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

Double algebraic genericity of universal harmonic functions on trees

C. A. Konidas
Abstract

It is well known that the set of universal functions on a tree contains a vector space except zero which is dense in the set of harmonic functions. In this paper we improve this result by proving that the set of universal functions on a tree contains two vector spaces except zero which are dense in the space of harmonic functions and intersect only at zero.

AMS classification numbers: 05C05, 60J45, 60J50, 30K99, 46M99
Keywords and phrases: Tree, boundary of a tree, harmonic functions, universal functions, algebraic genericity

1 Introduction

Let TT be the set of vertices of a rooted tree with root x0x_{0}. For each nn\in\mathbb{N} we define TnT_{n} to be the set of all vertices at distance nn from x0x_{0} and we also define T0={x0}T_{0}=\{x_{0}\}. Given a vertex xTx\in T, we shall write S(x)S(x) for the set of the children of x. We assume that all TnT_{n} are finite and for every xTx\in T the set S(x)S(x) has at least two elements. Since T=n=0TnT=\bigcup_{n=0}^{\infty}{T_{n}} and each TnT_{n} is finite and nonempty, the set TT is infinite denumerable. With each xTx\in T and yS(x)y\in S(x) we associate a real number q(x,y)>0q(x,y)>0, which we think of as the probability of transition from vertex xx to vertex yy, such that

yS(x)q(x,y)=1.\sum_{y\in S(x)}{q(x,y)}=1.

We define the boundary of TT, denoted by T\partial T, as the set of all infinite geodesics originating from x0x_{0}. More specifically, an element eTe\in\partial T is of the form e={znT:n}e=\{z_{n}\in T:n\in\mathbb{N}\}, where z1=x0z_{1}=x_{0} and zn+1S(zn)z_{n+1}\in S(z_{n}) for all nn\in\mathbb{N}. For each xTx\in T we define the boundary sector of xx as Bx={eT:xe}B_{x}=\{e\in\partial T:x\in e\}. Notice that {Bx:xTn}\{B_{x}:x\in T_{n}\} is a partition of T\partial T for all nn\in\mathbb{N}. For each xT{x0}x\in T\setminus\{x_{0}\}, we have that xTnx\in T_{n} for some nn\in\mathbb{N} and we assign probability p(Bx)=j=0n1q(yj,yj+1)p(B_{x})=\prod_{j=0}^{n-1}{q(y_{j},y_{j+1})} where y0=x0,yn=xy_{0}=x_{0},y_{n}=x and yj+1S(yj)y_{j+1}\in S(y_{j}) for all j{1,,n}j\in\{1,\dots,n\}. For each nn\in\mathbb{N} we consider n\mathcal{M}_{n} to be the σ\sigma-algerba on T\partial T generated by {Bx:xTn}\{B_{x}:x\in T_{n}\}, which is a finite partition of T\partial T and therefore a function h:Th:\partial T\to\mathbb{C} is n\mathcal{M}_{n}-measurable if and only if it is constant on every BxB_{x} for xTnx\in T_{n}. We can extend p(Bx)p(B_{x}) to a probability measure n\mathbb{P}_{n} on the measurable space (T,n)(\partial T,\mathcal{M}_{n}). One can easily check that nn+1\mathcal{M}_{n}\subseteq\mathcal{M}_{n+1} and n+1|n=\mathbb{P}_{n+1}|_{\mathcal{M}_{n}}=\mathbb{P} for all nn\in\mathbb{N}. By Kolmogorov’s Consistency Theorem there exists a probability measure \mathbb{P} on the measurable space (T,)(\partial T,\mathcal{M}), where \mathcal{M} is the completion of the σ\sigma-algebra on T\partial T generated by the set n=1n\bigcup_{n=1}^{\infty}{\mathcal{M}_{n}}, such that |n=n\mathbb{P}|_{\mathcal{M}_{n}}=\mathbb{P}_{n} for all nn\in\mathbb{N}. A function h:Th:\partial T\to\mathbb{C} is \mathcal{M}-measurable if and only if for every open set VV\subseteq\mathbb{C} we have h1(V)h^{-1}(V)\in\mathcal{M}. By identifying almost everywhere equal \mathcal{M}-measurable functions we obtain the space L0(T,)L^{0}(\partial T,\mathbb{C}) or L0(T)L^{0}(\partial T) which we endow with the metric

P(ψ,ϕ)=T|ψ(e)ϕ(e)|1+|ψ(e)ϕ(e)|𝑑(e),P(\psi,\phi)=\int_{\partial T}{\frac{\left\lvert\psi(e)-\phi(e)\right\rvert}{1+\left\lvert\psi(e)-\phi(e)\right\rvert}}\,d\mathbb{P}(e),

where ψ,ϕL0(T)\psi,\phi\in L^{0}(\partial T). The induced topology is that of convergence in probability.

