Fact-checked by Grok 2 weeks ago

Banach–Tarski paradox

The Banach–Tarski paradox is a in set-theoretic geometry asserting that a three-dimensional solid ball can be partitioned into a finite number of disjoint, non-measurable subsets, which can then be reassembled using only rigid motions—rotations and translations—to form two solid balls each identical in size and shape to the original. This result, proved by mathematicians and in their 1924 paper "Sur la décomposition des ensembles de points en parties respectivement congruentes," relies fundamentally on the from , which enables the construction of these highly pathological, non-measurable pieces that defy intuitive notions of volume preservation. The paradox arises because the decomposition violates everyday expectations about physical duplication: one cannot perform such a partitioning with measurable sets or using scissors and glue, as the pieces lack well-defined volumes under Lebesgue measure, rendering the theorem inapplicable to tangible matter. Subsequent refinements have shown that five pieces suffice for the doubling, and this is the minimal number required, though the original proof did not specify the exact count beyond finiteness. The theorem generalizes to higher dimensions but fails in one or two dimensions, highlighting the peculiar role of three-dimensional rotations in enabling paradoxical decompositions via free groups of rotations. Beyond its counterintuitive nature, the Banach–Tarski paradox has profound implications for measure theory, , and the foundations of , illustrating how the leads to counterexamples that challenge classical intuitions while underscoring the distinction between equidecomposability (via isometries) and measure equivalence. It has inspired ongoing research into measurable versions, amenable choice principles, and applications in , though it remains a cornerstone example of why accepting the yields "paradoxical" yet logically consistent results in infinite settings.

Historical Background

Banach and Tarski Publication

The Banach–Tarski paradox was first published in 1924 in the Polish mathematical journal Fundamenta Mathematicae, volume 6, pages 244–277, under the title "Sur la décomposition des ensembles de points en parties respectivement congruentes" by and . , a born on March 30, 1892, in , had begun his career with early contributions to , including work on monotone functions and integrations, and was on the cusp of pioneering developments in through his association with the Lwów School of Mathematics. Alfred Tarski, born on January 14, 1901, in , served as a key collaborator, bringing his emerging expertise in and logic; he had recently completed his doctorate at the in 1924, making him one of the institution's youngest recipients. In the paper, Banach and Tarski proved that any two bounded subsets of with nonempty interior, such as balls of different radii, are equidecomposable using finitely many pieces and isometries of the space, allowing one ball to be partitioned and reassembled into two identical copies of itself. They paraphrased the core result as follows: any two balls with different radii can be decomposed into the same finite number of disjoint, respectively congruent parts. The publication, appearing when Banach was 32 and Tarski 23, garnered immediate recognition for its counterintuitive implications, highlighting the non-measurable sets arising from the and defying classical notions of volume conservation in .

Earlier Works and Influences

The foundations of the Banach–Tarski paradox trace back to the rapid development of in the early , building on Georg Cantor's transfinite cardinalities and from the and , which highlighted paradoxes in infinite sets and prompted rigorous axiomatization. A key milestone was Zermelo's 1904 introduction of the in his proof of the , which allowed selection of elements from infinitely many sets without a explicit rule, but immediately raised concerns about its consistency and applications to and . In the 1910s, explorations of non-measurable sets emerged as a direct consequence of these foundational debates, revealing limitations in extending Lebesgue measure to all subsets of the real line. Giuseppe Vitali constructed the first explicit non-Lebesgue measurable set in 1905 using the axiom of choice, partitioning the real line into equivalence classes under rational translations to form the Vitali set that defied additive measure properties. Wacław Sierpiński advanced this in 1918 by proving that the axiom of choice implies the existence of non-measurable subsets of the unit interval, independent of the continuum hypothesis, through a construction involving well-ordering the reals and selecting representatives from Vitali-like classes. The most immediate precursor was Felix Hausdorff's 1914 paradox, which applied these ideas to geometric decompositions on . In Grundzüge der Mengenlehre, Hausdorff demonstrated that the unit sphere in three dimensions, minus of points, admits a paradoxical decomposition into three disjoint subsets A, B, and C such that rigid rotations map A onto the entire and B \cup C onto the entire . This result relied implicitly on the to select non-measurable pieces and exploited the fact that specific rotations—such as 180° about one and 120° about a perpendicular —generate a free subgroup of rank two in the rotation group SO(3), enabling the paradoxical equidecomposability without fixed points except on . Hausdorff's spherical paradox directly influenced the extension to solid balls by highlighting the role of free group actions in producing non-intuitive equidecomposabilities, though the ball's interior and center required additional techniques to handle radial issues. Early group-theoretic insights, including properties of explored in the 1910s and early 1920s, provided the algebraic framework for such constructions, paving the way for integrating rotations with paradoxical set decompositions.

