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

The Banach–Mazur–Schmidt and Banach–Mazur–McMullen games

Lior Fishman University of North Texas, Department of Mathematics, 1155 Union Circle #311430, Denton, TX 76203-5017, USA lior.fishman@unt.edu Vanessa Reams University of North Texas, Department of Mathematics, 1155 Union Circle #311430, Denton, TX 76203-5017, USA vanessareams@my.unt.edu  and  David Simmons University of York, Department of Mathematics, Heslington, York YO10 5DD, UK simmons.465@osu.edu https://sites.google.com/site/davidsimmonsmath/
Abstract.

We introduce two new mathematical games, the Banach–Mazur–Schmidt game and the Banach–Mazur–McMullen game, merging well-known games. We investigate the properties of the games, as well as providing an application to Diophantine approximation theory, analyzing the geometric structure of certain Diophantine sets.

1. Introduction

1.1. Schmidt’s game and the Banach–Mazur game

The Banach–Mazur game, dating back to 1935, is arguably the prototype for all infinite mathematical games. This game has been extensively studied and we refer the interested reader to [10, 8] for a thorough historical overview and recent developments. One of the most interesting aspects of the game is its connection to topology, namely that one of the players has a winning strategy if and only if the target set is comeager.

In 1966, W. M. Schmidt [9] introduced a two-player game referred to thereafter as Schmidt’s game. This game may be considered in a sense as a variant of the Banach–Mazur game. Schmidt invented the game primarily as a tool for studying certain sets which arise in number theory and Diophantine approximation theory. These sets are often exceptional with respect to both measure and category. The most significant example is the following. Let \mathbb{Q} denote the set of rational numbers. A real number xx is said to be badly approximable if there exists a positive constant c=c(x)c=c(x) such that |xpq|>cq2\left|x-\frac{p}{q}\right|>\frac{c}{q^{2}} for all pq\frac{p}{q}\in\mathbb{Q}. We denote the set of badly approximable numbers by BA\mathrm{BA}. This set plays a major role in Diophantine approximation theory, and is well-known to be both meager and Lebesgue null. Nonetheless, using his game, Schmidt was able to prove the following remarkable result:

Theorem 1.1 (Schmidt [9]).

Let (fn)n=1(f_{n})_{n=1}^{\infty} be a sequence of 𝒞1\mathcal{C}^{1} diffeomorphisms of \mathbb{R}. Then the Hausdorff dimension of the set n=1fn1(BA)\bigcap_{n=1}^{\infty}f^{-1}_{n}(\mathrm{BA}) is 11. In particular, n=1fn1(BA)\bigcap_{n=1}^{\infty}f^{-1}_{n}(\mathrm{BA}) is uncountable.

Remark 1.2.

We shall describe the games in the context of complete metric spaces. One could consider a more general framework of topological games, but as all of our applications and results are in this more restricted context, we prefer not to follow the most general presentation.

1.2. Description of games

Let (X,d)(X,d) be a complete metric space. In what follows, we denote by B(x,r)B(x,r) and B(x,r)B^{\circ}(x,r) the closed and open balls in the metric space (X,d)(X,d) centered at xx of radius rr, i.e.,

(1.1) B(x,r)=def{yX:d(x,y)r},B(x,r)=def{yX:d(x,y)<r}.B(x,r){\,\stackrel{{\scriptstyle\mathrm{def}}}{{=}}\,}\{y\in X:d(x,y)\leq r\},\;\;B^{\circ}(x,r){\,\stackrel{{\scriptstyle\mathrm{def}}}{{=}}\,}\{y\in X:d(x,y)<r\}.

Let Ω=defX×+\Omega{\,\stackrel{{\scriptstyle\mathrm{def}}}{{=}}\,}X\times\mathbb{R}_{+} be the set of formal balls in XX, and define a partial ordering on Ω\Omega by letting

(x2,r2)s(x1,r1) if r2+d(x1,x2)r1.(x_{2},r_{2})\leq_{s}(x_{1},r_{1})\text{ if }r_{2}+d(x_{1},x_{2})\leq r_{1}.

We associate to each pair (x,r)(x,r) a closed ball and an open ball in (X,d)(X,d) via the ‘ball’ functions B(,)B(\cdot,\cdot) and B(,)B^{\circ}(\cdot,\cdot) as in (1.1). Note that the inequality (x2,r2)s(x1,r1)(x_{2},r_{2})\leq_{s}(x_{1},r_{1}) clearly implies (but is not necessarily implied by111For example, if X={x0}X=\{x_{0}\} is a singleton then B(x0,r2)B(x0,r1)B(x_{0},r_{2})\subseteq B(x_{0},r_{1}) for all r1,r2>0r_{1},r_{2}>0, but the inequality (x0,r2)s(x0,r1)(x_{0},r_{2})\leq_{s}(x_{0},r_{1}) only holds if r2r1r_{2}\leq r_{1}.) the inclusion B(x2,r2)B(x1,r1)B(x_{2},r_{2})\subseteq B(x_{1},r_{1}). Nevertheless, the two conditions are equivalent when (X,d)(X,d) is a Banach space.

