Fact-checked by Grok 2 weeks ago

Inscribed square problem

The inscribed square problem, also known as the square peg problem or Toeplitz' , is an open question in plane geometry that conjectures every simple closed continuous curve—a curve—in the contains four points forming the vertices of a square. Proposed by German mathematician Otto Toeplitz in 1911 during a conference talk on , the problem asks whether such an inscribed square exists for any curve, regardless of its shape or smoothness, as long as it is a non-self-intersecting loop dividing the plane into an interior and exterior region. Toeplitz himself proved the conjecture for convex curves, but a full general proof eluded him, and the problem has resisted resolution for over a century despite extensive efforts. Early progress included Arnold Emch's 1916 proof for piecewise analytic curves and Lev Schnirelmann's 1929 result for curves with bounded , later refined for C^2- curves. In 1989, Stromquist established the existence for all C^1- curves using topological methods involving the Brouwer . The conjecture holds for polygonal curves, as shown by work on finite total without cusps, and for locally curves. More recent advances have extended the result to certain non-smooth cases, such as curves that are unions of two graphs with constant less than 1, proven by in 2017 via an approach that leverages conserved quantities like the \int y \, dx. Despite these partial successes, the general case for arbitrary continuous curves remains unsolved, with challenges arising from potential degenerate configurations and the lack of differentiability. The problem has inspired related questions, such as the inscribed rectangle problem, which was affirmatively resolved in 2020 for all smooth curves using .

Fundamentals

Problem Statement

The inscribed square problem, also known as the square peg problem, conjectures that every Jordan curve in the plane admits an inscribed square. A Jordan curve is defined as a continuous simple closed curve in \mathbb{R}^2, which is topologically equivalent to the unit circle S^1 and thus embeds the circle into the plane without self-intersections. This conjecture was posed by Otto Toeplitz in 1911. An inscribed square on such a curve consists of four distinct points on the that serve as the vertices of , with the sides of the square connecting consecutive points in the order they appear along the curve's parametrization. Equivalently, the square peg problem asks whether every simple closed in the plane contains a non-degenerate with all four vertices lying on the . Formally, consider a Jordan parametrized by \gamma: [0,1] \to \mathbb{R}^2 such that \gamma(0) = \gamma(1) and \gamma is injective on (0,1). The asserts the existence of parameters t_1 < t_2 < t_3 < t_4 in [0,1) where the points \gamma(t_1), \gamma(t_2), \gamma(t_3), and \gamma(t_4) form the vertices of a square, meaning the vectors \gamma(t_2) - \gamma(t_1), \gamma(t_3) - \gamma(t_2), \gamma(t_4) - \gamma(t_3), and \gamma(t_1) - \gamma(t_4) (with the last adjusted for closure) have equal lengths and adjacent sides are perpendicular.

Illustrative Examples

One of the simplest examples is the circle, where inscribed squares abound. Consider the unit circle parametrized by \gamma(\theta) = (\cos \theta, \sin \theta) for \theta \in [0, 2\pi). For any angle \theta, the points \gamma(\theta), \gamma(\theta + \pi/2), \gamma(\theta + \pi), and \gamma(\theta + 3\pi/2) form the vertices of an inscribed square, with side length \sqrt{2} and diagonal equal to the diameter 2. These squares can be rotated arbitrarily, yielding infinitely many distinct inscribed squares. To construct such a square geometrically, draw a diameter of the circle, then erect a perpendicular bisector through the center to intersect the circle at two additional points; connecting these four points yields the square. Ellipses also admit inscribed squares, though typically unique up to symmetry and not aligned with the principal axes unless the ellipse is a circle. For instance, in a non-circular ellipse, the inscribed square's vertices lie at points where the curve intersects lines of equal parametric advance adjusted for the eccentricity, ensuring equal side lengths and right angles. This existence stems from the ellipse's smoothness and affinity to the circle, where affine transformations map inscribed parallelograms, but specific square configurations persist due to the curve's convexity and differentiability. Among regular polygons, the square itself serves as a trivial example: the boundary curve is precisely an inscribed square. For an equilateral triangle, inscribed squares exist in multiple orientations; for example, one configuration places two adjacent vertices on one side of the triangle and the other two on the remaining sides, maximizing area when the square's base aligns symmetrically with the triangle's base. Construction involves erecting auxiliary squares outwardly on the sides and finding intersections of lines from the opposite vertex to the new points, yielding the inscribed square's position. There are three such maximal squares, one per side. A simple non-convex example is a star-shaped Jordan curve, such as a smooth, non-self-intersecting dimpled limaçon (e.g., r = 1 + 0.75 \cos \theta in polar coordinates), which is star-shaped with respect to an interior point and admits an inscribed square. For star-shaped C^2-curves, existence of inscribed cyclic quadrilaterals, including squares, follows from known results for smooth curves. To find inscribed squares in these cases conceptually, consider pairs of points on the curve as potential opposite vertices; their midpoint must lie on the curve's axis of symmetry or the intersection of perpendicular bisectors of adjacent sides. For the circle, this reduces to intersecting perpendicular diameters. Pseudocode for a symmetric case like the circle might proceed as: select initial \theta; compute points P_1 = \gamma(\theta), P_2 = \gamma(\theta + \pi/2), etc.; verify distances |P_1 P_2| = |P_2 P_3| = |P_3 P_4| = |P_4 P_1| and angles 90°; rotate \theta for variants. This approach builds intuition for more complex curves by emphasizing midpoint loci and perpendicularity constraints.