Mathematical Foundations

Formal Statement

The Banach–Tarski paradox concerns the equidecomposability of bounded sets with nonempty interior in Euclidean space of dimension at least three. Specifically, consider the unit ball B = \{ x \in \mathbb{R}^3 : \|x\| \leq 1 \}, which is a bounded set containing interior points. Two subsets A, C \subseteq \mathbb{R}^3 are equidecomposable (via isometries) if there exist finite partitions A = \bigsqcup_{i=1}^n A_i and C = \bigsqcup_{i=1}^n C_i into pairwise disjoint sets, together with isometries g_i of \mathbb{R}^3 (i.e., distance-preserving bijections, such as compositions of rotations and translations, belonging to the group SO(3) \ltimes \mathbb{R}^3), such that g_i(A_i) = C_i for each i = 1, \dots, n. Equidecomposability implies that the sets can be decomposed into finitely many pieces and reassembled into each other using rigid motions without stretching or tearing. The formal statement of the paradox is the following theorem: In \mathbb{R}^3, the unit ball B is equidecomposable to two disjoint copies of itself; that is, there exists a finite n, a partition of B into disjoint sets A_1, \dots, A_n, and isometries g_1, \dots, g_n such that the g_i(A_i) are pairwise disjoint and their union is the disjoint union of two unit balls. This decomposition yields two balls each identical to the original, yet the pieces do not preserve volume in the Lebesgue sense because each A_i is non-Lebesgue measurable. The result extends to any ball in \mathbb{R}^3 of the same radius, regardless of its center, as translations and rotations are isometries. The construction relies on the axiom of choice to select the non-measurable pieces.

Axiom of Choice Role

The (AC), first explicitly formulated by in 1904, asserts that for any collection of nonempty sets, there exists a choice function that selects exactly one element from each set. This axiom was introduced to prove the , enabling every set to be well-ordered, but it immediately sparked controversy among mathematicians. Critics such as , René-Louis Baire, and argued that AC lacked intuitive justification and could lead to non-constructive proofs without explicit methods for making the choices. Despite the debates, AC became a standard axiom in Zermelo-Fraenkel with choice (ZFC), underpinning many results in modern mathematics. In the context of the Banach–Tarski paradox, AC plays an indispensable role by enabling the construction of the required paradoxical decomposition. Specifically, the proof relies on AC to select a set of representatives from the cosets of a certain free subgroup within the rotation group acting on the sphere, which generates non-measurable sets essential for the decomposition. Without AC, these selections cannot be guaranteed in a constructive manner, and the resulting pieces would not exhibit the necessary paradoxical properties. This dependence highlights how AC facilitates the creation of pathological sets that defy intuitive notions of volume preservation under rigid motions. The necessity of AC is further evidenced by results in without choice. In Zermelo-Fraenkel set theory (ZF) alone, it is consistent that no such paradoxical decomposition of the unit ball exists, as demonstrated in Robert M. Solovay's 1970 model where ZF is augmented with the axiom of dependent choice but all sets of reals are Lebesgue measurable. In this model, the absence of non-measurable sets prevents the Banach–Tarski construction, underscoring that the paradox is not a theorem of ZF but requires the full strength of AC. These foundational aspects reveal AC's profound implications for and measure , illustrating its capacity to yield counterintuitive outcomes like duplication through finite partitions. The paradox thus serves as a stark example of how AC extends classical mathematics beyond measurable phenomena, prompting ongoing philosophical and mathematical scrutiny of its acceptance.