For Schmidt’s game, fix α,β(0,1)\alpha,\beta\in(0,1) and SXS\subseteq X. The set SS will be called the target set. Schmidt’s (α,β,S)(\alpha,\beta,S)-game is played by two players, whom we shall call Alice and Bob. The game starts with Bob choosing a pair ω1=(x1,r1)Ω\omega_{1}=(x_{1},r_{1})\in\Omega. Alice and Bob then take turns choosing pairs ωn=(xn,rn)sωn\omega^{\prime}_{n}=(x^{\prime}_{n},r^{\prime}_{n})\leq_{s}\omega_{n} and ωn+1=(xn+1,rn+1)sωn\omega_{n+1}=(x_{n+1},r_{n+1})\leq_{s}\omega^{\prime}_{n}, respectively. These pairs are required to satisfy

(1.2) rn=αrn and rn+1=βrn.r_{n}^{\prime}=\alpha r_{n}\text{ and }r_{n+1}=\beta r_{n}^{\prime}\,.

Since the game is played on a complete metric space and since the diameters of the nested balls

(1.3) B(ω1)B(ωn)B(ωn)B(ωn+1)B(\omega_{1})\supseteq\ldots\supseteq B(\omega_{n})\supseteq B(\omega^{\prime}_{n})\supseteq B(\omega_{n+1})\supseteq\ldots

tend to zero as nn\rightarrow\infty, the intersection of these balls is a singleton {x}\{x_{\infty}\}. Call Alice the winner if xSx_{\infty}\in S; otherwise Bob is declared the winner. A strategy consists of a description of how one of the players should act based on the opponent’s previous moves. A strategy is winning if it guarantees the player a win regardless of the opponent’s moves. If Alice has a winning strategy for Schmidt’s (α,β,S)(\alpha,\beta,S)-game, we say that SS is an (α,β)(\alpha,\beta)-winning set. If SS is (α,β)(\alpha,\beta)-winning for all (equiv. for all sufficiently small) β(0,1)\beta\in(0,1), we say that SS is an α\alpha-winning set. If SS is α\alpha-winning for some (equiv. for all sufficiently small) α(0,1)\alpha\in(0,1), we say that SS is winning. (To see that “for all” and “for some” may be replaced by “for all sufficiently small”, cf. [9, Lemmas 8 and 9].)

In what follows we shall need a variation of Schmidt’s game introduced by C. T. McMullen [6], the absolute winning game.222Technically, the game we describe below was not defined by McMullen, who only considered the special case X=\use@mathgroup\M@U\symAMSbRdX=\use@mathgroup\M@U\symAMSb R^{d}. The version of the absolute winning game described below appeared first in [3, §4]. Given β(0,1)\beta\in(0,1) and SXS\subseteq X, the (β,S)(\beta,S)-absolute game is played as follows: As before the game starts with Bob choosing a pair ω1=(x1,r1)Ω\omega_{1}=(x_{1},r_{1})\in\Omega, and Alice and Bob then take turns choosing pairs ωn\omega_{n} and ωn\omega^{\prime}_{n}. However, instead of requiring ωn+1sωnsωn\omega_{n+1}\leq_{s}\omega^{\prime}_{n}\leq_{s}\omega_{n}, now there is no restriction on Alice’s choice ωn=(xn,rn)\omega^{\prime}_{n}=(x^{\prime}_{n},r^{\prime}_{n}), and Bob’s choice ωn+1=(xn+1,rn+1)\omega_{n+1}=(x_{n+1},r_{n+1}) must be chosen to satisfy

(1.4) ωn+1sωn and d(xn,xn+1)rn+rn+1.\omega_{n+1}\leq_{s}\omega_{n}\text{ and }d(x^{\prime}_{n},x_{n+1})\geq r_{n}^{\prime}+r_{n+1}.

The second condition states that the balls ωn\omega^{\prime}_{n} and ωn+1\omega_{n+1} are “formally disjoint”, so we can think of Alice has having “deleted” the ball B(ωn)B(\omega^{\prime}_{n}). The restrictions on the radii in the absolute game are

(1.5) rnβrn and rn+1βrn.r_{n}^{\prime}\leq\beta r_{n}\text{ and }r_{n+1}\geq\beta r_{n}.