We consider the space T\mathbb{C}^{T} endowed with the metric

ρ(f,g)=n=112n|f(zn)g(xn)|1+|f(zn)g(zn)|,\rho(f,g)=\sum_{n=1}^{\infty}{\frac{1}{2^{n}}\frac{\left\lvert f(z_{n})-g(x_{n})\right\rvert}{1+\left\lvert f(z_{n})-g(z_{n})\right\rvert}},

where f,gTf,g\in\mathbb{C}^{T} and {zn:n}\{z_{n}:n\in\mathbb{N}\} is a enumeration of TT. The induced topology is that of pointwise convergence. A function fTf\in\mathbb{C}^{T} is called harmonic if and only if for all xTx\in T it holds that

f(x)=yS(x)q(x,y)f(y).f(x)=\sum_{y\in S(x)}{q(x,y)f(y)}.

The set of all harmonic functions shall be denoted by H(T,)H(T,\mathbb{C}) or H(T)H(T). For every fH(T)f\in H(T) and nn\in\mathbb{N} we define a function ωn(f):T\omega_{n}(f):\partial T\to\mathbb{C} by ωn(f)(e)=f(z)\omega_{n}(f)(e)=f(z), where zz is the unique element of eTne\cap T_{n} and eTe\in\partial T. Notice that ωn(f)\omega_{n}(f) is constant on each BxB_{x} for xTnx\in T_{n} and thus it is n\mathcal{M}_{n}-measurable. A harmonic function fH(T)f\in H(T) is called universal if and only if the sequence {ωn(f)}n=1\{\omega_{n}(f)\}_{n=1}^{\infty} is dense in L0(T)L^{0}(\partial T). The set of universal functions shall be denoted by U(T,)U(T,\mathbb{C}) or U(T)U(T). The following theorem was proved in [3].

Theorem 1.1.

The set U(T){0}U(T)\cup\{0\} contains a vector space which is dense in H(T)H(T), that is we have algebraic genericity for the set U(T)U(T).

In this paper we will improve this result by proving the following theorem.

Theorem 1.2.

There exist two vector spaces F1,F2F_{1},F_{2} contained in U(T){0}U(T)\cup\{0\} which are dense in H(T)H(T) such that F1F2={0}F_{1}\cap F_{2}=\{0\}, that is we have double algebraic genericity for the set U(T)U(T).

In order to prove this we will consider two particular subsets of U(T)U(T) which have been studied before in [1, 2, 3, 4]. We say that a harmonic function f:Tf:T\to\mathbb{C} is frequently universal if and only if for every nonempty open set VL0(T)V\subseteq L^{0}(\partial T) the set {n:ωn(f)V}\{n\in\mathbb{N}:\omega_{n}(f)\in V\} has strictly positive lower density. The set of frequently universal functions shall be denoted by UFM(T,)U_{FM}(T,\mathbb{C}) or UFM(T)U_{FM}(T). We say that a harmonic function f:Tf:T\to\mathbb{C} belongs to the class X(T,)X(T,\mathbb{C}) or X(T)X(T) if and only if for every nonempty open set VL0(T)V\subseteq L^{0}(\partial T) the set {n:ωn(f)V}\{n\in\mathbb{N}:\omega_{n}(f)\in V\} has upper density equal to one. Algebraic genericity for the set UFM(T)U_{FM}(T) and the fact that UFM(T)U_{FM}(T) and X(T)X(T) are disjoint have been proven in [4]. So, in order to prove Theorem 1.2, we will prove algebraic genericity for the set X(T)X(T).

2 A more general setting

We will provide a more general proof of our result. In particular we will replace \mathbb{C} by any normed space over \mathbb{C} and we will consider a generalized definition of harmonic functions introduced in [4]. However, in this section we will not assume that EE is a normed space as we will use some of the following results when considering a space in which the metric is not induced by a norm.

