\infty)\to G:=\mathbf{G}(\mathbb{R}) definable in a polynomially bounded o-minimal structure. By Theorem 4.5, there exists a subsequence of the measures μTi,φ,x\mu_{T_{i},\varphi,x} as in (1.1) that converges to a probability measure μ\mu as ii\to\infty. By Proposition 5.1, μ\mu is invariant under a nontrivial unipotent one-parameter subgroup. We will let WW be the subgroup generated by all unipotent one-parameter subgroups that preserve μ\mu. Then, using Ratner’s theorem describing finite ergodic invariant measures for WW, and the linearization technique (which is applicable due to the (C,α)(C,\alpha)-good property for the norm of trajectories of φ(t)\varphi(t) on representations of SL(n,)\text{SL}(n,\mathbb{R})), we will prove that φ(t)\varphi(t) is contained in a certain algebraic subgroup NN such that WNW\trianglelefteq N and that HφNH_{\varphi}\subseteq N. Our goal will be to prove that HφWH_{\varphi}\subset W. To achieve this goal, we would like to show that if φ(t)\varphi(t) is not bounded modulo WW, then the image of φ\varphi in N/WN/W, which is definable, is, in fact, non-contracting, and hence its corresponding P.S. group in N/WN/W is unipotent. Consequently, we will show that μ\mu is invariant under a unipotent one-parameter subgroup of NN that is not contained in WW. This will contradict our choice of WW.

Remark 5.11.

As the following example demonstrates, one cannot expect a non-contracting curve modulo a subgroup to have a unipotent P.S. group. Consider φ(t):=[tt201/t]\varphi(t):=\begin{bmatrix}t&t^{2}\\ 0&1/t\end{bmatrix} for t1t\geq 1. Then φ\varphi is non-contracting in SL(2,)\text{SL}(2,\mathbb{R}), and its P.S. group is U:={u(s):=[1s01]:s}U:=\{u(s):=\begin{bmatrix}1&s\\ 0&1\end{bmatrix}:s\in\mathbb{R}\}. Now consider the group PSL(2,)P\leq\text{SL}(2,\mathbb{R}) of upper triangular matrices. Then, P/UP/U is naturally identified with the group ASL(2,)A\leq\text{SL}(2,\mathbb{R}) of diagonal matrices. In particular, φ(t)\varphi(t) modulo HH is a diagonal curve. In view of our result below, this is explained by the fact that the hull of the curve is not contained in PP.

We will prove the following.

Proposition 5.12.

Let ψ:[0,)SL(n,)\psi:[0,\infty)\to\text{SL}(n,\mathbb{R}) be an unbounded, non-contracting curve definable in a polynomially bounded o-minimal structure. Suppose that ψHψ\psi\subseteq H_{\psi}, where HψH_{\psi} is the hull of ψ\psi. Let NSL(n,)N\leq\text{SL}(n,\mathbb{R}) be a Zariski closed subgroup such that HψNH_{\psi}\leq N, and let HNH\trianglelefteq N be a closed normal subgroup. Let q:NN/Hq:N\to N/H be the natural map.

Assume that the image of ψ\psi in N/HuN/H_{u} is unbounded. Then, there exists r<1r<1 such that

(5.14) limtq(ψ(hs,r(t)))q(ψ(t))1=p(s),s,\lim_{t\to\infty}q(\psi(h_{s,r}(t)))q(\psi(t))^{-1}=p(s),s\in\mathbb{R},

where p(s),sp(s),s\in\mathbb{R} is a (non-trivial) one-parameter subgroup of N/HN/H. Here, hs,r(t)h_{s,r}(t) is given by (5.1). Importantly, there is a one-parameter unipotent subgroup UNU\leq N such that

q(U)={p(s):s}.q(U)=\{p(s):s\in\mathbb{R}\}.
Proof of Proposition 5.12.

Let H~u:=zcl(Hu)\tilde{H}_{u}:=\text{zcl}(H_{u}) with HuH_{u} being the subgroup generated by all one-parameter unipotent subgroups contained in HH. Since HNH\trianglelefteq N, it follows that HuNH_{u}\trianglelefteq N, and by Zariski density we get that H~uN\tilde{H}_{u}\trianglelefteq N. Now, HuH_{u} is of finite index in H~u\tilde{H}_{u}, therefore, since ψ\psi is unbounded in N/HuN/H_{u} it follows that ψ\psi is unbounded in N/H~uN/\tilde{H}_{u}.

By [Spr98, Theorem 5.5.3] and since H~u\tilde{H}_{u} has no non-trivial real character, there is a finite dimensional rational representation ρ:SL(n,)GL(V)\rho:\text{SL}(n,\mathbb{R})\to\text{GL}(V) with an vVv\in V such that

H~u={gSL(n,):ρ(g)v=v}.\tilde{H}_{u}=\{g\in\text{SL}(n,\mathbb{R}):\rho(g)v=v\}.

Now let

W:={xV:ρ(h)x=x,hH~u}.W:=\{x\in V:\rho(h)x=x,\forall h\in\tilde{H}_{u}\}.

We have that vWv\in W so that WW is not the trivial subspace. Because H~u\tilde{H}_{u} is a normal subgroup of NN, we get that the action of NN keeps WW invariant. Now recall that one-parameter unipotent subgroups generate HψH_{\psi}. Therefore, det(ρ(Hψ)|W)={1}\det(\rho(H_{\psi})|_{W})=\{1\}. In particular, det(ρ(ψ)|W)={1}\det(\rho(\psi)|_{W})=\{1\}.

Since ψ\psi is not contracting, we have in particular that limtρ(ψ(t))w0,wW{0}.\lim_{t\to\infty}\rho(\psi(t))w\neq 0,~{}\forall w\in W\smallsetminus\{0\}. Thus, by Lemma 5.9, there is an r<1r<1 such that

limtρ(ψ(hs,r(t))ψ(t)1)|W=:p¯(s),s,\lim_{t\to\infty}\rho(\psi(h_{s,r}(t))\psi(t)^{-1})|_{W}=:\bar{p}(s),s\in\mathbb{R},

where p¯(s)\overline{p}(s) is a one-parameter unipotent subgroup. It is a general fact if f:G1G2f:G_{1}\to G_{2} is a homomorphism of algebraic groups, and U2G2U_{2}\leq G_{2} is a unipotent one-parameter subgroup, then there is a unipotent one-parameter subgroup U1G1U_{1}\leq G_{1} such that f(U1)=U2f(U_{1})=U_{2}. This follows by Jordan decomposition. Thus, there is a unipotent one-parameter subgroup UNU\leq N such that ρ(U)|W={p¯(s):s}\rho(U)|_{W}=\{\bar{p}(s):s\in\mathbb{R}\}. Now consider the diagram:

(5.15) N{N}N/Hu{N/H_{u}}N/H~u{N/\tilde{H}_{u}}GL(W){\text{GL}(W)}ϑ\scriptstyle{\vartheta}ρ|W\scriptstyle{\rho|_{W}}τ\scriptstyle{\tau}τ~\scriptstyle{\tilde{\tau}}

Since images of algebraic groups under algebraic homomorhpisms are closed (see e.g. [Spr98, Proposition 2.2.5]), we conclude that τ~\tilde{\tau} is an injective proper map. Since HuH_{u} is of finite index in H~u\tilde{H}_{u}, the map τ\tau is a covering map with finite fibers, and it follows that τ~τ\tilde{\tau}\circ\tau is a proper map with finite fibers. Then, we obtain for each ss\in\mathbb{R},

(5.16) limtϑ(ψ(hs,r(t))ψ(t)1)=limtϑ(ψ(hs,r(t)))ϑ(ψ(t))1=:p~(s).\lim_{t\to\infty}\vartheta(\psi(h_{s,r}(t))\psi(t)^{-1})=\lim_{t\to\infty}\vartheta(\psi(h_{s,r}(t)))\vartheta(\psi(t))^{-1}=:\tilde{p}(s).

Now we show that p~(s),s\tilde{p}(s),s\in\mathbb{R} is a one-parameter group, where sp~(s)s\mapsto\tilde{p}(s) is an homomorphism from the additive group of real numbers. This argument appears in [Pou13] and given here for completeness. Note that for r<1r<1 and for all s,ls,l\in\mathbb{R} that

hs,rhl,r=hs+l,r.h_{s,r}\circ h_{l,r}=h_{s+l,r}.

Denote ξ(t):=ϑ(ψ(t))\xi(t):=\vartheta(\psi(t)). Now if (5.14) holds for r<1r<1, then for s,ls,l\in\mathbb{R} we have

p~(s)p~(l)\displaystyle\tilde{p}(s)\tilde{p}(l) =(limtξ(hs,r(t))ξ(t)1)(limtξ(hl,r(t))ξ(t)1)\displaystyle=(\lim_{t\to\infty}\xi(h_{s,r}(t))\xi(t)^{-1})(\lim_{t\to\infty}\xi(h_{l,r}(t))\xi(t)^{-1})
=(limtξ(hs,r(hl,r(t)))ξ(hl,r(t))1)(limtξ(hl,r(t))ξ(t)1)\displaystyle=(\lim_{t\to\infty}\xi(h_{s,r}(h_{l,r}(t)))\xi(h_{l,r}(t))^{-1})(\lim_{t\to\infty}\xi(h_{l,r}(t))\xi(t)^{-1})
=limtξ(hs,r(hl,r(t)))ξ(hl,r(t))1ξ(hl,r(t))ξ(t)1\displaystyle=\lim_{t\to\infty}\xi(h_{s,r}(h_{l,r}(t)))\xi(h_{l,r}(t))^{-1}\xi(h_{l,r}(t))\xi(t)^{-1}
=limtξ(hs+l,r(t))ξ(t)1=p~(s+l).\displaystyle=\lim_{t\to\infty}\xi(h_{s+l,r}(t))\xi(t)^{-1}=\tilde{p}(s+l).

Since the above diagram commutes, we get that ϑ(U)={p~(s):s}\vartheta(U)=\{\tilde{p}(s):s\in\mathbb{R}\} is a non-trivial subgroup of N/HuN/H_{u}. Finally, consider the natural quotient map η:N/HuN/H\eta:N/H_{u}\to N/H, and note that q=ηϑq=\eta\circ\vartheta. Let p(s):=η(p~(s))p(s):=\eta(\tilde{p}(s)). Then, due to (5.16), we get by continuity that for all ss\in\mathbb{R},

p(s)=limtηϑ(ψ(hs,r(t))ψ(t)1)=limtq(ψ(hs,r(t)))q(ψ(t))1.p(s)=\lim_{t\to\infty}\eta\circ\vartheta(\psi(h_{s,r}(t))\psi(t)^{-1})=\lim_{t\to\infty}q(\psi(h_{s,r}(t)))q(\psi(t))^{-1}.