However, since (1.5) is insufficient to ensure that the diameters of the nested balls (1.3) tend to zero, it may happen that the intersection I=defnB(ωn)I{\,\stackrel{{\scriptstyle\mathrm{def}}}{{=}}\,}\bigcap_{n}B(\omega_{n}) is not a singleton. If this occurs, we call Alice the winner if ISI\cap S\neq{\diameter}; otherwise Bob is declared the winner. It may also happen that Bob has no legal moves, and for technical reasons in this case it is better to declare Alice the winner (cf. [2, comments before Lemma 2.1]). However, for sufficiently nice spaces (i.e. uniformly perfect spaces; cf. [3, Lemma 4.3]) and for sufficiently small β\beta, such a situation cannot arise. If Alice has a winning strategy for the β\beta-absolute game with a given target set SS, then SS is called β\beta-absolute winning, and if this is true for every β>0\beta>0, then SS is called absolute winning. Every absolute winning set on a uniformly perfect set is winning [3, Proposition 4.4(ii)].

The Banach–Mazur game’s rules are the same as for Schmidt’s game except for the fact that no restricting parameters are given, i.e., at each of the player’s turns, they may choose as small a radius as they please, and just like in the absolute winning game, if the intersection of the players’ balls is not a singleton, we declare Alice the winner if this intersection with the target set is nonempty. It is well-known that Alice has a winning strategy if and only if the target set is comeager [7].

Acknowledgements. The first-named author was supported in part by the Simons Foundation grant #245708. The third-named author was supported in part by the EPSRC Programme Grant EP/J018260/1. The authors thank the anonymous referee for helpful comments.

2. The Banach–Mazur–Schmidt and Banach–Mazur–McMullen games

We now define two new games: the Banach–Mazur–Schmidt (BMS) game and the Banach–Mazur–McMullen (BMM) game. In the BMS (resp. BMM) game, Bob starts, playing according to the Banach–Mazur game rules, while Alice is dealt a parameter β(0,1)\beta\in(0,1) and follows the rules for Schmidt’s game (resp. the absolute winning game). More precisely: in the β\beta-BMS game, Bob and Alice take turns choosing pairs ωn\omega_{n} and ωn\omega^{\prime}_{n} satisfying ωn+1sωnsωn\omega_{n+1}\leq_{s}\omega^{\prime}_{n}\leq_{s}\omega_{n}, while Alice’s choices are additionally required to satisfy (1.2). And in the β\beta-BMM game, Bob and Alice take turns choosing pairs ωn\omega_{n} and ωn\omega^{\prime}_{n} satisfying (1.4), but Bob’s moves are not required to satisfy (1.5) even though Alice’s are.

If Alice has a winning strategy for the β\beta-BMS (resp. β\beta-BMM) game, then we call the target set β\beta-BMS (resp. β\beta-BMM) winning, and if a set it β\beta-BMS (resp. β\beta-BMM) winning for all sufficiently small β(0,1)\beta\in(0,1), then we call it BMS-winning (resp. BMM-winning).

Our first theorem geometrically characterizes the β\beta-BMS and β\beta-BMM winning sets, but first we need the following definition:

Definition 2.1.

Fix β>0\beta>0. A set EXE\subseteq X is said to be uniformly β\beta-porous if there exists r0>0r_{0}>0 such that for every ball B(x,r)XB(x,r)\subseteq X with rr0r\leq r_{0}, there exists B(y,βr)B(x,r)B^{\circ}(y,\beta r)\subseteq B(x,r) such that B(y,βr)E=B^{\circ}(y,\beta r)\cap E={\diameter}.

Theorem 2.2.

Let (X,d)(X,d) be a separable complete metric space and fix β(0,1)\beta\in(0,1). Then a Borel set TXT\subseteq X is β\beta-BMS winning if and only if XTX\setminus T can be written as the countable union of uniformly β\beta-porous sets. Moreover, TT is β\beta-BMM winning if and only if XTX\setminus T is countable.

Example 2.3.

The Cantor set CC\subseteq\mathbb{R} is uniformly 1/51/5-porous,333Let B(x,r)\use@mathgroup\M@U\symAMSbRB(x,r)\subseteq\use@mathgroup\M@U\symAMSb R be a ball, and we will show that there exists B(y,r/5)B(x,r)B^{\circ}(y,r/5)\subseteq B(x,r) such that B(y,r/5)C=B^{\circ}(y,r/5)\cap C={\diameter}. By a zooming argument, we can without loss of generality assume that B(x,r)[0,1/3]B(x,r)\cap[0,1/3]\neq{\diameter} and B(x,r)[2/3,1]B(x,r)\cap[2/3,1]\neq{\diameter}. By a symmetry argument, we can without loss of generality assume that x1/2x\geq 1/2. If r/51/6r/5\leq 1/6, then we let y=1/2y=1/2, and if 3r/51/23r/5\geq 1/2, then we let y=1/2+4r/5y=1/2+4r/5. Either way we get B(y,r/5)[1/3,x+r]B(x,r)B^{\circ}(y,r/5)\subseteq[1/3,x+r]\subseteq B(x,r) and B(y,r/5)C=B^{\circ}(y,r/5)\cap C={\diameter}. so by Theorem 2.2, C\mathbb{R}\setminus C is 1/51/5-BMS winning.