Let EE be a separable topological vector space over \mathbb{C}, that is metrizable with a metric dd. We consider the space ETE^{T} with the metric

ρ(f,g)=n=112nd(f(zn),g(zn))1+d(f(zn),g(zn)),\rho(f,g)=\sum_{n=1}^{\infty}{\frac{1}{2^{n}}\frac{d(f(z_{n}),g(z_{n}))}{1+d(f(z_{n}),g(z_{n}))}},

where f,gETf,g\in E^{T} and {zn:n}\{z_{n}:n\in\mathbb{N}\} is a enumeration of TT. Since EE is separable and TT denumerable, the metric space (ET,ρ)(E^{T},\rho) is separable.

Definition 2.1.

A function f:TEf:T\to E is called generalized harmonic if and only if for all xTx\in T it holds that

f(x)=yS(x)w(x,y)f(y),f(x)=\sum_{y\in S(x)}{w(x,y)f(y)},

where w(x,y){0}w(x,y)\in\mathbb{C}\setminus\{0\} such that yS(x)w(x,y)=1\sum_{y\in S(x)}{w(x,y)}=1. The set of generalized harmonic functions shall be denoted by H(T,E)H(T,E).

From now we will drop the word generalized and refer to the elements of H(T,E)H(T,E) simply as harmonic functions. Notice that the set H(T,E)H(T,E) is a separable metric space as a metric subspace of the separable metric space (ET,ρ)(E^{T},\rho).

A function h:TEh:\partial T\to E is \mathcal{M}-measurable if and only if for every open set VEV\subseteq E we have that h1(V)h^{-1}(V)\in\mathcal{M}. By identifying almost everywhere equal \mathcal{M}-measurable functions we obtain the space L0(T,E)L^{0}(\partial T,E). We endow L0(T,E)L^{0}(\partial T,E) with the metric

P(ψ,ϕ)=Td(ψ(e),ϕ(e))1+d(ψ(e),ϕ(e))𝑑(e),P(\psi,\phi)=\int_{\partial T}{\frac{d(\psi(e),\phi(e))}{1+d(\psi(e),\phi(e))}}\,d\mathbb{P}(e),

where ψ,ϕL0(T,E)\psi,\phi\in L^{0}(\partial T,E). The topology induced by PP is that of convergence in probability. Because EE is separable and by construction of the measure space (T,,)(\partial T,\mathcal{M},\mathbb{P}) there exists a dense sequence {hn}n=1\{h_{n}\}_{n=1}^{\infty} in L0(T,E)L^{0}(\partial T,E), such that each hnh_{n} is k(n)\mathcal{M}_{k(n)}-measurable for some k(n)k(n)\in\mathbb{N}.

In general, the set of all measurable functions from a measurable space to some metrizable topological vector space needs not be a vector space. However, in this setting we have the following proposition concerning the structure of L0(T,E)L^{0}(\partial T,E).

Proposition 2.2.

Let EE be a separable topological vector space over \mathbb{C}, metrizable with a metric dd. The following are true.

  1. (i)

    The space L0(T,E)L^{0}(\partial T,E) contains no isolated points.

  2. (ii)

    The set L0(T,E)L^{0}(\partial T,E) when considered with pointwise addition and \mathbb{C}-scalar multiplication is a vector space.

  3. (iii)

    If the metric dd of EE is translation invariant, then addition on L0(T,E)L^{0}(\partial T,E) is continuous.

  4. (iv)

    If the metric dd of EE is induced by a norm, then scalar multiplication on L0(T,E)L^{0}(\partial T,E) is continuous.

Proof.
  1. (i)

    This was proved as a Lemma in [3], without using any of the other claims stated in this proposition.

  2. (ii)