Proof Construction

High-Level Overview

The Banach–Tarski paradox asserts that a solid ball in three-dimensional can be partitioned into finitely many disjoint subsets, which can then be rigidly moved—via rotations and translations—to form two complete balls identical in size and shape to the original. This result, while mathematically rigorous, appears deeply counterintuitive because it contradicts everyday notions of ; however, the pieces involved are non-measurable sets, lacking a well-defined , which allows them to bypass the additivity property of . The core idea of the proof leverages the group of rotations in three dimensions, denoted SO(3), to achieve a paradoxical decomposition first on the unit sphere S^2. A pivotal insight is that SO(3) contains a free subgroup of rank 2, isomorphic to the on two generators F_2, consisting of two independent rotations whose compositions enable "duplication" through specific word equations in the group. This permits the sphere to be broken into pieces that, when rotated appropriately, reassemble into two copies of the sphere. To extend this to the solid ball, the decomposition is applied radially along rays emanating from the center (excluding the center itself), transforming the spherical pieces into conical sectors that fill the ball minus the origin. The center point, being a set of measure zero, is assigned arbitrarily to one of the resulting balls without affecting the overall construction. A standard construction achieves the paradoxical decomposition of the ball with five pieces, and it has been shown by Raphael M. Robinson in 1947 that five is the minimal number required.

Step 1: Group Actions on the Sphere

The rotation group SO(3), consisting of all $3 \times 3 orthogonal matrices with $1, acts on the unit sphere S^2 = { x \in \mathbb{R}^3 : |x| = 1 }by [matrix multiplication](/page/Matrix_multiplication), preserving the sphere's [geometry](/page/Geometry) through rigid rotations. This action is continuous and transitive onS^2$, excluding fixed points of individual rotations, which form countable sets. The F_2 on two generators a and b is the with \langle a, b \mid \rangle, where elements are finite reduced words in a^{\pm 1} and b^{\pm 1}, and the group operation is concatenation followed by reduction to eliminate inverses. This structure allows F_2 to act freely on itself by left , leading to paradoxical decompositions intrinsic to its non-amenable nature. To embed F_2 into SO(3), consider rotations \alpha and \beta around perpendicular axes, such as the x-axis and y-axis, each by the angle \theta = \arccos(1/3) \approx 70.53^\circ. The subgroup generated by \alpha and \beta is isomorphic to F_2, as the irrational multiple of \pi in \theta ensures no unexpected relations beyond free group axioms, with the action on S^2 being free except at countable fixed-point sets. This embedding enables a paradoxical action: there exists a word w \in F_2 such that the equation w F_2 = F_2 \sqcup a F_2 holds, where \sqcup denotes disjoint union, allowing duplication of the group via finitely many pieces under left multiplication. Hausdorff first utilized such a free subgroup of SO(3) in 1914 to establish the paradoxical decomposability of S^2 into finitely many congruent pieces reassemblable into two copies of itself, excluding countable point sets.

Step 2: Paradoxical Decomposition

The action of the subgroup G \cong F_2 of \mathrm{SO}(3), generated by two specific rotations \sigma and \tau, on the sphere S^2 fixes exactly two points, the north and south poles Y, which lie on the common rotation axis of \sigma and \tau. The induced action on S^2 \setminus Y is , meaning that for any non-identity g \in G and x \in S^2 \setminus Y, g x \neq x. Consequently, S^2 \setminus Y partitions into disjoint orbits \{G x \mid x \in S^2 \setminus Y\}, where each orbit G x = \{g x \mid g \in G\} is in bijection with G via the action. By the , select a transversal X \subset S^2 \setminus Y containing exactly one representative from each , ensuring S^2 \setminus Y = \bigsqcup_{g \in G} g X, a . This identifies S^2 \setminus Y with G up to the G-action, transferring paradoxical properties from G to the sphere. The set X cannot be chosen measurably, rendering the resulting pieces non-measurable with respect to the on S^2. The free group G \cong F_2 on generators \sigma, \tau admits a paradoxical decomposition derived from its word structure. The non-identity elements of G partition into four disjoint sets based on the first letter of their reduced words: A = W(\sigma^{-1}) (words starting with \sigma^{-1}), B = W(\sigma) (starting with \sigma), C = W(\tau^{-1}) (starting with \tau^{-1}), and D = W(\tau) (starting with \tau). These sets satisfy the relations G = B \sqcup \sigma A and G = D \sqcup \tau C, where the unions are disjoint; for instance, \sigma A = G \setminus B because left multiplication by \sigma maps words starting with \sigma^{-1} to those not starting with \sigma, due to the reduced word property preventing cancellation that would alter the initial letter inappropriately. The identity element e can be adjoined to B (or any set) without disrupting the relations, as it corresponds to a negligible singleton orbit. Transferring to the sphere yields four disjoint pieces A' = \bigsqcup_{g \in A} g X, B' = \bigsqcup_{g \in B} g X, C' = \bigsqcup_{g \in C} g X, D' = \bigsqcup_{g \in D} g X, whose union is S^2 \setminus Y. The relations on G imply B' \sqcup \sigma A' = S^2 \setminus Y and D' \sqcup \tau C' = S^2 \setminus Y, both disjoint unions, achieved by applying the rotations \sigma and \tau to the pieces. Thus, the four pieces rearrange via elements of \mathrm{SO}(3) into two disjoint copies of S^2 \setminus Y, establishing the paradoxical decomposition. The non-measurability of A', B', C', D' follows directly from that of X, as the unions preserve the pathological selection.