A slightly more general example:

Example 2.4.

Given s0s\geq 0, a closed set KK is called Ahlfors ss-regular if there exists a measure μ\mu whose support equals KK and a constant C>0C>0 such that for all xKx\in K and 0<r10<r\leq 1,

C1rsμ(B(x,r))Crs.C^{-1}r^{s}\leq\mu(B(x,r))\leq Cr^{s}.

If KdK\subseteq\mathbb{R}^{d} is an Ahlfors ss-regular set with s<ds<d, or more generally if KXK\subseteq X is Ahlfors ss-regular, XX is Ahlfors δ\delta-regular, and s<δs<\delta, then a simple calculation shows that KK is uniformly porous,444Suppose otherwise, and let β>0\beta>0 be small. Then there exists a ball B(x,r)B(x,r) such that every ball B(y,βr)B(x,r)B^{\circ}(y,\beta r)\subseteq B(x,r) intersects KK. Let AB(x,r/2)A\subseteq B(x,r/2) be a maximal 3βr3\beta r-separated set. Since XX is Ahlfors δ\delta-regular, #(A)C1βδ\#(A)\geq C_{1}\beta^{-\delta} for some constant C1>0C_{1}>0. For each xAx\in A let f(x)Kf(x)\in K be chosen so that f(x)B(x,βr)f(x)\in B^{\circ}(x,\beta r). Then B={f(x):xA}B=\{f(x):x\in A\} is a βr\beta r-separated subset of B(x,r)KB(x,r)\cap K, so since KK is Ahlfors ss-regular, we have #(B)C2βs\#(B)\leq C_{2}\beta^{-s} for some constant C2>0C_{2}>0. Since #(A)=#(B)\#(A)=\#(B), this is a contradiction for sufficiently small β\beta. Thus KK is uniformly porous. An alternate proof of this fact may be found in [1, Lemma 3.12]. so by Theorem 2.2, T=XKT=X\setminus K is BMS-winning.

If XX is Ahlfors δ\delta-regular, then for every β\beta there exists sβ<δs_{\beta}<\delta such that every uniformly β\beta-porous set TT has upper box-counting dimension sβ\leq s_{\beta} [4, Theorem 4.7]. Since the Hausdorff and packing dimensions of a set are bounded above by its upper box dimension, we get the following corollary of Theorem 2.2:

Corollary 2.5.

If XX is Ahlfors δ\delta-regular and TXT\subseteq X is BMS-winning, then the Hausdorff and packing dimensions of XTX\setminus T are <δ<\delta.

The following can be proven either using Theorem 2.2 or by a method similar to [9, Theorem 2].

Corollary 2.6.

The intersection of countably many β\beta-BMS (resp. β\beta-BMM) winning sets is β\beta-BMS (resp. β\beta-BMM) winning.

Proof of Theorem 2.2.

(\Rightarrow): Suppose Alice has a winning strategy. By [9, Theorem 7], she has a positional winning strategy, i.e. a map ff which inputs a move of Bob and tells her what she should do next. Let \mathcal{B} denote the set of balls in XX, so that f:f:\mathcal{B}\to\mathcal{B} and if f(B(x,r))=B(y,s)f(B(x,r))=B(y,s), then s=βrs=\beta r. Let

g(B(x,r))={B(y,s)BMS gameB(x,r)B(y,s)BMM game.g(B(x,r))=\begin{cases}B^{\circ}(y,s)&\text{BMS game}\\ B^{\circ}(x,r)\setminus B(y,s)&\text{BMM game}\end{cases}.

For each mm\in\mathbb{N}, let

Km=XB=B(x,r)0<r1/mg(B).K_{m}=X\setminus\bigcup_{\begin{subarray}{c}B=B(x,r)\in\mathcal{B}\\ 0<r\leq 1/m\end{subarray}}g(B).
Claim 2.7.

XTmKmX\setminus T\subseteq\bigcup_{m\in\mathbb{N}}K_{m}.

Proof.