    It suffices to prove that if f:TEf:\partial T\to E and g:TEg:\partial T\to E are two \mathcal{M}-measurable functions and aa\in\mathbb{C} is a complex number, then the functions f+gf+g and afaf are \mathcal{M}-measurable. Since the set {hn:n}\{h_{n}:n\in\mathbb{N}\} is dense in L0(T,E)L^{0}(\partial T,E) we can find a sequence in that set that convergences to ff in probability. We can then find a subsequence of that sequence that convergences to ff pointwise almost everywhere. By reindexing we can find a sequence {fn}n=1\{f_{n}\}_{n=1}^{\infty} in the set {hn:n}\{h_{n}:n\in\mathbb{N}\} such that fnff_{n}\to f pointwise almost everywhere and each fnf_{n} is τ(n)\mathcal{M}_{\tau(n)}-measurable for some τ(n)\tau(n)\in\mathbb{N}. By the same argument we can find a sequence {gn}n=1\{g_{n}\}_{n=1}^{\infty} in the set {hn:n}\{h_{n}:n\in\mathbb{N}\} such that gngg_{n}\to g pointwise almost everywhere and each gng_{n} is (n)\mathcal{M}_{\ell(n)}-measurable for some (n)\ell(n)\in\mathbb{N}. Thus we can find a set AA\in\mathcal{M} with (A)=1\mathbb{P}(A)=1 such that fn(e)f(e)f_{n}(e)\to f(e) and gn(e)g(e)g_{n}(e)\to g(e) for all eAe\in A. Since addition on EE is continuous we have that fn(e)+gn(e)f(e)+g(e)f_{n}(e)+g_{n}(e)\to f(e)+g(e) for all eAe\in A. By setting r(n)=max{τ(n),(n)}r(n)=\max\{\tau(n),\ell(n)\} for all nn\in\mathbb{N} we have that τ(n)r(n)\mathcal{M}_{\tau(n)}\subseteq\mathcal{M}_{r(n)} and (n)r(n)\mathcal{M}_{\ell(n)}\subseteq\mathcal{M}_{r(n)} for all nn\in\mathbb{N}. Therefore, both fnf_{n} and gng_{n} are r(n)\mathcal{M}_{r(n)}-measurable for all nn\in\mathbb{N} and so they are constant on all the sets BxB_{x} for xTr(n)x\in T_{r(n)}. This implies that the sum fn+gnf_{n}+g_{n} is constant on all the sets BxB_{x} for xTr(n)x\in T_{r(n)} and thus, the sum fn+gnf_{n}+g_{n} is r(n)\mathcal{M}_{r(n)}-measurable and therefore \mathcal{M}-measurable. So, we have that f+gf+g is an almost everywhere pointwise limit of the sequence of \mathcal{M}-measurable functions {fn+gn}n=1\{f_{n}+g_{n}\}_{n=1}^{\infty}, the values of which are elements of the metric space (E,d)(E,d), and the measure space (T,,)(\partial T,\mathcal{M},\mathbb{P}) is complete. This implies that the function f+gf+g is \mathcal{M}-measurable. The proof that afaf is \mathcal{M}-measurable is similar, it requires the fact that multiplication on EE is continuous, and is omitted.

  3. (iii)

    The proof of this claim is similar to the proof of the respective claim about real-valued functions, which is well known. It suffices to replace the absolute value in \mathbb{R} by a translation invariant metric dd. Thus, the proof is omitted.

  4. (iv)

    The proof of this claim is similar to the proof of the respective claim about real-valued functions, which is well known. It suffices to replace the absolute value in \mathbb{R} by a norm induced metric dd. Thus, the proof is omitted.

For every fH(T,E)f\in H(T,E) and nn\in\mathbb{N} we define a function ωn(f):TE\omega_{n}(f):\partial T\to E by ωn(f)(e)=f(z)\omega_{n}(f)(e)=f(z), where zz is the unique element of eTne\cap T_{n} and eTe\in\partial T. Notice that ωn(f)\omega_{n}(f) is constant on each BxB_{x} for xTnx\in T_{n} and thus it is n\mathcal{M}_{n}-measurable.

Definition 2.3.

A harmonic function fH(T,E)f\in H(T,E) is called universal if and only if the sequence {ωn(f)}n=1\{\omega_{n}(f)\}_{n=1}^{\infty} is dense in L0(T,E)L^{0}(\partial T,E). The set of universal harmonic functions shall be denoted with U(T,E)U(T,E).

The following theorem was proved in [3].

Theorem 2.4.

Let EE be a separable metrizable topological vector space over \mathbb{C}. The set U(T,E){0}U(T,E)\cup\{0\} contains a vector space which is dense in H(T,E)H(T,E), that is we have algebraic genericity for the set U(T,E)U(T,E).

We will now consider some classes of universal functions.

Definition 2.5.

A harmonic function fH(T,E)f\in H(T,E) is called frequently universal if and only if for every nonempty open set VL0(T,E)V\subseteq L^{0}(\partial T,E) the set {n:ωn(f)V}\{n\in\mathbb{N}:\omega_{n}(f)\in V\} has strictly positive lower density. The set of frequently universal functions shall be denoted by UFM(T,E)U_{FM}(T,E).