Step 3: Handling the Radius and Center

To extend the paradoxical from the unit sphere S^2 to the full unit B = \{ x \in \mathbb{R}^3 : \|x\| \le 1 \}, the interior points are addressed through radial extensions along rays from the . Specifically, for a of S^2 into finitely many pieces A_1, \dots, A_n that are equidecomposable via rotations (isometries fixing the ), the punctured B_0 = B \setminus \{0\} is partitioned into corresponding radial pieces P_i = \{ r a : a \in A_i, \, 0 < r \le 1 \} for i = 1, \dots, n. These P_i form "solid cones" or sectors filling B_0, preserving the equidecomposability because rotations in SO(3) map rays from the to rays, thus sending P_i to P_j whenever A_i maps to A_j. The origin \{0\} presents a special case, as it lies on every ray and is fixed by all rotations fixing the origin. This singleton set has Lebesgue measure zero and can be arbitrarily assigned to any one of the pieces, say P_1 \cup \{0\}, without altering the paradoxical property, since adding a measure-zero set does not affect the non-measurability or the congruence mappings in the decomposition. This assignment ensures the full ball B is decomposed into the same finite number of pieces as the sphere—specifically, five pieces in the standard construction—maintaining the total count without requiring additional partitions. This radial approach handles varying radii inherently, as each covers all distances from near the to the , allowing the to scale uniformly across the ball's interior. For isometries not fixing the (such as translations), the preservation of equidecomposability holds because the ray structure is invariant under origin-fixing maps used in the core decomposition, while translations can be applied post-decomposition to reposition entire pieces without disrupting the radial integrity. The challenge of the fixed under rotations is thus resolved by its separate, negligible treatment, avoiding any need to decompose it further.

Step 4: Reassembly into Two Balls

The reassembly phase of the Banach–Tarski paradox utilizes isometries of three-dimensional —specifically rotations and translations—to rearrange the five decomposed pieces of the unit ball into two disjoint unit balls of the same size as the original. One ball is positioned at the origin, and the other is shifted to the point (3,0,0) to ensure disjointness. This step relies on the paradoxical properties established in the decomposition of , extending them to the ball by applying group elements from the rotation group SO(3). Explicitly, the five pieces, labeled P_1 through P_5, are assigned as follows: in the standard five-piece construction, one of the four pieces from the spherical decomposition is further partitioned into two to handle the fixed points appropriately. Pieces corresponding to certain combinations are mapped to the at the using rotations from the free subgroup generators σ and a conjugate τ' = ρ τ ρ^{-1} (where ρ is a suitable to adjust axes), with angles consistent with (1/3) to maintain freeness. The remaining pieces are mapped to the at (3,0,0) via similar rotations followed by a by the (3,0,0). These mappings ensure that the images cover the two target balls exactly without gaps or overlaps. Since each transformation is an , every piece remains congruent to its preimage in the original decomposition, preserving the geometric structure without distortion or scaling. The reassembled pieces fill the two balls without gaps or overlaps, as the disjoint union of the images equals the two target balls, and the original pieces partition the initial ball completely, leaving no material unused. This completeness arises directly from the equidecomposability established in prior steps. The pieces themselves are highly irregular, non-Lebesgue measurable sets with fractal-like complexity, constructed via the , rendering the reassembly visually and intuitively incomprehensible—far removed from any physical dissection of tangible objects.