By contradiction, suppose p(XT)mKmp\in(X\setminus T)\setminus\bigcup_{m\in\mathbb{N}}K_{m}. We claim that Bob can beat Alice’s strategy by using the following counter-strategy: always choose a ball BB\in\mathcal{B} such that pg(B)p\in g(B). Obviously, if he can successfully apply this strategy then this is a contradiction, since then the intersection point will be pXTp\in X\setminus T, a win for Bob, but Alice’s strategy was supposed to be a winning strategy. We prove by induction that he can apply the strategy. If he applied it to choose his previous move BnB_{n}, then pg(Bn)p\in g(B_{n}), and from the definition of gg, this guarantees the existence of a neighborhood B(p,2/m)B(p,2/m) of pp such that any ball contained in B(p,2/m)B(p,2/m) constitutes a legal move for Bob. Such a neighborhood also exists if it is the first turn and no one has made a move. Now since pKmp\notin K_{m}, there exists B=B(x,r)B=B(x,r)\in\mathcal{B} with 0<r1/m0<r\leq 1/m and pg(B)p\in g(B). Then BB constitutes a legal move for Bob, since pg(B)Bp\in g(B)\subseteq B and thus BB(p,2/m)B\subseteq B(p,2/m). ∎

Now if Alice and Bob are playing the BMS game, then for every B=B(x,r)B=B(x,r)\in\mathcal{B} such that 0<r1/m0<r\leq 1/m, we have g(B)Km=g(B)\cap K_{m}={\diameter}, so by definition, KmK_{m} is uniformly β\beta-porous, completing the proof.

On the other hand, suppose that Alice and Bob are playing the BMM game. Then for every B=B(x,r)B=B(x,r)\in\mathcal{B} such that 0<rβ1/2/m0<r\leq\beta^{1/2}/m, we have g(B(x,β1/2r))Km=g(B(x,\beta^{-1/2}r))\cap K_{m}={\diameter}, so BKm=B(x2,β1/2r)KmB\cap K_{m}=B(x_{2},\beta^{1/2}r)\cap K_{m} for some x2Xx_{2}\in X. Continuing this process we get a sequence (xk)1(x_{k})_{1}^{\infty} with x1=xx_{1}=x such that

BKm=B(x2,β1/2r)Km=B(x3,βr)Km=B\cap K_{m}=B(x_{2},\beta^{1/2}r)\cap K_{m}=B(x_{3},\beta r)\cap K_{m}=\cdots

So diam(BKm)=0\operatorname{diam}(B\cap K_{m})=0 and thus BKmB\cap K_{m} is either empty or a singleton. Since XX is separable, this implies that KmK_{m} is countable, completing the proof.

(\Leftarrow): Suppose that Alice and Bob are playing the BMS game and that XT=1EnX\setminus T=\bigcup_{1}^{\infty}E_{n}, where each EnE_{n} is uniformly β\beta-porous. For each nn, Alice can avoid the set EnE_{n} in a finite number of moves as follows: make dummy moves until Bob’s radius is smaller than the r0r_{0} which occurs in Definition 2.1, then make the move B(y,βr)B(x,r)B(y,\beta r)\subseteq B(x,r), where xx, yy, and rr are as in Definition 2.1, then make one more move to avoid the set B(y,βr)B(y,βr)B(y,\beta r)\setminus B^{\circ}(y,\beta r). By avoiding each set EnE_{n} in turn, Alice can ensure that the intersection point is in TT.

On the other hand, suppose that Alice and Bob are playing the BMM game and that XTX\setminus T is countable. If (xn)1(x^{\prime}_{n})_{1}^{\infty} is an enumeration of XTX\setminus T, then let Alice’s nnth move be ωn=(xn,rn)\omega^{\prime}_{n}=(x^{\prime}_{n},r^{\prime}_{n}) for some legal rnr^{\prime}_{n}. This ensures that the intersection point is in TT. ∎

3. Application to Diophantine approximation

Recall that the exponent of irrationality of a vector 𝐱d\mathbf{x}\in\mathbb{R}^{d} is the number

ω(𝐱)=lim sup𝐩/qdlog𝐱𝐩/qlog(q),\omega(\mathbf{x})=\limsup_{\mathbf{p}/q\in\mathbb{Q}^{d}}\frac{-\log\|\mathbf{x}-\mathbf{p}/q\|}{\log(q)},

where the limsup is taken along any enumeration of d\mathbb{Q}^{d}. The set

{𝐱d:ω(𝐱)=1+1/d}\{\mathbf{x}\in\mathbb{R}^{d}:\omega(\mathbf{x})=1+1/d\}

is of full Lebesgue measure and is winning for Schmidt’s game, while the set

{𝐱d:ω(𝐱)=}\{\mathbf{x}\in\mathbb{R}^{d}:\omega(\mathbf{x})=\infty\}

is comeager, so it is winning for the Banach–Mazur game. A natural question is whether their union is winning for the BMS game. The following result shows that the answer is no:

Theorem 3.1.

Let ψ:(0,)\psi:\mathbb{N}\to(0,\infty) be a decreasing function such that q1+1/dψ(q)0q^{1+1/d}\psi(q)\to 0, and let

(3.1) S={𝐱d:0<lim inf𝐩/qd𝐱𝐩/qψ(q)<}.S=\left\{\mathbf{x}\in\mathbb{R}^{d}:0<\liminf_{\mathbf{p}/q\in\mathbb{Q}^{d}}\frac{\|\mathbf{x}-\mathbf{p}/q\|}{\psi(q)}<\infty\right\}.