Definition 2.6.

A harmonic function fH(T,E)f\in H(T,E) is said to belong to the class X(T,E)X(T,E) if and only if for every nonempty open set VL0(T,E)V\subseteq L^{0}(\partial T,E) the set {n:ωn(f)V}\{n\in\mathbb{N}:\omega_{n}(f)\in V\} has upper density equal to one.

The following theorem was proved in [3].

Theorem 2.7.

Let EE be a separable metrizable topological vector space over \mathbb{C}. The class X(T,E)X(T,E) is GδG_{\delta} and dense in H(T,E)H(T,E).

Notice that fX(T,E)f\in X(T,E) if and only if for every nonempty non-dense open set VL0(T,E)V\subseteq L^{0}(\partial T,E) the set {n:ωn(f)V}\{n\in\mathbb{N}:\omega_{n}(f)\in V\} has lower density equal to zero. From this observation and the existence of a nonempty non-dense open set VL0(T,E)V\subseteq L^{0}(\partial T,E) we can deduce the following proposition, which was proved in a different way in [4].

Proposition 2.8.

Let EE be a separable metrizable topological vector space over \mathbb{C}. The sets UFM(T,E)U_{FM}(T,E) and X(T,E)X(T,E) are disjoint, that is

UFM(T,E)X(T,E)=.U_{FM}(T,E)\cap X(T,E)=\varnothing.

3 Algebraic genericity

In this section we will assume that the metric dd of EE is induced by a norm as we want L0(T,E)L^{0}(\partial T,E) to be a topological vector space when considered with the topology of convergence in probability. Under this assumption we can prove the following theorem, the proof of which can be found in [4, 3].

Theorem 3.1.

Let EE be normed space over \mathbb{C}. The set UFM(T,E){0}U_{FM}(T,E)\cup\{0\} contains a vector space which is dense in H(T,E)H(T,E).

We will now proceed to proving the same result for the class X(T,E)X(T,E). For this we will consider generalized harmonic functions f:TEf:T\to E^{\mathbb{N}}, where EE^{\mathbb{N}} is a separable metrizable topological vector space when considered with its product metric d^\hat{d} which is translation invariant. Notice that a function f=(f1,f2,)f=(f_{1},f_{2},\dots) belongs to set H(T,E)H(T,E^{\mathbb{N}}) if and only if the function fnf_{n} belongs to the set H(T,E)H(T,E) for all nn\in\mathbb{N}. By Proposition 2.2 we know that L0(T,E)L^{0}(\partial T,E) is a topological vector space and L0(T,E)L^{0}(\partial T,E^{\mathbb{N}}) is a vector space in which addition is continuous, each considered with the topology of convergence in probability. We will now prove two lemmas that will lead to the result.

Lemma 3.2.

Let EE be normed space over \mathbb{C}. If f=(f1,f2,)X(T,E)f=(f_{1},f_{2},\dots)\in X(T,E^{\mathbb{N}}) then the vector space span{fn:n}\operatorname{span}\{f_{n}:n\in\mathbb{N}\} is contained in the set X(T,E){0}X(T,E)\cup\{0\}.

Proof.

Let a1f1++asfsa_{1}f_{1}+\dots+a_{s}f_{s} be an arbitrary element of the set span{fn:n}\operatorname{span}\{f_{n}:n\in\mathbb{N}\} with as0a_{s}\neq 0. Let VV be a nonempty open subset of L0(T,E)L^{0}(\partial T,E). There exist some ε>0\varepsilon>0 and hVh\in V such that B(h,ε)VB(h,\varepsilon)\subseteq V. For i{1,,s}i\in\{1,\dots,s\} we define

bi={ai,if ai01,if ai=0.b_{i}=\begin{cases}a_{i},&\text{if }a_{i}\neq 0\\ 1,&\text{if }a_{i}=0\end{cases}.

We set δ=1sε\delta=\displaystyle{\frac{1}{s\cdot\varepsilon}} and consider the set

V^=1b1B(0,δ)×1b2B(0,δ)××1bsB(h,δ)×L0(T,E)×.\widehat{V}=\frac{1}{b_{1}}B(0,\delta)\times\frac{1}{b_{2}}B(0,\delta)\times\cdots\times\frac{1}{b_{s}}B(h,\delta)\times L^{0}(\partial T,E)\times\cdots.