Finally, since UHuU\not\subset H_{u} and since UU is unipotent, it follows that q(U)N/Hq(U)\leq N/H is not trivial, and q(U)=ηϑ(U)={p(s):s}.q(U)=\eta\circ\vartheta(U)=\{p(s):s\in\mathbb{R}\}.

5.3. Inclusion of the hull in a Lie group

Here, we will prove the following.

Proposition 5.13.

Let GSL(n,)G\leq\text{SL}(n,\mathbb{R}) be a connected unimodular closed subgroup, and let φ:[0,)G\varphi:[0,\infty)\to G be a non-contracting, continuous curve, definable in a polynomially bounded o-minimal structure. Then, HφGuH_{\varphi}\leq G_{u} where HφH_{\varphi} is the hull of φ\varphi and GuGG_{u}\leq G is the closed subgroup generated by the one-parameter unipotent subgroups contained in GG. In particular, both the hull and the range of the correcting curve of φ\varphi are in GG.

The proof will be based on Proposition 5.12 and the following key lemma.

Lemma 5.14.

Let 𝔤\mathfrak{g} be a Lie subalgebra of 𝔤𝔩(n,R)\mathfrak{gl}(n,R). Let 𝔤u\mathfrak{g}_{u} be the subalgebra generated by nilpotent matrices in 𝔤\mathfrak{g}. Then for any X𝔤X\in\mathfrak{g},

Trace(adX)=Trace(adX|𝔤u).\text{Trace}(\text{ad}X)=\text{Trace}(\text{ad}X|_{\mathfrak{g}_{u}}).
Proof.

We will be using repeatedly the following fact – If TT is a linear map on a vector space VV preserving a subspace WW, and TV/WT_{V/W} is the factored map on V/WV/W, then

Trace(T|V)=Trace(T|W)+Trace(TV/W),\text{Trace}(T|_{V})=\text{Trace}(T|_{W})+\text{Trace}(T_{V/W}),

where T|WT|_{W} is the restriction of TT to WW.

Let X𝔤X\in\mathfrak{g}. Then adX\text{ad}X acts trivially on 𝔤/[𝔤,𝔤]\mathfrak{g}/[\mathfrak{g},\mathfrak{g}]. Hence,

Trace(adX|𝔤)=Trace(adX|[𝔤,𝔤]).\text{Trace}(\text{ad}X|_{\mathfrak{g}})=\text{Trace}(\text{ad}X|_{[\mathfrak{g},\mathfrak{g}]}).

Let 𝔤~\tilde{\mathfrak{g}} denote the Lie algebra of the Zariski closure of the Lie group associated with 𝔤\mathfrak{g} in GL(n,)\text{GL}(n,\mathbb{R}). By [Che51, Chapter II, Theorem 13], [𝔤~,𝔤~]=[𝔤,𝔤][\tilde{\mathfrak{g}},\tilde{\mathfrak{g}}]=[\mathfrak{g},\mathfrak{g}]. Thus,

Trace(adX|𝔤)=Trace(adX|[𝔤~,𝔤~]).\text{Trace}(\text{ad}X|_{\mathfrak{g}})=\text{Trace}(\text{ad}X|_{[\tilde{\mathfrak{g}},\tilde{\mathfrak{g}}]}).

Let 𝔲\mathfrak{u} denote the radical of the [𝔤~,𝔤~][\tilde{\mathfrak{g}},\tilde{\mathfrak{g}}]. Then 𝔲\mathfrak{u} consists of nilpotent matrices; we can derive this from [Rag72, 2.5 A structure theorem]. We also note that [𝔤~,𝔤~][\tilde{\mathfrak{g}},\tilde{\mathfrak{g}}] and 𝔲\mathfrak{u} are ideals in 𝔤~\tilde{\mathfrak{g}}. Since [𝔤~,𝔤~]/𝔲[\tilde{\mathfrak{g}},\tilde{\mathfrak{g}}]/\mathfrak{u} is semisimple, the trace of the derivation on it corresponding to adX\text{ad}X is zero. Therefore,

Trace(adX|[𝔤~,𝔤~])=Trace(adX|𝔲).\text{Trace}(\text{ad}X|_{[\tilde{\mathfrak{g}},\tilde{\mathfrak{g}}]})=\text{Trace}(\text{ad}X|_{\mathfrak{u}}).

Now let 𝔳\mathfrak{v} denote the radical of 𝔤u\mathfrak{g}_{u}. Since 𝔤u\mathfrak{g}_{u} is a Lie subalgebra generated by nilpotent matrices, 𝔳\mathfrak{v} consists of nilpotent matrices. Also, 𝔤u\mathfrak{g}_{u}, and hence 𝔳\mathfrak{v} are ideals of 𝔤\mathfrak{g}. As above, since 𝔤u/𝔳\mathfrak{g}_{u}/\mathfrak{v} is semisimple, we conclude that

Trace(adX|𝔤u)=Trace(adX|𝔳).\text{Trace}(\text{ad}X|_{\mathfrak{g}_{u}})=\text{Trace}(\text{ad}X|_{\mathfrak{v}}).

It remains to prove:

Trace(adX|𝔳)=Trace(adX|𝔲).\text{Trace}(\text{ad}X|_{\mathfrak{v}})=\text{Trace}(\text{ad}X|_{\mathfrak{u}}).

Now 𝔲[𝔤~,𝔤~]=[𝔤,𝔤]𝔤\mathfrak{u}\subset[\tilde{\mathfrak{g}},\tilde{\mathfrak{g}}]=[\mathfrak{g},\mathfrak{g}]\subset\mathfrak{g}. Hence, 𝔲𝔤u\mathfrak{u}\subset\mathfrak{g}_{u}, and so 𝔲𝔳\mathfrak{u}\subset\mathfrak{v}. Moreover, 𝔳[𝔤~,𝔤~]\mathfrak{v}\cap[\tilde{\mathfrak{g}},\tilde{\mathfrak{g}}] is an ideal in 𝔤~\tilde{\mathfrak{g}}, so it is contained in 𝔲\mathfrak{u}. Thus,

𝔳[𝔤~,𝔤~]=𝔲.\mathfrak{v}\cap[\tilde{\mathfrak{g}},\tilde{\mathfrak{g}}]=\mathfrak{u}.

Therefore [X,𝔳]𝔳[𝔤~,𝔤~]=𝔲[X,\mathfrak{v}]\subset\mathfrak{v}\cap[\tilde{\mathfrak{g}},\tilde{\mathfrak{g}}]=\mathfrak{u}. Therefore, the action of adX\text{ad}X on 𝔳/𝔲\mathfrak{v}/\mathfrak{u} is zero. So,

Trace(adX|𝔳)=Trace(adX|𝔲).\text{Trace}(\text{ad}X|_{\mathfrak{v}})=\text{Trace}(\text{ad}X|_{\mathfrak{u}}).

Corollary 5.15.

Let GG be a closed connected Lie subgroup of GL(n,)\text{GL}(n,\mathbb{R}). Suppose that GG is unimodular. Let GuG_{u} be the closed subgroup generated by unipotent one-parameter subgroups contained in GG and let 𝔤u\mathfrak{g}_{u} denote the Lie algebra of GuG_{u}. Then for any gGg\in G,

det(Ad(g)|𝔤u)=1.\det(\text{Ad}(g)|_{\mathfrak{g}_{u}})=1.
Proof.

Suppose that g=exp(X)g=\exp(X). Then,

det(Ad(g)|𝔤u)=exp(Trace(adX|𝔤u)).\det(\text{Ad}(g)|_{\mathfrak{g}_{u}})=\exp(\text{Trace}(\text{ad}X|_{\mathfrak{g}_{u}})).

Then, Lemma 5.14 shows that det(Ad(g)|𝔤u)=det(Ad(g))\det(\text{Ad}(g)|_{\mathfrak{g}_{u}})=\det(\text{Ad}(g)). Since GG is connected and unimodular, we have that det(Ad(g))=1\det(\text{Ad}(g))=1. Then, det(Ad(g)|𝔤u)=1\det(\text{Ad}(g)|_{\mathfrak{g}_{u}})=1 for all gg in some identity neighborhood in GG. Since the identity neighborhood generates GG, the corollary follows. ∎

Proof of Proposition 5.13.

Let φ:[0,)G\varphi:[0,\infty)\to G be a non-contracting, continuous curve, definable in a polynomially bounded o-minimal structure. Let β\beta be a correcting curve for φ\varphi, and denote ψ(t):=β(t)φ(t)\psi(t):=\beta(t)\varphi(t).

Consider the exterior product of the Adjoint representation of SL(n,)\text{SL}(n,\mathbb{R}) on dim𝔤u𝔰𝔩(n,)\wedge^{\dim\mathfrak{g}_{u}}\mathfrak{sl}(n,\mathbb{R}). Let pdim𝔤u𝔤u{0}p\in\wedge^{\dim\mathfrak{g}_{u}}\mathfrak{g}_{u}\setminus\{0\}, and let NN be the stabilizer of pp in SL(n,)\text{SL}(n,\mathbb{R}). Note that NN is Zariski closed, NN is observable, and GuNG_{u}\trianglelefteq N. By Corollary 5.15, GNG\subset N. Since NN is observable, and φN\varphi\subset N, we conclude that HφNH_{\varphi}\subseteq N by the minimality of the hull, Theorem 2.4,(1). Now, by Remark 2.6, we have that ψHψ=Hφ\psi\subset H_{\psi}=H_{\varphi}. Assume for contradiction that ψ\psi is not bounded modulo GuG_{u}. Then, by Proposition 5.12, there exists a one-parameter unipotent subgroup of {u(s)}sN\{u(s)\}_{s\in\mathbb{R}}\leq N such that

q(u(s))=limtq(ψ(hs,r(t)))q(ψ(t))1,s.q(u(s))=\lim_{t\to\infty}q(\psi(h_{s,r}(t)))q(\psi(t))^{-1},s\in\mathbb{R}.