Extensions and Variations

Infinite Copies from One Ball

The Banach–Tarski paradox, which decomposes a into finitely many pieces to reassemble into two identical copies, generalizes to produce any finite number n \geq 2 of copies from a single using a similar finite . This extension relies on free groups of higher into the special SO(3), which acts on the sphere via rotations. Specifically, a free group of k admits a paradoxical that allows partitioning a set into pieces reassemblable into $2^k copies, and by adjusting the rank and combining decompositions, any finite n copies can be obtained with finitely many pieces overall. One method to achieve this for arbitrary finite n involves chaining duplications iteratively: start with the two-copy , apply it to each resulting copy to double again, and repeat finitely many times, yielding $2^m copies after m steps with a total number of that is the product of the counts at each stage, remaining finite. For numbers not powers of 2, higher-rank subgroups enable direct decompositions into n copies by exploiting the non-amenable structure of these groups, where no finitely additive measure exists, preventing volume preservation. This contrasts with amenable groups, such as abelian groups, which lack such paradoxical decompositions and admit means, ensuring that no finite partitioning can yield multiple copies without scaling. The construction extends further to infinitely many copies by leveraging the fact that the free group of rank 2 contains free subgroups of countably infinite rank, allowing a decomposition of the ball into countably infinitely many pieces that reassemble into countably infinitely many unit balls. Although this requires infinitely many pieces, the process remains finite in "steps" in the sense of the group action's structure, amplifying the paradoxical intuition of creating volume from no net addition. However, this phenomenon fails in one or two dimensions for balls, as the rotation groups SO(1) (trivial) and SO(2) (abelian, hence amenable) contain no non-abelian free subgroups, precluding any such paradoxical decompositions.

Von Neumann Paradox in 2D

In 1929, established that the \mathbb{R}^2 admits no finite paradoxical decomposition using isometries, meaning it cannot be partitioned into finitely many pieces that can be reassembled via rotations and translations into two copies of itself. This result contrasts sharply with the three-dimensional case of the Banach–Tarski paradox, where such a finite decomposition is possible. Von Neumann's proof hinges on the structure of the of the plane, which is the \mathrm{SO}(2) \ltimes \mathbb{R}^2, where \mathrm{SO}(2) denotes the group of rotations around the origin. The key insight is that \mathrm{SO}(2) is abelian and commutative, implying that it contains no free non-abelian of 2, which is essential for constructing paradoxical decompositions in the manner used for \mathrm{SO}(3) in three dimensions. Specifically, the of \mathrm{SO}(2) is trivial, preventing the generation of the independent rotations needed to "duplicate" sets through disjoint orbits, as occurs in higher dimensions. This algebraic limitation ensures that any attempt to mimic the Banach–Tarski construction in two dimensions fails for finite pieces, preserving the existence of a finitely additive, - and rotation-invariant measure on all subsets of the plane (up to the ). introduced the concept of amenability in this work to characterize groups admitting such invariant means, showing that amenable groups like the two-dimensional avoid finite paradoxical decompositions. Although finite decompositions are impossible with isometries, paradoxical decompositions become feasible with countably infinitely many pieces. For instance, the plane can be partitioned into countably many that, through translations alone (a of isometries), can be reassembled into two full copies of the plane. This relies on the to select a suitable basis for \mathbb{R}^2 as a over \mathbb{Q}, allowing the plane to be expressed as a countable disjoint union of translates that can be rearranged to cover twice the original area without overlap. Such countable paradoxes highlight the role of infinity in circumventing the amenability obstruction present for finite cases, though they lack the striking counterintuitive nature of finite-piece versions due to the involvement of uncountably many "exotic" sets. Von Neumann's 1929 analysis, building directly on the 1924 Banach–Tarski result, underscored the dimensional dependence of these phenomena, emphasizing how the abelian nature of \mathrm{SO}(2) deprives two dimensions of the non-abelian "richness" that enables the three-dimensional paradox.