Then for every β\beta, Bob has a strategy to ensure that the intersection point is in SS. In particular, XSX\setminus S is not BMS-winning.

Corollary 3.2.

The set (3.1) cannot be written as the union of countably many uniformly β\beta-porous sets for any 0<β<10<\beta<1.

Note that for any c>1+1/dc>1+1/d, the set {𝐱d:ω(𝐱)=c}\{\mathbf{x}\in\mathbb{R}^{d}:\omega(\mathbf{x})=c\} contains a set of the form (3.1), so it also cannot be written as the union of countably many uniformly β\beta-porous sets.

Proof.

Fix 0<β<10<\beta<1. For each 𝐩/qd\mathbf{p}/q\in\mathbb{Q}^{d}, write

B(𝐩/q)=B(𝐩/q,ψ(q)),B(𝐩/q)=B(𝐩/q,(1+6β1)ψ(q)).B(\mathbf{p}/q)=B(\mathbf{p}/q,\psi(q)),\;\;B^{\prime}(\mathbf{p}/q)=B(\mathbf{p}/q,(1+6\beta^{-1})\psi(q)).

We will give a strategy for Bob to force the intersection point to lie in infinitely many of the sets B(𝐩/q)B^{\prime}(\mathbf{p}/q), but only finitely many of the sets B(𝐩/q)B(\mathbf{p}/q). Accordingly, we fix Q0Q_{0}\in\mathbb{N} large to be determined, and we call a ball A=B(𝐱,r)dA=B(\mathbf{x},r)\subseteq\mathbb{R}^{d} good if for every 𝐩/qd\mathbf{p}/q\in\mathbb{Q}^{d} such that AB(𝐩/q)A\cap B(\mathbf{p}/q)\neq{\diameter} and qQ0q\geq Q_{0}, we have

(3.2) ψ(q)r/3.\psi(q)\leq r/3.

Intuitively, if AA is a good ball then Bob should still be able to win and avoid all of the sets B(𝐩/q)B(\mathbf{p}/q), after Alice has just played AA.

Claim 3.3.

If A=B(𝐱,r)A=B(\mathbf{x},r) is a good ball, then there exists a ball B=B(𝐲,s)AB=B(\mathbf{y},s)\subseteq A such that BB(𝐩0/q0)B\subseteq B^{\prime}(\mathbf{p}_{0}/q_{0}) for some 𝐩0/q0d\mathbf{p}_{0}/q_{0}\in\mathbb{Q}^{d} with ψ(q0)<r\psi(q_{0})<r, and such that for every 𝐩/qd\mathbf{p}/q\in\mathbb{Q}^{d} such that BB(𝐩/q)B\cap B(\mathbf{p}/q)\neq{\diameter} and qQ0q\geq Q_{0}, we have

ψ(q)βs/3.\psi(q)\leq\beta s/3.

In other words, if Alice’s previous choice was good, then Bob can move so that Alice’s next choice must be good, while at the same time moving sufficiently close to a rational point.

Proof.

By the Simplex Lemma [5, Lemma 4], there exists an affine hyperplane d\mathcal{L}\subseteq\mathbb{R}^{d} such that for all 𝐩/qdB(𝐱,2r)\mathbf{p}/q\in\mathbb{Q}^{d}\cap B(\mathbf{x},2r)\setminus\mathcal{L}, we have

(3.3) qcdrd/(d+1),q\geq c_{d}r^{-d/(d+1)},

where cd>0c_{d}>0 is a constant depending on dd. Choose a ball A~=B(𝐱~,r/3)A𝒩(,r/3)\widetilde{A}=B(\widetilde{\mathbf{x}},r/3)\subseteq A\setminus\mathcal{N}(\mathcal{L},r/3), where 𝒩(,t)={𝐱d:d(𝐱,)t}\mathcal{N}(\mathcal{L},t)=\{\mathbf{x}\in\mathbb{R}^{d}:d(\mathbf{x},\mathcal{L})\leq t\} is the closed tt-thickening of \mathcal{L}. Note that for all 𝐩/qd\mathbf{p}/q\in\mathbb{Q}^{d}, if A~B(𝐩/q)\widetilde{A}\cap B(\mathbf{p}/q)\neq{\diameter} and qQ0q\geq Q_{0}, then AB(𝐩/q)A\cap B(\mathbf{p}/q)\neq{\diameter}, so by (3.2), d(𝐩/q,A~)ψ(q)r/3d(\mathbf{p}/q,\widetilde{A})\leq\psi(q)\leq r/3. Thus 𝐩/q𝒩(A~,r/3)B(𝐱,2r)\mathbf{p}/q\in\mathcal{N}(\widetilde{A},r/3)\subseteq B(\mathbf{x},2r)\setminus\mathcal{L}, so (3.3) holds.