Also quq\circ u is a nontrivial unipotent one-parameter subgroup of N/GuN/G_{u}. Let β=limtβ(t)\beta_{\infty}=\lim_{t\to\infty}\beta(t). Then, for all ss\in\mathbb{R},

q(β1u(s)β)=limtq(φ(hs,r(t)))q(φ(t))1,q(\beta_{\infty}^{-1}u(s)\beta_{\infty})=\lim_{t\to\infty}q(\varphi(h_{s,r}(t)))q(\varphi(t))^{-1},

and since φGN\varphi\subset G\subset N, we get that {q(β1u(s)β):s}\{q(\beta_{\infty}^{-1}u(s)\beta_{\infty}):s\in\mathbb{R}\} is a nontrivial one-parameter subgroup of G/GuG/G_{u}. It then follows that {β1u(s)β:s}\{\beta_{\infty}^{-1}u(s)\beta_{\infty}:s\in\mathbb{R}\} is a nontrivial one-parameter unipotent subgroup contained in GG not contained in GuG_{u}, a contradiction. Finally, since φ\varphi is bounded modulu GuG_{u}, and since zcl(Gu)\text{zcl}(G_{u}) is observable, we get that HφGuH_{\varphi}\subseteq G_{u} by the minimality of the hull, Theorem 2.4,(1). ∎

6. Linearization

We begin with preliminaries needed for the linearization technique. Suppose that ΓSL(n,)\Gamma\leq\text{SL}(n,\mathbb{R}) is a discrete subgroup, and let μ\mu be a probability measure on X:=SL(n,)/ΓX:=\text{SL}(n,\mathbb{R})/\Gamma. Suppose that there exists a subgroup WGW\leq G of positive dimension which is generated by one-parameter unipotent subgroups contained in WW such that μ\mu is invariant under the WW-left translates. By the celebrated Ratner’s theorem (see [Rat91]) every WW-ergodic component of μ\mu is homogeneous.

Definition 6.1.

Let \mathcal{H} be the class of all closed connected subgroups HH of SL(n,)\text{SL}(n,\mathbb{R}) such that H/HΓH/H\cap\Gamma admits an HH-invariant probability measure and the closed subgroup HuHH_{u}\leq H which is generated by all unipotent one-parameter subgroups of HH acts ergodically on H/HΓH/H\cap\Gamma with respect to the HH-invariant probability measure.

We note the following results:

  • [Rat91, Theorem 1.1]: The collection \mathcal{H} is countable.

For HH\in\mathcal{H}, define

N(H,W)=\displaystyle N(H,W)= {gSL(n,):WgHg1}, and\displaystyle\{g\in\text{SL}(n,\mathbb{R}):W\subset gHg^{-1}\},\text{ and}
S(H,W)=\displaystyle S(H,W)= H,HHN(H,W).\displaystyle\bigcup_{H^{\prime}\in\mathcal{H},H^{\prime}\subsetneq H}N(H^{\prime},W).

In what follows, we denote by π:SL(n,)X\pi:\text{SL}(n,\mathbb{R})\to X the canonical projection. Consider

(6.1) TH(W)=π(N(H,W)S(H,W))=π(N(H,W))π(S(H,W)),T_{H}(W)=\pi(N(H,W)\smallsetminus S(H,W))=\pi(N(H,W))\smallsetminus\pi(S(H,W)),

for the second equality see [MS95, Lemma 2.4].

Theorem 6.2.

[MS95, Theorem 2.2] Let ν\nu be a WW-invariant Borel probability measure on XX. For HH\in\mathcal{H}, let νH\nu_{H} denote the restriction of ν\nu on TH(W)T_{H}(W). Then ν\nu decomposes as:

ν=HνH,\nu=\sum_{H\in\mathcal{H}^{*}}\nu_{H},

where \mathcal{H}^{*}\subset\mathcal{H} is a countable set of subgroups, each is a representative of a distinct Γ\Gamma-conjugacy class.

Moreover, for HH\in\mathcal{H} it holds that νH\nu_{H} is WW-invariant, and any WW-ergodic component of νH\nu_{H} is the unique gHg1gHg^{-1}-invariant probability measure on the closed orbit gHΓ/ΓXgH\Gamma/\Gamma\subset X, for some gN(H,W)S(H,W)g\in N(H,W)\smallsetminus S(H,W).

6.1. o-minimal curves in representations – a dichotomy theorem

For any dd\in\mathbb{N}, consider the action of SL(n,)\text{SL}(n,\mathbb{R}) on

(6.2) Vd=d𝔰𝔩(n,),V_{d}=\bigwedge^{d}\mathfrak{sl}(n,\mathbb{R}),

induced by the Adjoint representation of SL(n,)\text{SL}(n,\mathbb{R}) on its Lie algebra 𝔰𝔩(n,)\mathfrak{sl}(n,\mathbb{R}); that is, g(i=1dxi):=i=1dAdg(xi)g\cdot(\bigwedge_{i=1}^{d}x_{i}):=\bigwedge_{i=1}^{d}\text{Ad}_{g}(x_{i}) and extended linearly. Let HH\in\mathcal{H}, and let 𝔥\mathfrak{h} be the Lie algebra of HH and dH=dim𝔥d_{H}=\dim\mathfrak{h}. Fix a vector pHdH𝔥{0}p_{H}\in\bigwedge^{d_{H}}\mathfrak{h}\setminus\{0\} .

Theorem 6.3.

Let HH\in\mathcal{H}. Then:

  1. (1)

    [DM93, Proof of Proposition 3.2] Let 𝔴\mathfrak{w} be the Lie algebra of WW, and consider the variety

    (6.3) AH:={vVdH:vw=0VdH+1,w𝔴}.A_{H}:=\{v\in V_{d_{H}}:v\wedge w=0\in V_{d_{H}+1},\forall w\in\mathfrak{w}\}.

    Then, g.pHAHg.p_{H}\in A_{H} if and only if gN(H,W)g\in N(H,W). In other words, let ηH:SL(n,)VdH\eta_{H}:\text{SL}(n,\mathbb{R})\to V_{d_{H}} denote the orbit map

    (6.4) ηH(g)=g.pH,gSL(n,),\eta_{H}(g)=g.p_{H},\,\forall g\in\text{SL}(n,\mathbb{R}),

    then ηH1(AH)=N(H,W).\eta_{H}^{-1}(A_{H})=N(H,W).

  2. (2)

    [DM93, Theorem 3.4] Γ.pH\Gamma.p_{H} is discrete in VdHV_{d_{H}}.

We present the following variant of [MS95, Proposition 3.4].

Theorem 6.4.

Let φ:[0,)SL(n,)\varphi:[0,\infty)\to\text{SL}(n,\mathbb{R}) be a continuous, unbounded, non-contracting curve that is definable in a polynomially bounded o-minimal structure. Let β:[0,)SL(n,)\beta:[0,\infty)\to\text{SL}(n,\mathbb{R}) be a correcting curve of φ\varphi and let HφH_{\varphi} be the hull of φ\varphi such that β(t)φ(t)Hφ\beta(t)\varphi(t)\subset H_{\varphi} for all tt, see Definition 2.5. Fix HH\in\mathcal{H}. Let K1AH\emph{\text{K}}_{1}\subset A_{H} be a compact set, and let 0<ϵ<10<\epsilon<1. Then there exists a closed set 𝒮π(S(H,W))\mathcal{S}\subset\pi(S(H,W)) such that the following holds.

For a given compact set KX𝒮K\subset X\smallsetminus\mathcal{S}, there exists a neighborhood Ψ\Psi of K1\emph{\text{K}}_{1} in VdHV_{d_{H}} and such that for any xXx\in X, at least one of the following is satisfied:

  1. (1)

    There exists an wπ1(x).pHΨw\in\pi^{-1}(x).p_{H}\cap\Psi such that

    (6.5) Hφ.w=w.H_{\varphi}.w=w.
  2. (2)

    For all sufficiently large T>0T>0 (depending on xx),

    (6.6) |{t[0,T]:β(t)φ(t).xKπ(ηH1(Ψ¯))}|ϵT.|\{t\in[0,T]:\beta(t)\varphi(t).x\in K\cap\pi(\eta_{H}^{-1}(\overline{\Psi}))\}|\leq\epsilon T.

    Here, ηH\eta_{H} is the orbit map (6.4).

First, recall some observations and results that will be used in the proof. We let N(H)N(H) be the normalizer of HH in SL(n,)\text{SL}(n,\mathbb{R}), and denote

NΓ:=N(H)Γ.N_{\Gamma}:=N(H)\cap\Gamma.

We recall that for γNΓ\gamma\in N_{\Gamma} it holds that

(6.7) det(Adγ|𝔥){±1},\det(\text{Ad}_{\gamma}|_{\mathfrak{h}})\in\{\pm 1\},

see [DM93, Lemma 3.1].

In view of (6.7), either NΓ.pH={pH}N_{\Gamma}.p_{H}=\{p_{H}\} or NΓ.pH={±pH}N_{\Gamma}.p_{H}=\{\pm p_{H}\}. If the latter case is true, we will denote by V~H:=VdH/{±1}\tilde{V}_{H}:=V_{d_{H}}/\{\pm 1\}, the space VdHV_{d_{H}} modulo the equivalence relation defined by identifying vVdHv\in V_{d_{H}} with v-v. Otherwise, namely if NΓ.pH={pH}N_{\Gamma}.p_{H}=\{p_{H}\}, then we denote V~dH:=VdH\tilde{V}_{d_{H}}:=V_{d_{H}}.

Proposition 6.5.

[MS95, Proposition 3.2] Let DD be a compact subset of AHA_{H}, and consider

(6.8) S(D)={gηH1(D):gγηH1(D) for some γΓNΓ}SL(n,).S(D)=\{g\in\eta_{H}^{-1}(D):g\gamma\in\eta_{H}^{-1}(D)\text{ for some }\gamma\in\Gamma\smallsetminus N_{\Gamma}\}\subset\text{SL}(n,\mathbb{R}).

Then:

  1. (1)

    S(D)S(H,W)S(D)\subset S(H,W).

  2. (2)

    π(S(D))\pi(S(D)) is closed in XX.

  3. (3)

    Suppose that KXπ(S(D))K\subset X\smallsetminus\pi(S(D)) is compact. Then, there is a compact neighborhood ΦVdH\Phi\subseteq V_{d_{H}} of DD such that for every yπ(ηH1(Φ))Ky\in\pi(\eta_{H}^{-1}({\Phi}))\cap K, the set (π1(y).p~H)Φ~(\pi^{-1}(y).\tilde{p}_{H})\cap\tilde{{\Phi}} consists of a single element. Here Φ~\tilde{{\Phi}} are the images of pHp_{H} and Φ{\Phi} in V~dH\tilde{V}_{d_{H}}.

Proof of Theorem 6.4.

We assume for simplicity that NΓ.pH={pH}N_{\Gamma}.p_{H}=\{p_{H}\}, namely V~dH=VdH\tilde{V}_{d_{H}}=V_{d_{H}}. The case for which NΓ.pH={±pH}N_{\Gamma}.p_{H}=\{\pm p_{H}\} is treated in essentially the same way and is left for the readers. Let φ:[0,)SL(n,)\varphi:[0,\infty)\to\text{SL}(n,\mathbb{R}) be an unbounded, continuous non-contracting curve definable in a polynomially bounded o-minimal structure. Fix HH\in\mathcal{H}. Let K1AH\text{K}_{1}\subset A_{H} be a compact subset and 0<ϵ<10<\epsilon<1 be given. We will apply Proposition 3.10 with ψ(t)\psi(t) being the image of β(t)φ(t)\beta(t)\varphi(t) in the representation VdH{V}_{d_{H}}. Let K2AH\text{K}_{2}\subset A_{H} be a compact set given by Proposition 3.10 with K1K2\text{K}_{1}\subset\text{K}_{2}.