Higher Dimensions and Other Spaces

The Banach–Tarski paradox extends to Euclidean balls in \mathbb{R}^n for all n \geq 3. In these dimensions, the special orthogonal group SO(n) contains a free non-abelian subgroup of rank 2, which is non-amenable and enables a paradoxical decomposition of the unit ball into finitely many pieces that can be reassembled into two copies of itself using isometries. This construction mirrors the three-dimensional case but leverages rotations in higher-dimensional spheres to generate the necessary free subgroup action. In contrast, the paradox fails in one dimension. The isometry group of \mathbb{R} is amenable (isomorphic to the semidirect product \mathbb{R} \rtimes \mathbb{Z}/2\mathbb{Z}), preventing any finite paradoxical decomposition of line segments; any equidecomposability must preserve Lebesgue measure additively. More generally, paradoxical decompositions arise in any space whose isometry group is non-amenable. Alfred Tarski proved that a group admits a paradoxical decomposition if and only if it is non-amenable, providing a characterization that applies to actions on sets. Thus, bounded sets with nonempty interior in such spaces can be paradoxically decomposed using group actions. Extensions appear in non-Euclidean settings, including spaces where the Isom(\mathbb{H}^n) for n \geq 2 is non-amenable and contains free subgroups, allowing paradoxical decompositions of hyperbolic balls. Similarly, smooth manifolds with non-amenable fundamental groups admit such decompositions via their deck transformations. Stan Wagon details further generalizations to spheres S^n for n \geq 2 (excluding fixed points).

Modern Developments

Minimal Piece Counts

The Tarski number associated with the Banach–Tarski paradox refers to the minimal integer k such that the unit ball in \mathbb{R}^3 can be partitioned into k non-measurable pieces that, using isometries (rotations and translations), can be reassembled to form two unit balls identical to the original. In three dimensions, this Tarski number for the ball is 5, meaning five pieces suffice and fewer do not. The original 1924 proof by Banach and Tarski established the existence of a finite decomposition but relied on a construction effectively requiring a large number of pieces (on the order of thousands) to extend the paradoxical decomposition from the sphere to the full ball while handling radial directions via the axiom of choice. This was reduced to six pieces in subsequent refinements, but Raphael M. Robinson proved in 1947 that five is both achievable and the minimum for the ball. For the unit sphere S^2 itself (excluding the interior), the corresponding minimal decomposition requires only four pieces to duplicate it into two copies via rotations in SO(3). The extra piece needed for the ball arises from the need to account for the fixed point and the orbits under the , which the sphere decomposition does not encounter. In general, for the special orthogonal group SO(n) acting on the (n-1)-sphere in dimensions n ≥ 3, the Tarski number is 4, as these groups contain a free non-abelian subgroup on two generators, allowing a paradoxical decomposition with the minimal four pieces; however, exact minimality for the corresponding ball in higher dimensions remains open, with known bounds growing linearly with n in some constructions. These refinements highlight the finite yet counterintuitive nature of the paradox, preserving its core surprise—that volume is not preserved under such non-measurable partitions—while minimizing the complexity of the decomposition.