Let 𝐩0/q0d\mathbf{p}_{0}/q_{0}\in\mathbb{Q}^{d} be chosen to minimize q0q_{0}, subject to the conditions A~B(𝐩0/q0)\widetilde{A}\cap B(\mathbf{p}_{0}/q_{0})\neq{\diameter} and q0Qq_{0}\geq Q. Let s=3β1ψ(q0)s=3\beta^{-1}\psi(q_{0}). Since q1+1/dψ(q)0q^{1+1/d}\psi(q)\to 0, if Q0Q_{0} is sufficiently large then (3.3) implies s<r/3s<r/3. Thus there exists a ball B=B(𝐲,s)B~B=B(\mathbf{y},s)\subseteq\widetilde{B} such that BB(𝐩0/q0)B\cap B(\mathbf{p}_{0}/q_{0})\neq{\diameter} and thus BB(𝐩0/q0)B\subseteq B^{\prime}(\mathbf{p}_{0}/q_{0}). Now if 𝐩/qd\mathbf{p}/q\in\mathbb{Q}^{d} satisfies BB(𝐩/q)B\cap B(\mathbf{p}/q)\neq{\diameter} and qQ0q\geq Q_{0}, then qq0q\geq q_{0}, so

ψ(q)ψ(q0)=βs/3.\psi(q)\leq\psi(q_{0})=\beta s/3.\qed

By choosing Q0Q_{0} sufficiently large, we can guarantee that the ball B(𝟎,1)B(\mathbf{0},1) is good. Let Bob’s strategy consist of responding to Alice’s moves AA with the balls BB given in Claim 3.3, letting A=B(𝟎,1)A=B(\mathbf{0},1) for the first move. Then by induction, Alice’s moves will always be good, which implies that the intersection point 𝐳\mathbf{z} is not contained in any of the balls B(𝐩/q)B(\mathbf{p}/q). But by construction, 𝐳\mathbf{z} is contained in infinitely many balls B(𝐩0/q0)B(\mathbf{p}_{0}/q_{0}). Thus 𝐳S\mathbf{z}\in S, where SS is as in (3.1). ∎

A natural point of comparison for the exponent of irrationality function is the Lagrange spectrum function

L(𝐱)=lim inf𝐩/qd𝐱𝐩/qq1+1/dL(\mathbf{x})=\liminf_{\mathbf{p}/q\in\mathbb{Q}^{d}}\frac{\|\mathbf{x}-\mathbf{p}/q\|}{q^{1+1/d}}\cdot

While the condition ω(𝐱)>1+1/d\omega(\mathbf{x})>1+1/d is equivalent to 𝐱\mathbf{x}’s being very well approximable, the condition L(𝐱)>0L(\mathbf{x})>0 is equivalent to 𝐱\mathbf{x}’s being badly approximable. We have shown above that the levelsets of the exponent of irrationality function cannot be written as the countable union of uniformly β\beta-porous sets for any 0<β<10<\beta<1. To contrast this we prove:

Theorem 3.4.

For all dd\in\mathbb{N} and 0<ε<10<\varepsilon<1, the set WAd(ε):={𝐱d:L(𝐱)ε}\mathrm{WA}_{d}(\varepsilon):=\{\mathbf{x}\in\mathbb{R}^{d}:L(\mathbf{x})\leq\varepsilon\} is β\beta-BMS winning, where β=(ε/3)d+1(0,1/2)\beta=(\varepsilon/3)^{d+1}\in(0,1/2).

Corollary 3.5.

The set BAd(ε)=dWAd(ε)\mathrm{BA}_{d}(\varepsilon)=\mathbb{R}^{d}\setminus\mathrm{WA}_{d}(\varepsilon) can be written as the union of countably many uniformly β\beta-porous sets.

Proof.

Alice’s strategy will be as follows: move near a rational point 𝐩n/qn\mathbf{p}_{n}/q_{n}, then make a move disjoint from 𝐩n/qn\mathbf{p}_{n}/q_{n}, then wait long enough so that Bob’s move Bn=B(𝐱n,rn)B_{n}=B(\mathbf{x}_{n},r_{n}) satisfies

rn<2(3/ε)3(qnd(Bn,𝐩n/qn))d+1,r_{n}<2(3/\varepsilon)^{3}(q_{n}d(B_{n},\mathbf{p}_{n}/q_{n}))^{d+1},

then repeat. So suppose that Bob has just made the move B=B(𝐱,r)=B(𝐱n,rn)B=B(\mathbf{x},r)=B(\mathbf{x}_{n},r_{n}), and we will show how Alice can move near a new rational point 𝐩n+1/qn+1\mathbf{p}_{n+1}/q_{n+1}. Let QQ be the unique number such that r=2(3/ε)d/Q1+1/dr=2(3/\varepsilon)^{d}/Q^{1+1/d}. By Dirichlet’s theorem, there exists 𝐩/q=𝐩n+1/qn+1\mathbf{p}/q=\mathbf{p}_{n+1}/q_{n+1} with qQq\leq Q such that