In the notations of Proposition 6.5, we take the compact set D:=K2D:=\text{K}_{2} and let 𝒮:=π(S(K2)).\mathcal{S}:=\pi(S(\text{K}_{2})). Let KX𝒮K\subset X\smallsetminus\mathcal{S} be a compact set given in the hypothesis of the proposition. Let ΦVdH\Phi\subset{V}_{d_{H}} be a compact neighborhood of K2\text{K}_{2} which satisfies the outcome of Proposition 6.5,(3). As a consequence, for any t0t\geq 0, w1,w2π1(x).pHw_{1},w_{2}\in\pi^{-1}(x).p_{H}, and gSL(n,)g\in\text{SL}(n,\mathbb{R}), if gxKgx\in K and β(t)φ(t).wiΦ\beta(t)\varphi(t).w_{i}\in\Phi for i=1,2i=1,2, then w1=w2w_{1}=w_{2}; let us call this the injective property of KK in Φ\Phi.

We now choose an open neighborhood Ψ\Psi of K1\text{K}_{1} in VdHV_{d_{H}}, with Ψ¯Φ̊\overline{\Psi}\subset\mathring{\Phi}, as in Proposition 3.10, where Φ̊\mathring{\Phi} denotes the interior of Φ\Phi.

Case 1There exists wπ1(x).pHΨ¯w\in\pi^{-1}(x).p_{H}\cap\overline{\Psi} such that Hφ.w=wH_{\varphi}.w=w for all t0t\geq 0.

Case 2For all wπ1(x).pHΨ¯w\in\pi^{-1}(x).p_{H}\cap\overline{\Psi} it holds that limtβ(t)φ(t).w=\lim_{t\to\infty}\beta(t)\varphi(t).w=\infty. By Theorem 2.4,(3), the two mutually exclusive cases exhaust all possibilities. To complete the proof, we now assume that Case 1 does not occur and Case 2 occurs. We want to show that (6.6) holds for all sufficiently large TT.

We take T0>0T^{\prime}_{0}>0 to satisfy the outcomes of Propositions 1.15 and 3.10, and consider

(6.9) J={tT0:β(t)φ(t)xπ(ηH1(Ψ¯))K}.J=\{t\geq T^{\prime}_{0}:\beta(t)\varphi(t)x\in\pi(\eta^{-1}_{H}(\overline{\Psi}))\cap K\}.

Then, to prove (6.6), we only need to show that

(6.10) |J[T0,T]|ϵT, for all sufficiently large TT0.\lvert J\cap[T_{0}^{\prime},T]\rvert\leq\epsilon T\text{, for all sufficiently large $T\geq T_{0}^{\prime}$.}

For every tJt\in J, there exists wπ1(x).pHw\in\pi^{-1}(x).p_{H} such that β(t)φ(t).wΨ¯Φ\beta(t)\varphi(t).w\in\overline{\Psi}\subset\Phi and we have β(t)φ(t)xK\beta(t)\varphi(t)x\in K. By the injective property of KK in Φ\Phi, we get that such a ww is unique, and denote it by w(t)w(t). We note that for any tJt\in J, limsβ(s)φ(s).w(t)=\lim_{s\to\infty}\beta(s)\varphi(s).w(t)=\infty, because otherwise w(t)w(t) is fixed by HφH_{\varphi} due to (3) of Theorem 2.4, and hence w(t)Ψ¯π1(x).pHw(t)\in\overline{\Psi}\cap\pi^{-1}(x).p_{H}, contradicting our assumption that Case 1 does not occur.

Therefore, given tJt\in J and its corresponding w(t)π1(x).pHw(t)\in\pi^{-1}(x).p_{H}, there exists the largest closed interval I(w(t)):=[t,t+][T0,T]I(w(t)):=[t^{-},t^{+}]\subset[T^{\prime}_{0},T] containing tt such that

  1. (1)

    β(s)φ(s).w(t)Φ\beta(s)\varphi(s).w(t)\in\Phi for all sI(w(t))s\in I(w(t)),

  2. (2)

    tJt^{-}\in J, and

  3. (3)

    β(t+)φ(t+).w(t)ΦΦ̊\beta(t^{+})\varphi(t^{+}).w(t)\in\Phi\smallsetminus\mathring{\Phi} or t+=Tt^{+}=T.

We note that I(w(t))I(w(t)) is uniquely determined by the tJt\in J. For any sJI(w(t))s\in J\cap I(w(t)), we have β(s)φ(s)xK\beta(s)\varphi(s)x\in K and β(s)φ(s)w(t)Φ\beta(s)\varphi(s)\cdot w(t)\in\Phi, so by the injectivity of KK in Φ\Phi, we have w(s)=w(t)w(s)=w(t). Therefore, by (6.9),

JI(w(t))\displaystyle J\cap I(w(t)) ={sI(w(t))J:β(s)φ(s)w(s)Ψ}\displaystyle=\{s\in I(w(t))\cap J:\beta(s)\varphi(s)\cdot w(s)\in\Psi\}
={sI(w(t))J:β(s)φ(s)w(t)Ψ}\displaystyle=\{s\in I(w(t))\cap J:\beta(s)\varphi(s)\cdot w(t)\in\Psi\}
(6.11) {sI(w(t)):β(s)φ(s)w(t)Ψ}.\displaystyle\subseteq\{s\in I(w(t)):\beta(s)\varphi(s)\cdot w(t)\in\Psi\}.

We claim that for any t1,t2Jt_{1},t_{2}\in J and their corresponding w(t1),w(t2)π1(x).pHw(t_{1}),w(t_{2})\in\pi^{-1}(x).p_{H}, we have

(6.12) either I(w(t1))=I(w(t2)) or I(w(t1))I(w(t2))=.\text{either }I(w(t_{1}))=I(w(t_{2}))\text{ or }I(w(t_{1}))\cap I(w(t_{2}))=\varnothing.

To verify this claim, suppose that I(w(t1))I(w(t2))I(w(t_{1}))\cap I(w(t_{2}))\neq\varnothing. Then either t1I(w(t2))t_{1}^{-}\in I(w(t_{2})) or t2I(w(t1))t_{2}^{-}\in I(w(t_{1})). Since t1,t2Jt_{1}^{-},t_{2}^{-}\in J, we have JI(w(t1))I(w(t2))J\cap I(w(t_{1}))\cap I(w(t_{2}))\neq\varnothing. So, by our earlier observation, w(t1)=w(t2)w(t_{1})=w(t_{2}). This proves the claim.

To prove (6.10), we only need to consider the case of JJ being unbounded, and J[T0,T]J\cap[T_{0}^{\prime},T]\neq\varnothing. Let t1=min(J[T0,T])t_{1}^{-}=\min(J\cap[T_{0}^{\prime},T]) and w1=w(t1)=[t1,t1+]w_{1}=w(t_{1}^{-})=[t_{1}^{-},t_{1}^{+}]. We inductively define w1,,wl(T)w_{1},\ldots,w_{l(T)} as follows: For any i1i\geq 1 such that wiw_{i} is defined, if J(ti+,T]J\cap(t_{i}^{+},T]\neq\varnothing, let ti+1=inf(J(ti+,T])t_{i+1}^{-}=\inf(J\cap(t_{i}^{+},T]), and wi+1:=w(ti+1)w_{i+1}:=w(t_{i+1}^{-}), and I(wi+1)=[ti+1,ti+1+]I(w_{i+1})=[t_{i+1}^{-},t_{i+1}^{+}]. Otherwise, we let l(T)=il(T)=i and stop. Thus,

(6.13) J[T0,T]I(w1)I(wl(T)).J\cap[T_{0}^{\prime},T]\subseteq I(w_{1})\sqcup...\sqcup I(w_{l(T)}).

Here among {w1,,wl(T)}π1(x).pH\{w_{1},\ldots,w_{l(T)}\}\subset\pi^{-1}(x).p_{H}, each wiw_{i} appears at most finitely many times, because β(s)φ(s)wi\beta(s)\varphi(s)w_{i}\to\infty as ss\to\infty. Since π1(x).pH\pi^{-1}(x).p_{H} is a discrete subset of VdHV_{d_{H}}, and since and JJ is unbounded, wl(T)\lVert w_{l(T)}\rVert\to\infty as TT\to\infty.

So, by (6.11) and (6.13),

(6.14) |J[T0,T]|k=1l(T)|{tI(wk):β(t)φ(t).wkΨ}|.\lvert J\cap[T_{0}^{\prime},T]\rvert\leq\sum_{k=1}^{l(T)}\left|\{t\in I(w_{k}):\beta(t)\varphi(t).w_{k}\in\Psi\}\right|.

Let 1il(T)11\leq i\leq l(T)-1. Then ti+<tlT<Tt_{i}^{+}<t_{l_{T}}^{-}<T. So, by the defining property (3) of I(wi)=[ti,ti+]I(w_{i})=[t_{i}^{-},t_{i}^{+}], we have β(ti+)φ(ti+)wiΦΦ̊\beta(t_{i}^{+})\varphi(t_{i}^{+})\cdot w_{i}\in\Phi\setminus\mathring{\Phi}. Now, by Proposition 3.10 we have that

|{tI(wk):β(t)φ(t).wkΨ}|ϵ|I(wk)|.\left|\{t\in I(w_{k}):\beta(t)\varphi(t).w_{k}\in\Psi\}\right|\leq\epsilon|I(w_{k})|.

Therefore,

(6.15) k=1l(T)1|{tI(wk):β(t)φ(t).wkΨ}|ϵ(TT0).\sum_{k=1}^{l(T)-1}\left|\{t\in I(w_{k}):\beta(t)\varphi(t).w_{k}\in\Psi\}\right|\leq\epsilon(T-T_{0}).

Now, it is possible that β(tl(T)+)φ(tl(T)+)wl(T)Φ̊\beta(t_{l(T)}^{+})\varphi(t_{l(T)}^{+})\cdot w_{l(T)}\in\mathring{\Phi} and tl(T)+=Tt_{l(T)}^{+}=T. So, we will estimate the l(T)l(T)-th term differently. By Proposition 1.15, and recalling (1.6), given ϵ>0\epsilon>0 we can pick M0>0M_{0}>0 such for any wVdHw\in V_{d_{H}}, if β(T0)φ(T0).wM0\lVert\beta(T^{\prime}_{0})\varphi(T^{\prime}_{0}).w\rVert\geq M_{0}, then

|{t[T0,T]:β(t)φ(t).wΨ}|ϵ(TT0).\left|\{t\in[T_{0}^{\prime},T]:\beta(t)\varphi(t).w\in\Psi\}\right|\leq\epsilon(T-T_{0}^{\prime}).

