Hostname: page-component-8448b6f56d-gtxcr Total loading time: 0 Render date: 2024-04-24T19:33:45.313Z Has data issue: false hasContentIssue false

ERDŐS–LIOUVILLE SETS

Published online by Cambridge University Press:  03 October 2022

TABOKA PRINCE CHALEBGWA
Affiliation:
The Fields Institute for Research in Mathematical Sciences, 222 College Street, Toronto, Ontario MST 3J1, Canada e-mail: taboka@aims.ac.za
SIDNEY A. MORRIS*
Affiliation:
School of Engineering, IT and Physical Sciences, Federation University Australia, PO Box 663, Ballarat, Victoria 3353, Australia and Department of Mathematical and Physical Sciences, La Trobe University, Melbourne, Victoria 3086, Australia
Rights & Permissions [Opens in a new window]

Abstract

In 1844, Joseph Liouville proved the existence of transcendental numbers. He introduced the set $\mathcal L$ of numbers, now known as Liouville numbers, and showed that they are all transcendental. It is known that $\mathcal L$ has cardinality $\mathfrak {c}$, the cardinality of the continuum, and is a dense $G_{\delta }$ subset of the set $\mathbb {R}$ of all real numbers. In 1962, Erdős proved that every real number is the sum of two Liouville numbers. In this paper, a set W of complex numbers is said to have the Erdős property if every real number is the sum of two numbers in W. The set W is said to be an Erdős–Liouville set if it is a dense subset of $\mathcal {L}$ and has the Erdős property. Each subset of $\mathbb {R}$ is assigned its subspace topology, where $\mathbb {R}$ has the euclidean topology. It is proved here that: (i) there exist $2^{\mathfrak {c}}$ Erdős–Liouville sets no two of which are homeomorphic; (ii) there exist $\mathfrak {c}$ Erdős–Liouville sets each of which is homeomorphic to $\mathcal {L}$ with its subspace topology and homeomorphic to the space of all irrational numbers; (iii) each Erdős–Liouville set L homeomorphic to $\mathcal {L}$ contains another Erdős–Liouville set $L'$ homeomorphic to $\mathcal {L}$. Therefore, there is no minimal Erdős–Liouville set homeomorphic to $\mathcal {L}$.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2022. Published by Cambridge University Press on behalf of Australian Mathematical Publishing Association Inc.

1 Introduction

It has been known for over 175 years that every Liouville number is transcendental and for 120 years that the set $\mathcal {L}$ of Liouville numbers is uncountable. Notwithstanding this, the set $\mathcal {L}$ is known to have Lebesgue measure zero. So in this sense, $\mathcal {L}$ is very small. Therefore, it is surprising that each real number equals the sum of two Liouville numbers. It is reasonable to ask if $\mathcal {L}$ is the smallest set, in some sense, with this property. In this paper, it is proved that there is an uncountable number of sets smaller than $\mathcal {L}$ which have this property. Indeed, there are $2^{\mathfrak {c}}$ such subsets of $\mathcal {L}$ no two of which are homeomorphic as subspaces of $\mathbb {R}$ .

2 Preliminaries

Remark 2.1. In 1844, Joseph Liouville proved the existence of transcendental numbers [Reference Angell2, Reference Baker3]. He introduced the set $\mathcal L$ of real numbers, now known as Liouville numbers, and showed that they are all transcendental. A real number x is said to be a Liouville number if for every positive integer n, there exists a pair of integers $(p,q)$ with $q>1$ such that

$$ \begin{align*} 0< \bigg\lvert x-\frac{p}{q}\bigg\rvert<\frac{1}{q^n}. \end{align*} $$

This definition of a Liouville number can be reformulated as follows. For a given irrational x, let $p_k/q_k=p_k(x)/q_k(x)$ , where $q_k(x)>0$ , denote the sequence of convergents of the continued fraction expansion of x; then for every $n\in \mathbb {N}$ , there are infinitely many k such that $q_{k+1}>q_k^n$ . A more restrictive class of Liouville numbers is obtained by requiring this inequality to hold for every $k> N=N(n)\in \mathbb {N}$ . Such numbers are called strong Liouville numbers.