Recent Proofs and Refinements

In 2022, a significant refinement was achieved by demonstrating that the Banach–Tarski paradox holds with a decomposition into at most six pieces, relying on the Hahn–Banach theorem rather than the full axiom of choice (AC). This proof operates within the framework of ZF set theory augmented by the axiom of dependent choice (DC), avoiding the stronger assumptions typically required for paradoxical decompositions. The construction leverages the Hahn–Banach extension theorem to separate certain orbits under the action of the free group on two generators, ensuring the pieces are non-measurable but explicitly definable without invoking AC's full power. The second edition of The Banach–Tarski Paradox by Grzegorz Tomkowicz and Stan Wagon, published in 2016, incorporates several post-2000 advancements, including new proofs for variants of the in different geometric settings and extended discussions on extensions. It addresses gaps in earlier treatments by providing detailed constructions for paradoxical decompositions in spaces with additional structure, such as those involving invariant means, and explores aspects through analyses of effective versions of the in descriptive . The book confirms that while explicit, constructive proofs remain challenging due to the non-measurable nature of the pieces, progress has been made in bounding the complexity of such decompositions without relying on heavy choice principles. Progress on Tarski numbers—the minimal number of pieces required for paradoxical decompositions in non-amenable groups—has seen confirmations but no major breakthroughs since the early . A 2020 discussion on MathOverflow summarizes for general groups, noting that Tarski numbers less than 4 are impossible for free groups, with 4, 5, and 6 confirmed as achievable in certain cases, while questions for other groups and higher-dimensional balls persist. These bounds apply to groups like SO(3), reinforcing earlier results. A mechanically verified proof of the Banach–Tarski theorem was formalized in 2022 using the ACL2(r) theorem prover, filling gaps in explicit constructions by providing a computer-checked decomposition of the unit ball into a finite number of pieces (52 in their explicit construction). This work highlights computability challenges, as the non-measurable sets involved resist algorithmic description, yet it verifies the paradox's validity under standard ZFC assumptions. Lower bounds for the number of pieces in higher dimensions, such as beyond three, continue to be open problems, with no definitive resolutions despite ongoing research into group actions on spheres. A 2025 preprint further explores minimal decompositions for forming n copies of the unit ball, proving that 3n-1 pieces suffice and are necessary in R^3 (confirming 5 for n=2).