Since, wl(T)\lVert w_{l(T)}\rVert\to\infty as TT\to\infty, we choose TxT0T_{x}\geq T_{0}^{\prime}, such that for all TTxT\geq T_{x}, β(T0)φ(T0).wl(T)M0\lVert\beta(T^{\prime}_{0})\varphi(T^{\prime}_{0}).w_{l(T)}\rVert\geq M_{0}. Then, for any TTxT\geq T_{x}, we have

(6.16) |{tI(wl(T)):β(t)φ(t).wl(T)Ψ}|ϵ(TT0).\left|\{t\in I(w_{l(T)}):\beta(t)\varphi(t).w_{l(T)}\in\Psi\}\right|\leq\epsilon(T-T_{0}^{\prime}).

Thus, by combining (6.14), (6.15) and (6.16), for all TTxT\geq T_{x} we get

|J[T0,T]|2ϵ(TT0).\lvert J\cap[T_{0}^{\prime},T]\rvert\leq 2\epsilon(T-T_{0}^{\prime}).

Therefore, (6.10) follows for 3ϵ3\epsilon in place of ϵ\epsilon. ∎

6.2. Proving Theorem 1.8

Let G:=𝐆()G:=\mathbf{G}(\mathbb{R}). Suppose that 𝒢G\mathcal{G}\leq G is a connected, closed Lie subgroup with a lattice Γ𝒢\Gamma\leq\mathcal{G}. Suppose that φ:[0,)G\varphi:[0,\infty)\to G be a continuous, unbounded, non-contracting curve for GG definable in a polynomially bounded o-minimal structure. Suppose that φ([0,))𝒢\varphi\left([0,\infty)\right)\subset\mathcal{G}. Let HφH_{\varphi} be the hull of φ\varphi in GG and let β\beta be a definable correcting curve in GG, so that βφHφ\beta\varphi\subset H_{\varphi}, see Definition 2.5. Then, by Proposition 5.13, Hφ𝒢H_{\varphi}\subset\mathcal{G} and β([0,))𝒢\beta([0,\infty))\subset\mathcal{G}. We choose an injective algebraic map ι:GSL(n,)\iota:G\hookrightarrow\text{SL}(n,\mathbb{R}) for some n2n\geq 2. Note that ι(Γ)\iota(\Gamma) is a discrete subgroup of SL(n,)\text{SL}(n,\mathbb{R}). Importantly, ιφ\iota\circ\varphi is non-contracting in SL(n,)\text{SL}(n,\mathbb{R}) by Lemma 2.15, and ι(Hφ)=Hιφ\iota(H_{\varphi})=H_{\iota\circ\varphi} by Lemma 2.8. By abuse of notations, in the following we replace φ\varphi with ιφ\iota\circ\varphi, Γ\Gamma with ι(Γ)\iota(\Gamma) and GG with ι(G)\iota(G).

Let x0𝒢/ΓSL(n,)/Γx_{0}\in\mathcal{G}/\Gamma\hookrightarrow\text{SL}(n,\mathbb{R})/\Gamma, and fix a limiting measure μ\mu of μT,βφ,x0\mu_{T,\beta\varphi,x_{0}} (as in (1.1)), with respect to the weak-\ast topology on the space of probability measures on SL(n,)/Γ\text{SL}(n,\mathbb{R})/\Gamma, TT\to\infty. We fix a sequence {Ti}\{T_{i}\} such that TiT_{i}\to\infty and

(6.17) limiμTi,βφ,x0=μ.\lim_{i\to\infty}\mu_{T_{i},\beta\varphi,x_{0}}=\mu.

By Theorem 4.5, μ\mu is a probability measure. We let WW be the subgroup of SL(n,)\text{SL}(n,\mathbb{R}) generated by all unipotent one-parameter subgroups under which the limiting measure μ\mu is invariant. By Proposition 5.1, WW has a positive dimension. Using Theorem 6.2, there exists HH\in\mathcal{H} (of smallest dimension) such that

(6.18) μ(π(N(H,W)))>0 and μ(π(S(H,W)))=0.\displaystyle\mu(\pi(N(H,W)))>0\text{ and }\mu(\pi(S(H,W)))=0.

Define

(6.19) N1(H):=ηH1(pH)={gSL(n,):g.pH=pH}.\displaystyle N^{1}(H):=\eta_{H}^{-1}(p_{H})=\{g\in\text{SL}(n,\mathbb{R}):g.p_{H}=p_{H}\}.

and recall

  • [DM93, Theorem 3.4]: N1(H)Γ/ΓXN^{1}(H)\Gamma/\Gamma\subset X is closed.

Lemma 6.6.

Fix HH\in\mathcal{H} such that (6.18) holds. Then there exists an g0SL(n,)g_{0}\in\text{SL}(n,\mathbb{R}) such that x0=π(g0)x_{0}=\pi(g_{0}) and

(6.20) g01β(t)φ(t)g0N1(H),t0,g_{0}^{-1}\beta(t)\varphi(t)g_{0}\subset N^{1}(H),~{}\forall t\geq 0,

Importantly,

(6.21) ν:=g01μ,\nu:=g_{0}^{-1}\mu,

is supported on π(N1(H))\pi(N^{1}(H)), ν\nu is HH-invariant, and the largest group generated by unipotent one-parameter subgroups preserving ν\nu is HuH_{u}.

Proof.

In view of (6.1), due to (6.18), there is a compact subset CN(H,W)S(H,W)C\subset N(H,W)\smallsetminus S(H,W) such that μ(π(C))=α\mu(\pi(C))=\alpha for some α>0\alpha>0, and π(C)π(S(H,W))=\pi(C)\cap\pi(S(H,W))=\varnothing. We will now apply Theorem 6.4. We consider the compact subset K1AH\text{K}_{1}\subset A_{H} given by K1:=C.pH\text{K}_{1}:=C.p_{H} and fix ϵ:=α2\epsilon:=\frac{\alpha}{2}. Let 𝒮π(S(H,W))\mathcal{S}\subset\pi(S(H,W)) be the closed subset given in Theorem 6.4. Now let Ψ\Psi be an arbitrary open neighborhood of K1\text{K}_{1}, and let KX𝒮K\subseteq X\smallsetminus\mathcal{S} be an arbitrary open neighborhood of π(C)\pi(C), which exists as π(C)\pi(C) is compact and 𝒮\mathcal{S} is closed. Then, Kπ(ηH1(Ψ))K\cap\pi(\eta_{H}^{-1}(\Psi)) is an open neighborhood of π(C)\pi(C). Since we assume (6.17), for all ii large enough we have

|{t[0,Ti]:β(t)φ(t).xKπ(ηH1(Ψ))}|ϵTi.|\{t\in[0,T_{i}]:\beta(t)\varphi(t).x\in K\cap\pi(\eta_{H}^{-1}(\Psi))\}|\geq\epsilon T_{i}.

Namely, condition 6.4,(2) fails, and so the first outcome of Theorem 6.4,(1) must hold. That is, there is a g0π1(x)g_{0}\in\pi^{-1}(x) such that g0.pHΨg_{0}.p_{H}\in\Psi and

(6.22) β(t)φ(t)g0.pH=g0.pH, for t0.\beta(t)\varphi(t)g_{0}.p_{H}=g_{0}.p_{H},\text{ for }t\geq 0.

Namely, β(t)φ(t)g0g0N1(H)\beta(t)\varphi(t)g_{0}\in g_{0}N^{1}(H) for all tT0t\geq T_{0}. Since π(N1(H))\pi(N^{1}(H)) is closed, we obtain that μ\mu is supported in π(g0N1(H))\pi(g_{0}N^{1}(H)).

Since π1(x).pH\pi^{-1}(x).p_{H} is discrete, and since we may choose the K1\text{K}_{1}-neighborhood Ψ\Psi arbitrarily, we obtain that g0.pHAHg_{0}.p_{H}\in A_{H}. Then, by Theorem 6.3,(1), we conclude that g0N(H,W)g_{0}\in N(H,W). Then g0N1(H)N(H,W)g_{0}N^{1}(H)\subseteq N(H,W).

In particular, we conclude that μ(π(g0N1(H)S(H,W))=1\mu(\pi(g_{0}N^{1}(H)\smallsetminus S(H,W))=1. By [MS95, Lemma 2.4], for each g0ng0N1(H)S(H,W)g_{0}n\in g_{0}N^{1}(H)\smallsetminus S(H,W), the group

H=(g0n)H(g0n)1=g0Hg01H^{\prime}=(g_{0}n)H(g_{0}n)^{-1}=g_{0}Hg_{0}^{-1}

is the smallest group containing WW such that the orbit Hπ(g0n)H^{\prime}\pi(g_{0}n) is closed. Then, by [Sha91, Theorem 2.3], WW acts ergodically on Hπ(g0n)H^{\prime}\pi(g_{0}n) with respect to g0μHg_{0}\mu_{H}, where μH\mu_{H} is the HH invariant measure on H/HΓHΓ/ΓH/H\cap\Gamma\cong H\Gamma/\Gamma. Namely, μ\mu-almost every WW-ergodic component of μ\mu is g0Hg01g_{0}Hg_{0}^{-1}-invariant. By performing ergodic decomposition on μ\mu, we obtain that μ\mu is g0Hg01g_{0}Hg_{0}^{-1}-invariant. Then, ν=g01μ\nu=g_{0}^{-1}\mu is HH-invariant. So HuH_{u} preserves ν\nu. Since g01Wg0Hg_{0}^{-1}Wg_{0}\leq H, we get Hu=g01Wg0H_{u}=g_{0}^{-1}Wg_{0} by the maximality of WW.

We will use the following notations: ψ(t):=g01β(t)φ(t)g0\psi(t):=g_{0}^{-1}\beta(t)\varphi(t)g_{0}, for t0t\geq 0, N:=N1(H)N:=N^{1}(H), Δ:=N1(H)Γ\Delta:=N^{1}(H)\cap\Gamma, and q:NN/Hq:N\to N/H denotes the natural quotient map.

Consider N¯:=q(N)\bar{N}:=q(N), Δ¯:=q(Δ)\bar{\Delta}:=q(\Delta). Since HΔ=ΔHH\Delta=\Delta H is closed, we have that Δ¯N¯\bar{\Delta}\leq\bar{N} is discrete (since ΔH\Delta H is closed, then ΔH/HΔ/ΔH\Delta H/H\cong\Delta/\Delta\cap H). We let q¯:N/ΔN¯/Δ¯\bar{q}:N/\Delta\to\bar{N}/\bar{\Delta} be the natural map. Notice that the fibers of q¯\bar{q} are HH-orbits. Namely,

q¯1(n¯Δ¯))=nHΔ/Δ=HnΔ/Δ.\bar{q}^{-1}(\bar{n}\bar{\Delta}))=nH\Delta/\Delta=Hn\Delta/\Delta.