We now fix some nn\in\mathbb{N} and suppose that ωn(f)V^\omega_{n}(f)\in\widehat{V}. Then for every i{1,,s1}i\in\{1,\dots,s-1\} we have

ωn(fi)1biB(0,δ) and ωn(fs)1bsB(h,δ).\omega_{n}(f_{i})\in\frac{1}{b_{i}}B(0,\delta)\text{~{}~{}~{}and~{}~{}~{}}\omega_{n}(f_{s})\in\frac{1}{b_{s}}B(h,\delta).

By the definition of b1,,bnb_{1},\dots,b_{n} and the linearity of ωn()\omega_{n}(\cdot) we conclude that for all i{1,,s1}i\in\{1,\dots,s-1\} we have

P(ωn(aifi),0)<δ and P(ωn(asfs),h)<δ.P(\omega_{n}(a_{i}f_{i}),0)<\delta\text{~{}~{}~{}and~{}~{}~{}}P(\omega_{n}(a_{s}f_{s}),h)<\delta.

The metric PP is translation invariant, since the metric dd is translation invariant, thus

P(ωn(a1f1++asfs),h)\displaystyle P(\omega_{n}(a_{1}f_{1}+\dots+a_{s}f_{s}),h) P(ωn(a1f1),0)++P(ωn(as1fs1),0)+P(ωn(asfs),h)\displaystyle\leqslant P(\omega_{n}(a_{1}f_{1}),0)+\dots+P(\omega_{n}(a_{s-1}f_{s-1}),0)+P(\omega_{n}(a_{s}f_{s}),h)
<sδ.\displaystyle<s\cdot\delta.

Since sδ=εs\cdot\delta=\varepsilon we conclude that ωn(a1f1++asfs)B(h,ε)\omega_{n}(a_{1}f_{1}+\dots+a_{s}f_{s})\in B(h,\varepsilon), where B(h,ε)VB(h,\varepsilon)\subseteq V. Therefore, we have proved that

{n:ωn(f)V^}{n:ωn(a1f1++asfs)V}.\{n\in\mathbb{N}:\omega_{n}(f)\in\widehat{V}\}\subseteq\{n\in\mathbb{N}:\omega_{n}(a_{1}f_{1}+\dots+a_{s}f_{s})\in V\}.

Notice that convergence in probability in L0(T,E)L^{0}(\partial T,E^{\mathbb{N}}), which is the same set as L0(T,E)L^{0}(\partial T,E)^{\mathbb{N}}, is equivalent to convergence in the product topology when considering L0(T,E)L^{0}(\partial T,E) with the topology of convergence in probability. Therefore V^\widehat{V} is an open subset of L0(T,E)L^{0}(\partial T,E^{\mathbb{N}}) and it is also nonempty. Since fX(T,E)f\in X(T,E), the set {n:ωn(f)V^}\{n\in\mathbb{N}:\omega_{n}(f)\in\widehat{V}\} has upper density equal to one, which implies that the set {n:ωn(a1f1++asfs)V}\{n\in\mathbb{N}:\omega_{n}(a_{1}f_{1}+\dots+a_{s}f_{s})\in V\} has upper density equal to one. Thus a1f1+asfsX(T,E)a_{1}f_{1}+\dots a_{s}f_{s}\in X(T,E). ∎

Lemma 3.3.

Let EE be normed space over \mathbb{C}. There exists some sequence {fn}n=1\{f_{n}\}_{n=1}^{\infty} dense in H(T,E)H(T,E) such that the function f=(f1,f2,)f=(f_{1},f_{2},\dots) belongs to the class X(T,E)X(T,E^{\mathbb{N}}).

Proof.

The metric space H(T,E)H(T,E) is separable, so let {φn:n}\{\varphi_{n}:n\in\mathbb{N}\} be a dense sequence in H(T,E)H(T,E). Since EE^{\mathbb{N}} when considered with the product metric d^\hat{d} is a separable topological vector space, Theorem 2.7 applies. Thus, the class X(T,E)X(T,E^{\mathbb{N}}) is dense in H(T,E)H(T,E^{\mathbb{N}}) and in particular nonempty, so let f=(f1,f2,)f=(f_{1},f_{2},\dots) be an element of X(T,E)X(T,E^{\mathbb{N}}). We now fix some nn\in\mathbb{N}. There exists some j0(n)j_{0}(n)\in\mathbb{N} such that

j=j0(n)12j<1n.\sum_{j=j_{0}(n)}^{\infty}{\frac{1}{2^{j}}}<\frac{1}{n}.