In 1962, Erdős [Reference Erdős8] proved that every real number is the sum of two Liouville numbers (and also the product of two Liouville numbers). He gave two proofs. One was a constructive proof. The other proof used the fact that the set $\mathcal {L}$ of all Liouville numbers is a dense $G_{\delta }$ -set in $\mathbb {R}$ and showed that every dense $G_{\delta }$ -set in $\mathbb {R}$ has this property.

Definition 2.2. A set W of complex numbers is said to have the Erdős property if every real number is a sum of two numbers in W.

Remark 2.3. Recall that if A and B are subsets of the set $\mathbb {C}$ of all complex numbers, then the sum-set is defined to be $A+B=\{a+b:a\in A, b\in B\}$ . So the subset W of $\mathbb {C}$ has the Erdős property if the sum-set $W+W$ contains the set $\mathbb {R}$ . (See [Reference Brown, Yin, Pollington and Moran4, Reference Pollington, Pollington and Moran13].)

Remark 2.4. By the theorem proved by Erdős mentioned above, the set $\mathcal {L}$ of all Liouville numbers has the Erdős property.

Remark 2.5. If W is a set with the Erdős property, then every set containing W also has the Erdős property.

Definition 2.6. A set W is said to be an Erdős–Liouville set if it has the Erdős property and is a dense subset of the set $\mathcal {L}$ of Liouville numbers.

Remark 2.7. It is not immediately obvious that there exist any Erdős–Liouville sets other than the set $\mathcal {L}$ itself. It is known that some sets of positive Lebesgue measure have the Erdős property, but they are not subsets of $\mathcal {L}$ as the set $\mathcal {L}$ is known to have measure zero. (See, for example, [Reference Chalebgwa and Morris5].) According to Petruska, [Reference Petruska12], Erdős asked if the set of strong Liouville numbers has the Erdős property. However, Petruska [Reference Petruska12] proved that it does not. He did this by showing that the sum of two strong Liouville numbers is either a Liouville number or a rational number. Hence, the sum of two strong Liouville numbers cannot equal any irrational number other than a Liouville number. However, it is proved in [Reference Conman and Poletsky7], in the text following Corollary 1.4 and in Section 3, that there does exist another Erdős–Liouville set. In [Reference Marques and Moreira10], the set of ultra-Liouville numbers is introduced and it is shown that this set is a dense $G_{\delta }$ -subset of $\mathcal {L}$ which is therefore an Erdős–Liouville set.

Remark 2.8. In the literature, there are various strengthenings of the Erdős result on Liouville numbers. We mention explicitly [Reference Alniaçik and Saias1, Reference Rieger14, Reference Schwarz15]. The paper [Reference Kumar, Thangadurai and Waldschmidt9] shows that the set of Liouville numbers has a property stronger than the Erdős property. Though we do not study such properties, we record here that the $\mathfrak {c}$ Erdős–Liouville sets we produce in Theorem 4.6 also possess this stronger property, while Theorem 3.6 and the proof of Theorem 4.6 show that there are only $\mathfrak {c}$ dense $G_{\delta }$ subsets of $\mathbb {R}$ . The relevant theorem from [Reference Kumar, Thangadurai and Waldschmidt9] describing this stronger property is the following result.

Theorem 2.9. Let G be a dense $G_{\delta }$ -subset of $\mathbb {R}$ , I an interval in $\mathbb {R}$ with nonempty interior, and f a continuous function from I to $\mathbb {R}$ which is nowhere locally constant. (This means that f is not constant on any nonempty open subinterval of I.) Then there exists an $x\in G\cap I$ such that $f(x)\in G$ . Indeed, there is an uncountable number of such x.

If we put $f(x)=r-x$ , for $r,x\in \mathbb {R}$ and $I=\mathbb {R}$ , we see that f satisfies the conditions of the theorem and thus G has the Erdős property. However, as observed in [Reference Kumar, Thangadurai and Waldschmidt9], if we put $I=(0,\sqrt {r})$ and $f(x) = \sqrt {r-x^2}$ , we see that for every Erdős–Liouville set G, every positive real number is the sum of two squares of numbers in G. Also, the argument in [Reference Kumar, Thangadurai and Waldschmidt9, pages 63–64] with $L^1= \{\exp (\alpha ): \alpha \in \mathcal {L}\}$ leads to the observation that $L^1\cap {\mathcal {L}}$ is an Erdős–Liouville set. Although it was not explicitly mentioned in [Reference Kumar, Thangadurai and Waldschmidt9], it follows by induction that if $L^n= L^{n-1}\cap \mathcal {L}$ , for $n\in \mathbb {N}$ , $n>1$ , then each $L^n$ is an Erdős–Liouville set. However, we do not know if the sets $L^n$ are distinct from each other and distinct from $\mathcal {L}$ .