Since μ\mu is a probability measure on N/ΔN/\Delta, it follows that ν¯:=q¯ν\bar{\nu}:=\bar{q}_{*}\nu is a probability measure on N¯/Δ¯\bar{N}/\bar{\Delta}.

Lemma 6.7.

Suppose that ψ(t)\psi(t) is unbounded in N/HuN/H_{u}. Then there exists an unipotent one-parameter subgroup UNU\leq N such that UU is not a subgroup of HH and ν¯\bar{\nu} is invariant under q(U)q(U) and q(U)q(U) is a (non-trivial) one-parameter subgroup.

Proof.

We note that the measure push-forward map qq_{*} is continuous, and in particular,

ν¯(f)=limi1Ti0Tif(q(ψ(t))Δ¯)𝑑t,fCc(N¯/Δ¯).\bar{\nu}(f)=\lim_{i\to\infty}\frac{1}{T_{i}}\int_{0}^{T_{i}}f(q(\psi(t))\bar{\Delta})dt,\forall f\in C_{c}(\bar{N}/\bar{\Delta}).

Thus, we may conclude the statement by applying Proposition 5.12, and Lemma 5.5 (see the proof of Lemma 5.6. The only difference here is that N/HN/H is not a subgroup of GL(n,)\text{GL}(n,\mathbb{R})). ∎

We now note the following, which is obtained by the ergodic decomposition for the HH-action on N/ΔN/\Delta (see [EW11, Theorem 8.20])

Lemma 6.8.

Let d(hΔ)d(h\Delta) denote the volume element of the HH-invariant probability measure on HΔ/ΔH\Delta/\Delta. For fCc(N/Δ)f\in C_{c}(N/\Delta) let:

(6.23) P(f)(q(n)Δ¯):=HΔ/Δf(nhΔ)d(hΔ).P(f)(q(n)\bar{\Delta}):=\int_{H\Delta/\Delta}f(nh\Delta)d(h\Delta).

Then,

N/Δf(nΔ)𝑑ν(nΔ)=N¯/Δ¯P(f)(q(n)Δ¯)𝑑ν¯(q(n)Δ¯).\displaystyle\int_{N/\Delta}f(n\Delta)d\nu(n\Delta)=\int_{\bar{N}/\bar{\Delta}}P(f)(q(n)\bar{\Delta})d\bar{\nu}(q(n)\bar{\Delta}).

We now have the following key observation.

Corollary 6.9.

It holds that ψ(t)Hu\psi(t)\in H_{u} for all tT0t\geq T_{0}, and ν\nu is the HH-invariant probability on Hπ(e)H\pi(e), where π(e)\pi(e) is the identity coset in X=SL(n,)/ΓX=\text{SL}(n,\mathbb{R})/\Gamma. Moreover, HH is the smallest subgroup such that g01Hφg0Hg_{0}^{-1}H_{\varphi}g_{0}\leq H and such that the orbit Hπ(e)H\pi(e) is closed.

Proof.

We first show that ψ(t)\psi(t) is bounded in N/HuN/H_{u}. In fact, if ψ(t)\psi(t) is unbounded in N/HuN/H_{u}, then by Lemma 6.7, ν¯\bar{\nu} is invariant under q(U)q(U), where UU is a one-parameter unipotent subgroup and q(U)q(U) is a non-trivial one-parameter subgroup of N¯=N/H\bar{N}=N/H. It then follows from Lemma 6.8 than ν\nu is invariant by UU and UU is not a subgroup of HH. Since ν\nu was invariant by HH as well, we get that ν\nu is invariant by a group generated by one-parameter subgroups which strictly includes HuH_{u}. This is a contradiction since HuH_{u} is a maximal group with this property.

Since the image of ψ(t)=g01β(t)φ(t)g0\psi(t)=g_{0}^{-1}\beta(t)\varphi(t)g_{0} is boundned in SL(n,)/Hu\text{SL}(n,\mathbb{R})/H_{u}, we get that the image of β(t)φ(t)\beta(t)\varphi(t) is bounded in SL(n,)/g0Hug01\text{SL}(n,\mathbb{R})/g_{0}H_{u}g_{0}^{-1}. Thus, since zcl(Hu)\text{zcl}(H_{u}) is observable (see [Gro97, Corollary 2.8]), we conclude by Theorem 2.4 that Hφg0Hug01H_{\varphi}\leq g_{0}H_{u}g_{0}^{-1}, and in particular β(t)φ(t)g0Hug01\beta(t)\varphi(t)\in g_{0}H_{u}g_{0}^{-1} for all tt. Namely, ψ(t)g01Hφg0Hu\psi(t)\in g_{0}^{-1}H_{\varphi}g_{0}\leq H_{u} for all tT0t\geq T_{0}. In particular, ν¯\bar{\nu} is the Dirac measure at the identity coset, and so ν=μH\nu=\mu_{H} by Lemma 6.8. In particular {ψ(t)π(e)}t0\{\psi(t)\pi(e)\}_{t\geq 0} is dense in Hπ(e)H\pi(e). In particular g01Hφg0π(e)g_{0}^{-1}H_{\varphi}g_{0}\pi(e) is dense in Hπ(e)H\pi(e), which finishes the proof.

The proof of the main Theorem 1.8 follows directly from the above corollary.

7. Curves with a largest possible hull

The following result gives a condition for the hull to be semi-simple.

Lemma 7.1.

Suppose that φ:[0,)SL(n,)\varphi:[0,\infty)\to\text{SL}(n,\mathbb{R}) is a continuous curve definable in an o-minimal structure such that φ\varphi has the following divergence property:

(7.1) limtφ(t).(v1vk)=,\lim_{t\to\infty}\varphi(t).(v_{1}\wedge\dots\wedge v_{k})=\infty,

for all 1kn11\leq k\leq n-1 and all linearly independent vectors v1,,vknv_{1},...,v_{k}\in\mathbb{R}^{n}. Then, the observable hull HφH_{\varphi} of φ\varphi is a connected semi-simple with no compact factors, and the action of HφH_{\varphi} on n\mathbb{R}^{n} is irreducible.

Proof.

Since zcl(Hφ)\text{zcl}(H_{\varphi}) is observable, we have a finite dimensional rational representation VV of SL(n,)\text{SL}(n,\mathbb{R}) and a vVv\in V such that zcl(Hφ)\text{zcl}(H_{\varphi}) is the isotropy group of vv. Note that φ(t).v,t0\varphi(t).v,t\geq 0 is bounded since φ\varphi is bounded in SL(n,)/Hφ\text{SL}(n,\mathbb{R})/H_{\varphi}. Since we assume (7.1), it follows by Lemma 2.1 that the orbit SL(n,).v\text{SL}(n,\mathbb{R}).v is Zariski closed. Then, Matsushima criterion [Mat60] implies that zcl(Hφ)\text{zcl}(H_{\varphi}) is reductive. Since HφH_{\varphi} is connected and generated by one-parameter unipotent subgroups, it follows that HφH_{\varphi} is a connected semi-simple group with no compact factors. Finally, we prove that HφH_{\varphi} acts on n\mathbb{R}^{n} irreducibly. Suppose for contradiction that n=V1V2\mathbb{R}^{n}=V_{1}\oplus V_{2} where both V1V_{1} and V2V_{2} are non-trivial sub-spaces invariant under HφH_{\varphi}. Let v1,,vkv_{1},...,v_{k} by a basis for V1V_{1}. Here 1k<n1\leq k<n because V2V_{2} is not trivial. Since semi-simple groups have no non-trivial characters, we obtain that Hφ.(v1vk)=v1vkH_{\varphi}.(v_{1}\wedge\cdots\wedge v_{k})=v_{1}\wedge\cdots\wedge v_{k}. Since φ([0,))\varphi([0,\infty)) is bounded in SL(n,)/Hφ\text{SL}(n,\mathbb{R})/H_{\varphi}, we get φ(t).(v1vk)\varphi(t).(v_{1}\wedge\cdots\wedge v_{k}) is bounded as tt\to\infty. This contradicts (7.1). ∎

Let φ\varphi be the curve given by (1.5). First, we will verify that φ\varphi satisfies the following stronger divergence property.

Lemma 7.2.

For any 1kn1\leq k\leq n, and 0vkn+10\neq v\in\wedge^{k}\mathbb{R}^{n+1}, limtφ(t).v\lim_{t\to\infty}\varphi(t).v\to\infty as tt\to\infty.

Proof.

We denote by e0,e1,,ene_{0},e_{1},...,e_{n} the canonical basis vectors of n+1\mathbb{R}^{n+1}, where eie_{i} denotes the vectors whose coordinates equal 0 besides the (i+1)(i+1)th position in which 11 appears. Fix an integer 1kn1\leq k\leq n and let \mathcal{I} be the collection of all subsets of {0,,n}\{0,\ldots,n\} of cardinality kk. Let φ\varphi be the curve given by (1.5). For any I:={i1,,ik}I:=\{i_{1},\ldots,i_{k}\}\in\mathcal{I}, where i1<i2<<iki_{1}<i_{2}<\cdots<i_{k}, let eI=ei1eike_{I}=e_{i_{1}}\wedge\cdots\wedge e_{i_{k}}. There are two cases for φ(t).eI\varphi(t).e_{I}:

If i1=0i_{1}=0, then

φ(t).eI=\displaystyle\varphi(t).e_{I}= f0(t)e0[fi2(t)e0+hi2(t)1ei2][fik(t)e0+hik(t)1eik]\displaystyle f_{0}(t)e_{0}\wedge[f_{i_{2}}(t)e_{0}+h_{i_{2}}(t)^{-1}e_{i_{2}}]\wedge\cdots\wedge[f_{i_{k}}(t)e_{0}+h_{i_{k}}(t)^{-1}e_{i_{k}}]
(7.2) =\displaystyle= f0(t)(l=2khil(t)1)eI;\displaystyle f_{0}(t)\left(\prod_{l=2}^{k}h_{i_{l}}(t)^{-1}\right)e_{I};

and if i11i_{1}\geq 1, then

φ(t).eI=\displaystyle\varphi(t).e_{I}= [fi1(t)e0+hi1(t)1ei1][fik(t)e0+hik(t)1eik]\displaystyle[f_{i_{1}}(t)e_{0}+h_{i_{1}}(t)^{-1}e_{i_{1}}]\wedge\cdots\wedge[f_{i_{k}}(t)e_{0}+h_{i_{k}}(t)^{-1}e_{i_{k}}]
=\displaystyle= l=1k(1)l1fil(t)(j{1,,k}{l}hij(t)1)e{0}I{il}\displaystyle\sum_{l=1}^{k}(-1)^{l-1}f_{i_{l}}(t)\left(\prod_{j\in\{1,\ldots,k\}\setminus\{l\}}h_{i_{j}}(t)^{-1}\right)e_{\{0\}\cup I\setminus\{i_{l}\}}
(7.3) +j{1,,k}hij(t)1eI.\displaystyle+\prod_{j\in\{1,\ldots,k\}}h_{i_{j}}(t)^{-1}e_{I}.