There also exists some N(n)N(n)\in\mathbb{N} such that

{z0,,zj0(n)}k=0N(n)Tk,\{z_{0},\dots,z_{j_{0}(n)}\}\subseteq\bigcup_{k=0}^{N(n)}{T_{k}},

where {zn:n}\{z_{n}:n\in\mathbb{N}\} is the denumeration of TT through which the metric ρ\rho was defined. We set gn(x)=φn(x)fn(x)g_{n}(x)=\varphi_{n}(x)-f_{n}(x) for all xk=0N(n)Tkx\in\bigcup_{k=0}^{N(n)}{T_{k}} and we set gn(x)=gn(x)g_{n}(x)=g_{n}(x^{-}) for all xk=N(n)+1Tkx\in\bigcup_{k=N(n)+1}^{\infty}{T_{k}}, where xx^{-} is the father of xx. Notice that the function gn:TEg_{n}:T\to E defined this way is a generalized harmonic function such that ωk(gn)=ωN(n)(gn)\omega_{k}(g_{n})=\omega_{N(n)}(g_{n}) for all k>N(n)k>N(n) and ρ(φnfn,gn)<1/n\rho(\varphi_{n}-f_{n},g_{n})<1/n. Since the metric dd is translation invariant it follows that the metric ρ\rho is translation invariant, thus ρ(fn+gn,gn)<1/n\rho(f_{n}+g_{n},g_{n})<1/n for all nn\in\mathbb{N}. Since H(T,E)H(T,E) has no isolated points, the sequence {fn+gn}n=1\{f_{n}+g_{n}\}_{n=1}^{\infty} is dense in H(T,E)H(T,E). We will now show that the function f+g=(f1+g1,f2+g2,)H(T,E)f+g=(f_{1}+g_{1},f_{2}+g_{2},\dots)\in H(T,E^{\mathbb{N}}) belongs to the class X(T,E)X(T,E^{\mathbb{N}}) which completes the proof. It suffices to show that for every nonempty open basic set VL0(T,E)V\subseteq L^{0}(\partial T,E) the set {n:ωn(f+g)V}\{n\in\mathbb{N}:\omega_{n}(f+g)\in V\} has upper density equal to one. Since convergence in probability in L0(T,E)L^{0}(\partial T,E^{\mathbb{N}}), which is the same set as L0(T,E)L^{0}(\partial T,E)^{\mathbb{N}}, is equivalent to convergence in the product topology when considering L0(T,E)L^{0}(\partial T,E) with the topology of convergence in probability, a nonempty basic open set of L0(T,E)L^{0}(\partial T,E^{\mathbb{N}}) can be written as

V=V1××Vs×L0(T,E)×,V=V_{1}\times\cdots\times V_{s}\times L^{0}(\partial T,E)\times\cdots,

where V1,,VsV_{1},\dots,V_{s} are nonempty open subsets of L0(T,E)L^{0}(\partial T,E). We set L=max{N(1),,N(s)}L=\max\{N(1),\dots,N(s)\}. Then, for all nLn\geqslant L and all j{1,,s}j\in\{1,\dots,s\} we have ωn(gj)=ωL(gj)\omega_{n}(g_{j})=\omega_{L}(g_{j}). We fix some nn\in\mathbb{N} with nLn\geqslant L and suppose that ωn(f)VωL(g)\omega_{n}(f)\in V-\omega_{L}(g). Then ωn(fj)VjωL(fj)\omega_{n}(f_{j})\in V_{j}-\omega_{L}(f_{j}) for all j{1,,s}j\in\{1,\dots,s\}, which implies that ωn(fj)+ωn(gj)Vj\omega_{n}(f_{j})+\omega_{n}(g_{j})\in V_{j}. But ωn(fj+gj)=ωn(fj)+ωn(gj)\omega_{n}(f_{j}+g_{j})=\omega_{n}(f_{j})+\omega_{n}(g_{j}) for all j{1,,n}j\in\{1,\dots,n\}, so we have that ωn(fj+gj)Vj\omega_{n}(f_{j}+g_{j})\in V_{j} for all j{1,,n}j\in\{1,\dots,n\}, which in turn implies that ωn(f+g)V\omega_{n}(f+g)\in V. Thus, we have proved that

{nL:ωn(f)VωL(g)}{n:ωn(f+g)V}.\{n\geqslant L:\omega_{n}(f)\in V-\omega_{L}(g)\}\subseteq\{n\in\mathbb{N}:\omega_{n}(f+g)\in V\}.