Proposition 2.10. Let S be a set of real numbers such that $W_1\supset S\supset W_2$ , where $W_1$ and $W_2$ are Erdős–Liouville sets. Then S is an Erdős–Liouville set.

Proof. As $S\supset W_2$ , by Remark 2.5, it has the Erdős property. Also as $W_2$ is dense in $\mathbb {R}$ , so too is S. Finally, as $S\subset W_1$ , it is a subset of $\mathcal {L}$ . Therefore, S is an Erdős–Liouville set.

3 Some topology

Before proving the existence of an uncountable number of Erdős–Liouville sets, we need to record some topology, some of which was laid bare in [Reference Chalebgwa and Morris5, Reference Chalebgwa and Morris6, Reference Morris11].

Definition 3.1. A topological space X is said to be topologically complete (or completely metrisable) if the topology of X is the same as the topology induced by a complete metric on X.

Of course, every complete metric space is topologically complete.

We denote by $\mathbb {P}$ the set of all irrational real numbers with the topology it inherits as a subspace of the euclidean space $\mathbb {R}$ .

A beautiful characterisation of the topological space $\mathbb {P}$ is given in [Reference van Mill16, Theorem 1.9.8].

Theorem 3.2. The space of all irrational real numbers $\mathbb {P}$ is topologically the unique nonempty, separable, metrisable, topologically complete, nowhere locally compact, and zero-dimensional space. $\Box $

This has a Corollary 3.3, [Reference van Mill16, Corollary 1.9.9], which is often proved using continued fractions.

Corollary 3.3. The space $\mathbb {P}$ is homeomorphic to the Tychonoff product $\mathbb {N}^{\aleph _0}$ of a countably infinite number of homeomorphic copies of the discrete space $\mathbb {N}$ of positive integers. Hence, $\mathbb {P}\times \mathbb {P}$ is homeomorphic to $\mathbb {P}$ . Indeed, $\mathbb {P}$ is homeomorphic to $\mathbb {P}^{\aleph _0}$ .

Remark 3.4. Recall that a subset X of a topological space Y is said to be a $G_{\delta }$ -set if it is a countable intersection of open sets in Y while X is said to be an $F_{\sigma }$ -set if it is a countable union of closed sets in Y. Obviously, a subset X of a topological space Y is a $G_{\delta }$ -set if and only if its complement is an $F_{\sigma }$ -set. We see immediately that in a metric space such as $\mathbb {R}$ , the set $\mathcal {T}$ of all transcendental real numbers is a $G_{\delta }$ -set as its complement is the countably infinite set $\mathbb {A}$ of all real algebraic numbers.

Now we connect the notion of $G_{\delta }$ -set in $\mathbb {R}$ to the property of being topologically complete.

Theorem 3.5 [Reference van Mill16, Theorem A.63].

A subset of a separable metric topologically complete space is a $G_{\delta }$ -set in that space if and only if it is topologically complete.

Using Theorems 3.2, 3.5 and Corollary 3.3, we obtain the following result.

Theorem 3.6. Every $G_{\delta }$ subset of the set $\mathbb {P}$ of all irrational real numbers is homeomorphic to $\mathbb {P}$ and to $\mathbb {N}^{\aleph _0}$ . In particular, the space $\mathcal {T}$ of all real transcendental numbers and the space $\mathcal {L}$ of all Liouville numbers, with their subspace topologies from $\mathbb {R}$ , are both homeomorphic to $\mathbb {P}$ and to $\mathbb {N}^{\aleph _0}$ .

These results and a similar one [Reference van Mill16, Theorem 1.9.6] characterising the space $\mathbb {Q}$ of all rational numbers with its euclidean topology, are used in [Reference Chalebgwa and Morris6, Reference Morris11] to describe transcendental groups and topological transcendental fields.