Let 𝐯kn+1{0}\mathbf{v}\in\wedge^{k}\mathbb{R}^{n+1}\setminus\{0\}. Then 𝐯=IcIeI\mathbf{v}=\sum_{I\in\mathcal{I}}c_{I}e_{I}, where cIc_{I}\in\mathbb{R}. Then,

(7.4) φ(t).𝐯=ICI(t)eI,\varphi(t).\mathbf{v}=\sum_{I\in\mathcal{I}}C_{I}(t)e_{I},

where CI(t)C_{I}(t)\in\mathbb{R}. The coefficients of our interest are CI(t)C_{I}(t) such that 0I0\in I. For any 1i2<<ikn1\leq i_{2}<\cdots<i_{k}\leq n, and J={i2,,ik}J=\{i_{2},\ldots,i_{k}\},

(7.5) C{0}J(t)=(jJhj(t)1)l{0,1,,k}Jfl(t)cJ{l}(1)jl,\displaystyle C_{\{0\}\cup J}(t)=\left(\prod_{j\in J}h_{j}(t)^{-1}\right)\sum_{l\in\{0,1,\ldots,k\}\setminus J}f_{l}(t)c_{J\cup\{l\}}(-1)^{j_{l}},

where jl=min{j{2,,k}:ij>l}j_{l}=\min\{j\in\{2,\ldots,k\}:i_{j}>l\}.

Let 𝐯={I:cI0}\mathcal{I}_{\mathbf{v}}=\{I\in\mathcal{I}:c_{I}\neq 0\}\neq\varnothing, as 𝐯0\mathbf{v}\neq 0. Let i1=min𝐯i_{1}=\min\cup\mathcal{I}_{\mathbf{v}}. We pick I𝐯I\subset\mathcal{I}_{\mathbf{v}} such that iiIi_{i}\in I. Let J=I{i1}J=I\setminus\{i_{1}\}. Then, cJ{l}=0c_{J\cup\{l\}}=0 for every l<i1l<i_{1}. Therefore, by (7.5)

(7.6) C{0}J(t)=(jJhj(t))1(fi1cI+l{i1+1,,k}Jfl(t)cJ{l}(1)jl).C_{\{0\}\cup J}(t)=\left(\prod_{j\in J}h_{j}(t)\right)^{-1}\left(f_{i_{1}}c_{I}+\sum_{l\in\{i_{1}+1,\ldots,k\}\setminus J}f_{l}(t)c_{J\cup\{l\}}(-1)^{j_{l}}\right).

By condition (1), i=1ndeghi=n\sum_{i=1}^{n}\deg h_{i}=n and degh1deghn\deg h_{1}\geq\cdots\geq\deg h_{n}. Therefore

(7.7) i=1jdeghij\sum_{i=1}^{j}\deg h_{i}\geq j

for each 1jn1\leq j\leq n. Since 1kn1\leq k\leq n and deghn>0\deg h_{n}>0, we get

(7.8) deghi2++deghikdegh1++deghk1ndeghn<n,\deg h_{i_{2}}+\cdots+\deg h_{i_{k}}\leq\deg h_{1}+\cdots+\deg h_{k-1}\leq n-\deg h_{n}<n,

and if i11i_{1}\geq 1, then by (7.7),

(7.9) deghi2++deghikn(degh1++deghi1)ni1.\deg h_{i_{2}}+\cdots+\deg h_{i_{k}}\leq n-(\deg h_{1}+\cdots+\deg h_{i_{1}})\leq n-i_{1}.

Since cI0c_{I}\neq 0, by our assumptions (2) and (3),

(7.10) deg(cIfi1+l{i1+1,,k}J(1)jlcJ{l}fl(t))nif i1=0>ni1if i11.\deg\left(c_{I}f_{i_{1}}+\sum_{l\in\{i_{1}+1,\ldots,k\}\setminus J}(-1)^{j_{l}}c_{J\cup\{l\}}f_{l}(t)\right)\begin{array}[]{ll}\geq n&\text{if $i_{1}=0$}\\ >n-{i_{1}}&\text{if $i_{1}\geq 1$}.\end{array}

By combining (7.6), (7.8), (7.9) and (7.10), we conclude that degC{0}J(t)>0\deg C_{\{0\}\cup J}(t)>0. Hence, the conclusion of the lemma follows from (7.4). ∎

Lemma 7.3.

Let {ρ(s):s}\{\rho(s):s\in\mathbb{R}\} be the P.S. group of φ\varphi (Definition 5.3). Then, there exists v=(v1,,vn)n{0}v=(v_{1},\ldots,v_{n})\in\mathbb{R}^{n}\setminus\{0\}, such that

(7.11) ρ(s):=[1sv01]SL(n+1,),s.\rho(s):=\begin{bmatrix}1&sv\\ 0&1\end{bmatrix}\in\text{SL}(n+1,\mathbb{R}),\,\forall s\in\mathbb{R}.
Proof.

Due to Lemma 7.2, by Lemma 5.9, ρ(s)\rho(s) is unipotent. Moreover, there exists r<1r<1 such that for any ss\in\mathbb{R},

(7.12) φ(hr,s(t))φ(t)1ρ(s),\varphi(h_{r,s}(t))\varphi(t)^{-1}\to\rho(s),

where hr,sh_{r,s} is as in (5.1). We note that φ(s)\varphi(s) is contained in the subgroup FF of SL(n+1,)\text{SL}(n+1,\mathbb{R}) whose nonzero entries are only on the diagonal and in the top row. Therefore, by (7.12), ρ(s)\rho(s) is a one-parameter unipotent subgroup of FF. Hence, ρ(s)\rho(s) is of the form given by (7.11). ∎

Proof of Proposition 1.10.

Let {e0,,en}\{e_{0},\ldots,e_{n}\} denote the standard basis of n+1\mathbb{R}^{n+1}. Let (v1,,vn)n(v_{1},\ldots,v_{n})\in\mathbb{R}^{n} be as in Lemma lem:ps group of our main example. Let W1W_{1} denote the ortho-complement of v=i=1nvieiv=\sum_{i=1}^{n}v_{i}e_{i} in the span of {e1,,en}\{e_{1},\ldots,e_{n}\}. Then the fixed points space of ρ(s)\rho(s) in n+1\mathbb{R}^{n+1} is W:=e0+W1W:=\mathbb{R}e_{0}+W_{1}.

Now let β\beta denote the correcting definable curve such that β(t)φ(t)Hφ\beta(t)\varphi(t)\subset H_{\varphi}. Since β(t)β()SL(n+1,)\beta(t)\to\beta(\infty)\in\text{SL}(n+1,\mathbb{R}) as tt\to\infty, by (7.12), we conclude that ρ1(s):=β()ρ(s)β()1Hφ\rho_{1}(s):=\beta(\infty)\rho(s)\beta(\infty)^{-1}\in H_{\varphi} for all ss\in\mathbb{R}. By Lemma 7.2 and Lemma 7.1, HφH_{\varphi} is a semisimple group acting irreducibly on n+1\mathbb{R}^{n+1}. Therefore, by Jacobson-Morosov theorem, there exists a homomorphism η:SL(2,)Hφ\eta:\text{SL}(2,\mathbb{R})\to H_{\varphi} such that η(u(s))=ρ1(s)\eta(u(s))=\rho_{1}(s), where u(s):=[1s01]SL(2,)u(s):=\begin{bmatrix}1&s\\ 0&1\end{bmatrix}\in\text{SL}(2,\mathbb{R}). As noted earlier, the fixed points space of ρ1\rho_{1} on n+1\mathbb{R}^{n+1} is β()W\beta(\infty)W, and it has dimension nn. Expressing the representation of SL(2,)\text{SL}(2,\mathbb{R}) on n+1\mathbb{R}^{n+1} via η\eta as a direct sum of irreducible representations of SL2()\text{SL}_{2}(\mathbb{R}), and noting that the dimension of the u(s)u(s)-fixed subspace is nn, we conclude that n+1=V1V2\mathbb{R}^{n+1}=V_{1}\oplus V_{2}, where V1V_{1} is isomorphic to the standard representation of SL(2,)\text{SL}(2,\mathbb{R}) on 2\mathbb{R}^{2}, and SL(2,)\text{SL}(2,\mathbb{R}) acts trivially on V2V_{2}. Therefore, by [SY24, Theorem A.7], either Hφ=SL(n+1,)H_{\varphi}=\text{SL}(n+1,\mathbb{R}) or n+1=2dn+1=2d for some d2d\geq 2 and HφH_{\varphi} is conjugate to the standard symplectic group SP(2d)\text{SP}(\mathbb{R}^{2d}).

Consider the case when HφSL(n+1,)H_{\varphi}\neq\text{SL}(n+1,\mathbb{R}). Then, there exists a basis w1,w1,w2,w2,,wd,wdw_{1},w^{\prime}_{1},w_{2},w^{\prime}_{2},\ldots,w_{d},w^{\prime}_{d} of 2d\mathbb{R}^{2d} such that HφH_{\varphi} stabilizes

𝐯=w1w1+w2w2++wnwn22d.\mathbf{v}=w_{1}\wedge w^{\prime}_{1}+w_{2}\wedge w^{\prime}_{2}+\cdots+w_{n}\wedge w^{\prime}_{n}\in\wedge^{2}\mathbb{R}^{2d}.

Now, by Lemma 7.2, φ(t)\varphi(t) cannot be bounded in SL(2d,)/Hφ\text{SL}(2d,\mathbb{R})/H_{\varphi}, contradictions the assumption that HφH_{\varphi} is the hull of φ\varphi. Therefore, Hφ=SL(n+1,)H_{\varphi}=\text{SL}(n+1,\mathbb{R}). This completes the proof of Proposition 1.10. ∎

Proof of Corollary 1.11.

If GG is a linear Lie group, the result follows immediately from Proposition 1.10 and Theorem 1.8. If GG is not a linear Lie group, for our proof to work, it is sufficient that after the composition with the Adjoint representation of GG on its Lie algebra 𝔤\mathfrak{g}, the curve is definable and non-contracting in GL(𝔤)\text{GL}(\mathfrak{g}). This property is valid in our situation. ∎

Appendix A Kempf’s Lemma for real algebraic groups

We note that [Kem78, Lemma 1.1(b)] is stated for reductive algebraic groups over an arbitrary field. We need the same result for any algebraic group defined over \mathbb{R}. Here, we verify that the proof given by Kempf is valid for real algebraic groups GG using the most naive language.