But the set VωL(g)V-\omega_{L}(g) is a nonempty open subset of L0(T,E)L^{0}(\partial T,E^{\mathbb{N}}) and fX(T,E)f\in X(T,E^{\mathbb{N}}), therefore the set {n:ωn(f)VωL(g)}\{n\in\mathbb{N}:\omega_{n}(f)\in V-\omega_{L}(g)\} has upper density equal to one. This implies that the set {nL:ωn(f)VωL(g)}\{n\geqslant L:\omega_{n}(f)\in V-\omega_{L}(g)\} has upper density equal to one and therefore the set {n:ωn(f+g)V}\{n\in\mathbb{N}:\omega_{n}(f+g)\in V\} has upper density equal to one. ∎

We can now prove the following theorem.

Theorem 3.4.

Let EE be normed space over \mathbb{C}. The set X(T,E){0}X(T,E)\cup\{0\} contains a vector space which is dense in H(T,E)H(T,E).

Proof.

By Lemma 3.3 there exists some f=(f1,f2,)X(T,E)f=(f_{1},f_{2},\dots)\in X(T,E^{\mathbb{N}}) such that the sequence {fn}n=1\{f_{n}\}_{n=1}^{\infty} is dense in H(T,E)H(T,E). By Lemma 3.2 the vector space span{fn:n}\operatorname{span}\{f_{n}:n\in\mathbb{N}\} is contained in the set X(T,E){0}X(T,E)\cup\{0\} and the sequence {fn}n=1\{f_{n}\}_{n=1}^{\infty} is contained in the vector space span{fn:n}\operatorname{span}\{f_{n}:n\in\mathbb{N}\}, which implies that vector space span{fn:n}\operatorname{span}\{f_{n}:n\in\mathbb{N}\} is dense in H(T,E)H(T,E). ∎

We can now prove the main result of this paper.

Theorem 3.5.

Let EE be normed space over \mathbb{C}. There exists two vector spaces F1,F2F_{1},F_{2} contained in U(T,E){0}U(T,E)\cup\{0\} which are dense in H(T,E)H(T,E) such that F1F2={0}F_{1}\cap F_{2}=\{0\}, that is we have double algebraic genericity for the set U(T,E)U(T,E).

Proof.

By Theorem 3.1 there exists a vector space F1F_{1} contained in UFM(T,E){0}U_{FM}(T,E)\cup\{0\}, which is a subset of U(T,E){0}U(T,E)\cup\{0\}, that is dense in H(T,E)H(T,E). By Theorem 3.4 there exists a vector space F2F_{2} contained in X(T,E){0}X(T,E)\cup\{0\}, which is a subset of U(T,E){0}U(T,E)\cup\{0\}, that is dense in H(T,E)H(T,E). By Proposition 2.8 we have that F1F2={0}F_{1}\cap F_{2}=\{0\}. ∎

Acknowledgements

I would like to thank V. Nestoridis for his assistance in my effort of understanding [3] and [4] and also for his remarks and advice concerning this paper.

References

  • [1] E. Abakumov, V. Nestoridis and M.. Picardello “Frequently dense harmonic functions and universal martingales on trees” In Proc. Amer. Math. Soc. 149.5, 2021, pp. 1905–1918 DOI: https://doi.org/10.1090/proc/15355
  • [2] E. Abakumov, V. Nestoridis and M.. Picardello “Universal properties of harmonic functions on trees” In Journal of Mathematical Analysis and Applications 445.2, 2017, pp. 1181–1187 DOI: https://doi.org/10.1016/j.jmaa.2016.03.078
  • [3] N. Biehler, E. Nestoridi and V. Nestoridis “Generalized harmonic functions on trees: Universality and frequent universality” In Journal of Mathematical Analysis and Applications 503.1, 2021, pp. 125277 DOI: https://doi.org/10.1016/j.jmaa.2021.125277
  • [4] N. Biehler, V. Nestoridis and A. Stavrianidi “Algebraic genericity of frequently universal harmonic functions on trees” In Journal of Mathematical Analysis and Applications 489.1, 2020, pp. 124132 DOI: https://doi.org/10.1016/j.jmaa.2020.124132

C.A. Konidas
National and Kapodistrian University of Athens
Department of Mathematics,
e-mail: [email protected]