4 The existence of $2^{\mathfrak {c}}$ Erdős–Liouville sets

Theorem 4.1. Let X be a topological space homeomorphic to $\mathbb {P}$ . Then X has a dense $G_{\delta }$ -set Y which is homeomorphic to $\mathbb {P}$ such that the cardinality of the set $X\setminus Y$ is $\mathfrak {c}$ , the cardinality of the continuum.

Proof. Consider the topological space $\mathcal {T}$ of all real transcendental numbers and the topological space $\mathcal {L}$ of all Liouville numbers. We saw in Corollary 2.6 and Remark 2.1 that $\mathcal {L}$ is a dense $G_{\delta }$ -set, and $\mathcal {T}$ and $\mathcal {L}$ are homeomorphic to $\mathbb {P}$ . Further, the cardinality of the set $\mathcal {T}\setminus \mathcal {L}$ is $\mathfrak {c}$ . As the properties of being a dense $G_{\delta }$ -set and having cardinality $\mathfrak {c}$ are preserved by homeomorphisms, the theorem is proved.

By Theorem 4.1 and Remark 2.1, we have the following corollary.

Corollary 4.2. The space $\mathcal {L}$ of all Liouville numbers has a dense $G_{\delta }$ -set $L_1 $ homeomorphic to $\mathcal {L}$ . Further, $L_1$ is an Erdős–Liouville set.

Theorem 4.3. If L is any Erdős–Liouville set homeomorphic to $\mathbb {P}$ , then it has a proper subset $L_1$ which is an Erdős–Liouville set homeomorphic to $\mathbb {P}$ . Therefore, there is no minimal Erdős–Liouville set homeomorphic to $\mathbb {P}$ . $\Box $

Our next theorem follows immediately from Corollary 4.2 and Theorem 4.1.

Theorem 4.4. There exist Erdős–Liouville sets $L_1, L_2,\ldots , L_n, \ldots, $ for $n\in \mathbb {N}$ , such that

$$ \begin{align*} \mathcal{L}\supset L_1\supset L_2\supset\cdots\supset L_n\supset\cdots \end{align*} $$

with each $L_n\setminus L_{n+1}$ having cardinality $\mathfrak {c}$ and each $L_{n+1}$ a $G_{\delta }$ -set in $L_n$ which is homeomorphic to $\mathbb {P}$ . $\Box $

Theorem 4.5. There exist $2^{\mathfrak {c}}$ Erdős–Liouville sets no two of which are homeomorphic.

Proof. First, we note that there are precisely $2^{\mathfrak {c}}$ subsets of the set $\mathcal {L}$ of all Liouville numbers as $\mathcal {L}$ has cardinality $\mathfrak {c}$ . So the cardinality of the set of Erdős–Liouville sets is not greater than $2^{\mathfrak {c}}$ .

Using the notation of Theorem 4.4, let W be any subset of $\mathcal {L}\setminus L_1$ . As $L_1$ is an Erdős–Liouville set and $L_1\subset \mathcal {L}$ , Remark 2.5 implies that $L_1\cup W$ is an Erdős–Liouville set. As there are $2^{\mathfrak {c}}$ subsets W of the set $\mathcal {L}\setminus L_1$ , it follows that there are $2^{\mathfrak {c}}$ distinct Erdős–Liouville sets. So it remains to show only that amongst these, there are $2^{\mathfrak {c}}$ no two of which are homeomorphic.

By the Laverentieff theorem, [Reference van Mill16, Theorem A8.5], there are at most $\mathfrak {c}$ subspaces of $\mathbb {R}$ which are homeomorphic. As there are $2^{\mathfrak {c}}$ distinct Erdős–Liouville sets, it follows that there are $2^{\mathfrak {c}}$ Erdős–Liouville sets no two of which are homeomorphic, as required.

Theorem 4.6. There exist $\mathfrak {c}$ Erdős–Liouville sets each of which is homeomorphic to $\mathcal {L}$ with its subspace topology. So each is homeomorphic to $\mathbb {P}$ .

Proof. Using the notation of Theorem 4.4, $\mathcal {L}\supset L_1$ , and the set $\mathcal {L}\setminus L_1$ has cardinality $\mathfrak {c}$ . Let $S= \{s_1, s_2, \ldots , s_n, \ldots \}$ be any countably infinite subset of $\mathcal {L}\setminus L_1$ . As $\mathcal {L}\setminus L_1$ has cardinality $\mathfrak {c}$ , there are $\mathfrak {c}$ distinct such subsets S. Then, $\mathcal {L}\setminus S= \bigcap _{i=1}^\infty (\mathcal {L}\setminus \{s_i\})$ .