(3.4) 𝐱𝐩q1qQ1/d\left\|\mathbf{x}-\frac{\mathbf{p}}{q}\right\|\leq\frac{1}{qQ^{1/d}}\cdot

Note that this inequality implies that 𝐩n+1/qn+1𝐩m/qm\mathbf{p}_{n+1}/q_{n+1}\neq\mathbf{p}_{m}/q_{m} for all mnm\leq n.

Case 1. 𝐱𝐩qr/2\left\|\mathbf{x}-\frac{\mathbf{p}}{q}\right\|\leq r/2. In this case, we have B(𝐩/q,βr)BB(\mathbf{p}/q,\beta r)\subseteq B. On the other hand,

βr=2(3/ε)dβQ1+1/d2(3/ε)dβq1+1/dεq1+1/d\beta r=\frac{2(3/\varepsilon)^{d}\beta}{Q^{1+1/d}}\leq\frac{2(3/\varepsilon)^{d}\beta}{q^{1+1/d}}\leq\frac{\varepsilon}{q^{1+1/d}}\cdot

So the move B(𝐩/q,βr)B(\mathbf{p}/q,\beta r) will bring Alice sufficiently close to the rational point 𝐩/q\mathbf{p}/q.

Case 2. 𝐱𝐩qr/2\left\|\mathbf{x}-\frac{\mathbf{p}}{q}\right\|\geq r/2. In this case, by (3.4) we have

1qQ1/dr2=(3/ε)dQ1+1/d,\frac{1}{qQ^{1/d}}\geq\frac{r}{2}=\frac{(3/\varepsilon)^{d}}{Q^{1+1/d}},

and rearranging gives

q(ε/3)dQ.q\leq(\varepsilon/3)^{d}Q.

Thus

r2𝐱𝐩qε3q1+1/d,\frac{r}{2}\leq\left\|\mathbf{x}-\frac{\mathbf{p}}{q}\right\|\leq\frac{\varepsilon}{3q^{1+1/d}},

and in particular

𝐱𝐩q+rεq1+1/d\left\|\mathbf{x}-\frac{\mathbf{p}}{q}\right\|+r\leq\frac{\varepsilon}{q^{1+1/d}}\cdot

It follows that BB(𝐩/q,ε/q1+1/d)B\subseteq B(\mathbf{p}/q,\varepsilon/q^{1+1/d}), so any move Alice makes will bring her sufficiently close to the rational point 𝐩/q\mathbf{p}/q. ∎

References

  • [1] M. Bonk, J. Heinonen, and S. Rohde, Doubling conformal densities, J. Reine Angew. Math. 541 (2001), 117–141.
  • [2] R. Broderick, L. Fishman, and D. S. Simmons, Badly approximable systems of affine forms and incompressibility on fractals, J. Number Theory 133 (2013), no. 7, 2186–2205.
  • [3] L. Fishman, D. S. Simmons, and M. Urbański, Diophantine approximation and the geometry of limit sets in Gromov hyperbolic metric spaces, http://arxiv.org/abs/1301.5630, preprint 2013.
  • [4] E. Järvenpää, M. Järvenpää, A. Käenmäki, T. Rajala, S. Rogovin, and V. Suomala, Packing dimension and Ahlfors regularity of porous sets in metric spaces, Math. Z. 266 (2010), no. 1, 83–105.
  • [5] S. Kristensen, R. Thorn, and S. L. Velani, Diophantine approximation and badly approximable sets, Advances in Math. 203 (2006), 132–169.
  • [6] C. T. McMullen, Winning sets, quasiconformal maps and Diophantine approximation, Geom. Funct. Anal. 20 (2010), no. 3, 726–740.
  • [7] J. C. Oxtoby, The Banach-Mazur game and Banach category theorem, Contributions to the theory of games, vol. 3, Annals of Mathematics Studies, no. 39, Princeton University Press, Princeton, N.J., 1957, pp. 159–163.
  • [8] J. P. Revalski, The Banach-Mazur game: History and recent developments, http://www1.univ-ag.fr/aoc/activite/revalski/Banach-Mazur_Game.pdf.
  • [9] W. M. Schmidt, On badly approximable numbers and certain games, Trans. Amer. Math. Soc. 123 (1966), 27–50.
  • [10] R. Telgársky, Topological games: on the 50th anniversary of the Banach-Mazur game, Rocky Mountain J. Math. 17 (1987), no. 2, 227–276.