For finite-dimensional real vector spaces VV and WW, we say that P:VWP:V\to W is a polynomial map if, after choosing bases for VV and WW, the map PP in those coordinates is a multi-variable polynomial in each coordinate. We say that the map is GG-equvairant if g.P(x)=P(g.x)g.P(x)=P(g.x) for all gGg\in G and xVx\in V.

Lemma A.1 (Kempf).

Let GG be the set of \mathbb{R}-points of an algebraic group defined over \mathbb{R}. Let VV be a rational representation of GG. Suppose that for vVv\in V, it holds that S:=zcl(G.v)G.vS:=\text{zcl}(G.v)\smallsetminus G.v is non-empty. Then there exists a rational representation WW and an equivariant polynomial map P:VWP:V\to W such that P(S)={0}P(S)=\{0\} and P(v)0P(v)\neq 0.

Proof.

First, note that since zcl(G.v)\text{zcl}(G.v) and G.vG.v are GG-invariant, we get that SS is GG-invariant. Let [V]\mathbb{R}[V] be the coordinate ring on the vector space VV and consider I[V]I\subseteq\mathbb{R}[V] the polynomial ideal of the polynomials vanishing on SS. For dd\in\mathbb{N}, let I(d)II(d)\subseteq I be the subset of polynomials in II with degree bounded by dd. We may choose dd such that the ideal generated by I(d)I(d) is II. Notice that I(d)I(d) is a finite dimensional vector space. Consider the right action of GG on [V]\mathbb{R}[V] defined by

(A.1) (f.g)(x):=f(g.x),gG,xV,f[V].(f.g)(x):=f(g.x),~{}g\in G,~{}x\in V,~{}f\in\mathbb{R}[V].

The action preserves the degree of the polynomial. So, it preserves I(d)I(d) because SS is GG-invariant. Let I(d)I(d)^{*} be the space of linear functionals on I(d)I(d). Using the representation (A.1), we get a left action on I(d)I(d)^{*} defined by

(A.2) g.l(f):=l(f.g),gG,lI(d),fI(d).g.l(f):=l(f.g),~{}g\in G,~{}l\in I(d)^{*},~{}f\in I(d).

Now consider the map P:VI(d)P:V\to I(d)^{*} defined by

(A.3) P(x):=evx,P(x):=\text{ev}_{x},

where evx\text{ev}_{x} is the functional mapping fI(d)f\in I(d) to f(x)f(x). It is straightforward to verify that PP is an equivariant polynomial map. Finally, we claim that P(v)=evvP(v)=\text{ev}_{v} is a non-zero functional. By the definition of SS and PP, P(x)=0P(x)=0 for all xSx\in S. It remains to show that P(v)0P(v)\neq 0. For that, recall the closed orbit lemma [Bor91, Section 1.8], which states that SS is Zariski closed. This implies that there exists a pI(d)p\in I(d) such that p(S)=0p(S)=0 and p(v)0p(v)\neq 0. Since P(v)(p)=p(v)P(v)(p)=p(v), P(v)0P(v)\neq 0. ∎

References

  • [BM00] M. Bekka and Matthias Mayer “Ergodic Theory and Topological Dynamics of Group Actions on Homogeneous Spaces” In Cambridge University Press, 2000
  • [BHM63] A. Białynicki-Birula, G. Hochschild and G.. Mostow “Extensions of representations of algebraic linear groups” In Amer. J. Math. 85, 1963, pp. 131–144
  • [Bir71] David Birkes “Orbits of Linear Algebraic Groups” In Annals of Mathematics 93.3 Annals of Mathematics, 1971, pp. 459–475 URL: http://www.jstor.org/stable/1970884
  • [Bôc01] Maxime Bôcher “Certain cases in which the vanishing of the Wronskian is a sufficient condition for linear dependence” In Trans. Amer. Math. Soc. 2.2, 1901, pp. 139–149 DOI: 10.2307/1986214
  • [Bor91] Armand Borel “Linear Algebraic Groups, 2nd ed.” Springer, 1991
  • [Che51] Claude Chevalley “Théorie des groupes de Lie. Tome II. Groupes algébriques” No. 1152, Actualités Scientifiques et Industrielles [Current Scientific and Industrial Topics] Hermann & Cie, Paris, 1951, pp. vii+189
  • [Dan81] S.. Dani “Invariant measures and minimal sets of horospherical flows” In Invent. Math. 64.2, 1981, pp. 357–385 DOI: 10.1007/BF01389173
  • [Dan84] S.. Dani “On orbits of unipotent flows on homogeneous spaces” In Ergodic Theory and Dynamical Systems 4.1, 1984, pp. 25–34 DOI: 10.1017/S0143385700002248
  • [Dan86] S.. Dani “On orbits of unipotent flows on homogeneous spaces, II” In Ergodic Theory and Dynamical Systems 6.2, 1986, pp. 167–182 DOI: 10.1017/S0143385700003382
  • [DM93] S.. Dani and G.. Margulis “Limit distributions of orbits of unipotent flows and values of quadratic forms” In Adv. Soviet Math., vol. 16, Amer. Math. Soc., Providence, RI, 1993
  • [Dri98] Lou Dries “Tame topology and o-minimal structures” 248, London Mathematical Society Lecture Note Series Cambridge University Press, Cambridge, 1998, pp. x+180 DOI: 10.1017/CBO9780511525919
  • [DM96] Lou Dries and Chris Miller “Geometric categories and o-minimal structures” In Duke Math. J. 84.2, 1996, pp. 497–540 DOI: 10.1215/S0012-7094-96-08416-1
  • [EW11] Manfred Einsiedler and Thomas Ward “Ergodic theory: with a view towards number theory” Springer, 2011
  • [Gro97] Frank D. Grosshans “Observable subgroups” In Algebraic Homogeneous Spaces and Invariant Theory Berlin, Heidelberg: Springer Berlin Heidelberg, 1997, pp. 5–32 DOI: 10.1007/BFb0093527
  • [Kem78] George R. Kempf “Instability in invariant theory” In Ann. of Math. (2) 108.2, 1978, pp. 299–316 DOI: 10.2307/1971168
  • [Kle07] Dmitry Kleinbock “Quantitative nondivergence and its Diophantine applications” In Clay Mathematics Proceedings, 2007
  • [KM98] Dmitry.. Kleinbock and Grigory.. Margulis “Flows on homogeneous spaces and Diophantine approximation on manifolds” In Ann. of Math. (2) 148.1, 1998, pp. 339–360 DOI: 10.2307/120997
  • [Mar71] Grigory.. Margulis “On the action of unipotent groups in a lattice space” In Mathematics of the USSR-Sbornik 15.4, 1971, pp. 549 DOI: 10.1070/SM1971v015n04ABEH001560
  • [Mat60] Yozô Matsushima “Espaces homogènes de Stein des groupes de Lie complexes” In Nagoya Math. J. 16, 1960, pp. 205–218
  • [Mil94] Chris Miller “Expansions of the real field with power functions” In Ann. Pure Appl. Logic 68.1, 1994, pp. 79–94 DOI: 10.1016/0168-0072(94)90048-5
  • [Mil94a] Chris Miller “Exponentiation is hard to avoid” In Proc. Amer. Math. Soc. 122, 1994, pp. 257–259
  • [Mil94b] Christopher Lee Miller “Polynomially bounded o-minimal structures”, 1994
  • [MS95] Shahar Mozes and Nimish A. Shah “On the space of ergodic invariant measures of unipotent flows” In Ergodic Theory Dynam. Systems 15, 1995, pp. 149–159
  • [PS99] Ya’acov Peterzil and Charles Steinhorn “Definable Compactness and Definable Subgroups of o-Minimal Groups” In Journal of the London Mathematical Society 59.3, 1999, pp. 769–786 DOI: 10.1112/S0024610799007528
  • [PS18] Ya’acov Peterzil and Sergei Starchenko “O-minimal flows on nilmanifolds” In Duke Mathematical Journal, 2018 URL: https://api.semanticscholar.org/CorpusID:53135635
  • [PRR23] Vladimir Platonov, Andrei Rapinchuk and Igor Rapinchuk “Algebraic Groups and Number Theory”, Cambridge Studies in Advanced Mathematics Cambridge University Press, 2023
  • [Pou13] Georgios Poulios “Peterzil-steinhorn subgroups of real algebraic groups” Thesis (Ph.D.)–University of Notre Dame ProQuest LLC, Ann Arbor, MI, 2013, pp. 61 URL: http://gateway.proquest.com/openurl?url_ver=Z39.88-2004&rft_val_fmt=info:ofi/fmt:kev:mtx:dissertation&res_dat=xri:pqm&rft_dat=xri:pqdiss:3732192
  • [Rag72] Madabusi S. Raghunathan “Discrete Subgroups of Lie Groups” Springer Berlin, Heidelberg, 1972
  • [Rat91] Marina Ratner “On Raghunathan’s Measure Conjecture” In Ann. of Math. 134.3 Ann. of Math., 1991, pp. 545–607
  • [Rat91a] Marina Ratner “Raghunathan’s topological conjecture and distributions of unipotent flows” In Duke Mathematical Journal 63.1 Duke University Press, 1991, pp. 235–280 DOI: 10.1215/S0012-7094-91-06311-8
  • [Rem36] Evgeny.. Remez “Sour une propriété de polynômes de Tchebysheff,” In Communicationes le l’Inst. des Sci. Kharkow, 1936, pp. 93–95
  • [Sha91] Nimish Shah “Uniformly distributed orbits of certain flows on homogeneous spaces” In Mathematische Annalen 289, 1991, pp. 315–334 DOI: 10.1007/BF01446574
  • [Sha96] Nimish A. Shah “Limit distributions of expanding translates of certain orbits on homogeneous spaces” In Proceedings of the Indian Academy of Sciences - Mathematical Sciences 106, 1996, pp. 105–125
  • [Sha94] Nimish A. Shah “Limit distributions of polynomial trajectories on homogeneous spaces” In Duke Math. J. 75, 1994, pp. 711–732
  • [SY24] Nimish A. Shah and Pengyu Yang “Equidistribution of expanding degenerate manifolds in the space of lattices”, 2024 arXiv:2112.13952 [math.DS]
  • [Spr98] Tonny.. Springer “Linear Algebraic Groups” Birkhäuser Boston, MA, 1998
  • [TB05] Nguyêñ Quôć Thǎńg and Dao Phuong Bǎć “Some rationality properties of observable groups and related questions” In Illinois J. Math. 49.2, 2005, pp. 431–444
  • [Zim84] Robert J. Zimmer “Ergodic theory and semisimple groups” 81, Monographs in Mathematics Birkhäuser Verlag, Basel, 1984, pp. x+209