Observing that $\mathcal {L}\supset \mathcal {L}\setminus S\supset L_1$ , Proposition 2.10 implies that each $\mathcal {L}\setminus S$ is an Erdős–Liouville set.

Noting that $\mathcal {L}$ is a $G_{\delta }$ -set in $\mathbb {R}$ , and each $\mathcal {L}\setminus \{s_i\}$ is an open set in $\mathcal {L}$ , it follows that $\mathcal {L}\setminus S$ is a $G_{\delta }$ -set. By Theorem 3.6, each of the $\mathfrak {c}$ sets $\mathcal {L}\setminus S$ is therefore homeomorphic to $ \mathcal {L}$ and $\mathbb {P}$ .

Footnotes

The first author’s research is supported by the Fields Institute for Research in Mathematical Sciences, via the Fields-Ontario postdoctoral Fellowship.

References

Alniaçik, K. and Saias, E., ‘Une remarque sur les G ${}_{\delta }$ -denses’, Arch. Math. 62 (1994), 425426.CrossRefGoogle Scholar
Angell, D., Irrationality and Transcendence in Number Theory (CRC Press, New York, 2022).Google Scholar
Baker, A., Transcendental Number Theory (Cambridge University Press, Cambridge, 1975).CrossRefGoogle Scholar
Brown, G. and Yin, Q.-H., ‘Some metric properties of sum sets’, in: Number Theory with an Emphasis on the Markhoff Spectrum (eds. Pollington, A. D. and Moran, W.) (Marcel Dekker Inc., New York, 1993), 1722.Google Scholar
Chalebgwa, T. P. and Morris, S. A., ‘A continuous homomorphism of a thin set onto a fat set’, Bull. Aust. Math. Soc., to appear. Published online (1 April 2022).CrossRefGoogle Scholar
Chalebgwa, T. P. and Morris, S. A., ‘Topological transcendental fields’, Axioms 11 (2022), Article no. 118; doi:10.3390/axioms11030118.CrossRefGoogle Scholar
Conman, D. and Poletsky, E., ‘Polynomial estimates, exponential curves and Diophantine approximation’, Math. Res. Lett. 17 (2010), 11251136.CrossRefGoogle Scholar
Erdős, P., ‘Representation of real numbers as sums and products of Liouville numbers’, Michigan Math. J. 9 (1962), 5960.10.1307/mmj/1028998621CrossRefGoogle Scholar
Kumar, K. S., Thangadurai, R. and Waldschmidt, M., ‘Liouville numbers and Shanuel’s conjecture’, Arch. Math. 102 (2014), 5970.10.1007/s00013-013-0606-0CrossRefGoogle Scholar
Marques, D. and Moreira, C. G., ‘On a variant of a question posed by K. Mahler concerning Liouville numbers’, Bull. Aust. Math. Soc. 91 (2015), 2933.10.1017/S0004972714000471CrossRefGoogle Scholar
Morris, S. A., ‘Transcendental groups’, Preprint, 2021, arXiv:2112.12450.Google Scholar
Petruska, G., ‘On strong Liouville numbers’, Indag. Math. (N.S.) 3 (1992), 211218.10.1016/0019-3577(92)90010-ICrossRefGoogle Scholar
Pollington, A. D., ‘Some metric properties of sum sets’, in: Number Theory with an Emphasis on the Markhoff Spectrum (eds. Pollington, A. D. and Moran, W.) (Marcel Dekker Inc., New York, 1993), 207214.Google Scholar
Rieger, G. J., ‘Über die Löbarkeit von Gleichungssystemen durch Liouville–Zahlen’, Arch. Math. 26 (1975), 4043.10.1007/BF01229700CrossRefGoogle Scholar
Schwarz, W., ‘Liouville Zahlen und der Satz von Baire’, Math.-Phys. Semesterber. 24 (1977), 8487.Google Scholar
van Mill, J., The Infinite-Dimensional Topology of Function Spaces (Elsevier, Amsterdam, 2001).Google Scholar