References

  1. [1]
    [PDF] Sur la décomposition des ensembles de points
    Sur la décomposition des ensembles de points en parties respectivement congruentes*. Par. S. Banach et A. Tarski. Nous étudions dans cette Note les notions de ...Missing: équinombreux | Show results with:équinombreux
  2. [2]
    [PDF] The Banach-Tarski Paradox - MIT
    May 17, 2007 · In the late 19th century, Georg Cantor was the first to formally investigate this question, thus founding the study of set theory as a ...
  3. [3]
  4. [4]
    [PDF] the banach-tarski paradox
    Aug 12, 2008 · It has also been proven that the paradox can be accomplished with as few as five pieces. So we have examined the Banach-Tarski Paradox inside ...
  5. [5]
    [PDF] The Banach-Tarski Paradox
    The Banach-Tarski Paradox is a famous theorem about the equivalence of sets. In this paper, using Karl Stromberg's version of the proof in [1] as a guide, ...
  6. [6]
  7. [7]
    [PDF] A Comprehensive Analysis of the Banach-Tarski Paradox
    1.1.1 History of Banach-Tarski paradox. The Banach-Tarski paradox was first stated in 1924. The idea of it is, as acknowledged by Banach and Tarski, based on ...
  8. [8]
    Stefan Banach (1892 - 1945) - Biography - MacTutor
    On 7 April 1922, by resolution of the Faculty Council, Dr Stefan Banach received his habilitation for a Docent of Mathematics degree. He was appointed Professor ...
  9. [9]
    Alfred Tarski (1901 - 1983) - Biography - MacTutor
    Alfred Tarski made important contributions in many areas of mathematics, including metamathematics, set theory, measure theory, model theory, and general ...
  10. [10]
    On Decomposition of Point Sets into Respectively Congruent Parts ...
    It appeared in volume 6 of the journal Fundamenta Mathematicae. This is its first translation. Its best-known result is often called the Banach–Tarski paradox: ...Missing: original | Show results with:original
  11. [11]
    [2108.05714] The Banach-Tarski Paradox - arXiv
    Aug 8, 2021 · The aim of the paper is to provide a comprehensive proof of the Banach-Tarski paradox, expanding in between the lines of the original volume.Missing: primary | Show results with:primary
  12. [12]
    The Axiom of Choice - Stanford Encyclopedia of Philosophy
    Jan 8, 2008 · In 1904 Ernst Zermelo formulated the Axiom of Choice (abbreviated as ... Building on the work of Hausdorff, Banach and Tarski derive from AC their ...Origins and Chronology of the... · Mathematical Applications of... · Bibliography
  13. [13]
    [PDF] Non-measurable sets - Universiteit Leiden
    Jul 23, 2018 · In this text we will cover the non-measurable sets of G. Vitali, W. Sierpinski,. E.B. Van Vleck and F. Bernstein, which are all constructed in a ...Missing: 1910s | Show results with:1910s
  14. [14]
    [PDF] The Hausdorff Paradox
    Hausdorff gave the first such construction in 1914; he showed that if φ and ρ are rotations through 180. ◦ and 120. ◦. , respectively, about axes containing ...
  15. [15]
    Sur la décomposition des ensembles de points en parties ... - EUDML
    Sur la décomposition des ensembles de points en parties respectivement congruentes · Volume: 6, Issue: 1, page 244-277 · ISSN: 0016-2736 ...Missing: équinombreux | Show results with:équinombreux
  16. [16]
    [PDF] THE AXIOM OF CHOICE
    Zermelo's purpose in introducing AC was to establish a central principle of Cantor's set theory, namely, that every set admits a well-ordering and so can also ...
  17. [17]
    The Banach-Tarski Paradox - Cambridge University Press
    'In 1985 Stan Wagon wrote The Banach-Tarski Paradox, which not only became ... Chapter 13 - The Role of the Axiom of Choice. pp 207-221. You have access ...
  18. [18]
    The axiom of choice and Banach-Tarski paradoxes
    We shall use the axiom of choice to prove an extremely wimpy version of the Banach Tarski paradox, to wit: Theorem. It is possible to take a subset of the ...
  19. [19]
    A Model of Set-Theory in Which Every Set of Reals is Lebesgue ...
    (1) The principle of dependent choice (= DC, cf. III. 2.7.) (2) Every set of reals is Lebesgue measurable (LM). (3) Every set of reals has the property of ...
  20. [20]
    Introduction (Chapter 1) - The Banach–Tarski Paradox
    We refer to the Banach–Tarski Paradox on duplicating spheres or balls, which is often stated in the following fanciful form: a pea may be taken apart into ...
  21. [21]
    The Banach–Tarski Paradox - Cambridge University Press
    'In 1985 Stan Wagon wrote The Banach-Tarski Paradox, which not only became the classic text on paradoxical mathematics, but also provided vast new areas for ...
  22. [22]
    The Banach-Tarski Paradox: Duplicating Spheres and Balls
    The Banach-Tarski Paradox - May 1985. ... Stan Wagon. Show author details. Stan Wagon: Affiliation: Macalester College, Minnesota. Chapter. Chapter; Accessibility.
  23. [23]
    [PDF] The Banach-Tarski Paradox
    May 3, 2017 · The heart of the proof is the fact that SO(3) contains a free group on two generators as discussed above. We proceed in several steps. • [Step 1] ...Missing: embedding | Show results with:embedding
  24. [24]
    [PDF] the banach-tarski paradox - UChicago Math
    Jan 9, 2015 · The Banach-Tarski paradox is a theorem which states that the solid unit ball can be partitioned into a finite number of pieces, which can then ...
  25. [25]
    Higher Dimensions (Chapter 6) - The Banach–Tarski Paradox
    Higher Dimensions · Grzegorz Tomkowicz, Stan Wagon, Macalester College, Minnesota; Book: The Banach–Tarski Paradox; Online publication: 05 June 2016; Chapter ...
  26. [26]
    The Tarski numbers of groups - ScienceDirect
    ... paradoxical decomposition if and only if it is non-amenable. The minimal ... paradoxical decompositions will naturally arise in the proof of Theorem A(b).
  27. [27]
    The Hahn–Banach theorem and a six-piece paradoxical ...
    Oct 25, 2021 · In this paper, we prove that it is at most six: Theorem (HB). The Banach–Tarski paradox holds using six pieces. 2. Key lemma. The key to the ...
  28. [28]
    Latest progress on Tarski numbers - MathOverflow
    Jul 29, 2020 · Tarski numbers <4 are forbidden, which suggests that 7 is the answer to the question "What is the smallest non negative integer that we do not ...Equivalence of the Banach–Tarski paradox - MathOverflowIs there any version of the Banach-Tarski paradox in ZF?More results from mathoverflow.net
  29. [29]
    A Complete, Mechanically-Verified Proof of the Banach-Tarski ...
    This paper presents a formal proof of the Banach-Tarski theorem in ACL2(r). The Banach-Tarski theorem states that a unit ball can be partitioned into a finite ...