Historical Background

Origins and Early Proofs

The inscribed square problem, also known as the square peg problem, was first posed by in 1911 during a talk titled "Über einige Aufgaben der Analysis situs" at the meeting of the Swiss Society of Natural Sciences in Solothurn. Toeplitz conjectured that every simple closed continuous curve in the plane—now known as a —admits four points that form the vertices of a square. This question arose as an extension of earlier results in geometric topology on inscribed polygons. The initial motivation for Toeplitz's conjecture was to generalize well-understood cases for specific curve types, such as circles (which admit infinitely many inscribed squares) and polygons (where the existence is intuitive but required proof), to arbitrary continuous Jordan curves. Toeplitz suggested the problem to his students as an exercise, highlighting its apparent simplicity yet challenging nature in bridging combinatorial geometry and topology. Although Toeplitz claimed a solution for convex curves in his 1911 talk, no detailed proof was published at the time. The first significant progress came in 1916 with a proof by Arnold Emch for a broad class of curves, including every polygonal Jordan curve, which is a piecewise linear simple closed curve with finitely many vertices. Emch demonstrated that such curves always admit an inscribed square, employing continuity arguments based on side lengths and angles of potential quadrilaterals. His method involves selecting pairs of points on the curve to form chords and considering rotations to identify perpendicular chords of equal length, which connect to form a rhombus; by continuously varying the orientation, the side lengths and angles adjust continuously, guaranteeing a configuration where they equalize to form a square due to the intermediate value theorem applied to the difference in side lengths. This approach relies on the analytic nature of the arcs, but since linear segments are analytic, it directly applies to polygons. Emch's work built on his earlier 1913 paper addressing convex curves and was prompted by a suggestion from mathematician Aubrey J. Kempner, who was aware of Toeplitz's conjecture. Early related developments included connections to classical results in differential geometry, such as the (established by Mukhopadhyaya in 1909), which asserts that every closed convex curve has at least four vertices (points of extremal curvature), providing a geometric foundation for analyzing inscribed polygons. Additionally, Emch's continuity-based oscillation arguments echo techniques in on the number of zeros of oscillatory functions, which underpins counting intersections in such proofs, though explicit links were not formalized until later works. These foundational efforts set the stage for extending the result beyond piecewise linear and analytic cases.

Key Milestones in the 20th Century

In the early decades of the 20th century, significant progress was made on the for convex curves. established foundational results, proving in 1913 that sufficiently smooth convex closed curves admit an inscribed square by analyzing secant lines and medians to identify perpendicular chords of equal length. He extended this in 1915 to convex curves under weaker smoothness assumptions, employing limit arguments that implicitly cover all convex cases. By 1916, Emch further generalized his approach to piecewise analytic convex curves with finitely many inflection points, ensuring the existence of tangents at nonsmooth vertices and leveraging continuity of medians. These works built on earlier polygonal proofs, but shifted focus to continuous curves using geometric continuity arguments. The 1920s brought advances for smoother nonconvex curves. In 1921, Konrad Zindler provided an independent proof for general convex curves, confirming Emch's results through variational methods that minimized certain functionals over inscribed quadrilaterals. Lev Schnirelmann achieved a breakthrough in 1929 (published posthumously in 1944) by proving the conjecture for curves slightly more regular than C², employing bordism arguments from algebraic topology to show that continuous deformations of the curve preserve the parity of inscribed squares. This topological approach marked a departure from purely geometric techniques and influenced later extensions to less regular classes. Mid-century efforts targeted analytic curves. In 1961, Richard P. Jerrard resolved the problem for analytic Jordan curves, utilizing complex analysis and fixed-point theorems on the Riemann sphere to guarantee the intersection of perpendicular chords. This result relied on analytic continuation properties, establishing the existence of inscribed squares via degree theory. Additional confirmations for convex curves appeared, such as C. M. Christensen's 1950 proof and Roger Fenn's 1970 application of the table theorem with mod-2 homology arguments. The late 20th century saw proofs for broader classes, including monotone and symmetric curves. In 1989, Walter Stromquist proved the conjecture for locally monotone continuous curves, constructing inscribed rhombi and showing that monotonicity ensures perpendicular sides without "special trapezoids"—isosceles trapezoids with equal non-parallel sides that are not squares. This avoided pathological configurations by pairing potential trapezoids topologically. In 1995, Mark J. Nielsen and Stephen E. Wright extended the result to curves symmetric with respect to a line or point, exploiting group actions to reduce the problem to fixed-point existence on invariant subsets. These developments highlighted the role of symmetry and monotonicity in circumventing obstructions for general Jordan curves.
YearAuthor(s)Key Result
1911Otto ToeplitzFormulation of the inscribed square conjecture for Jordan curves.
1913Arnold EmchProof for smooth convex curves via secants and medians.
1915Arnold EmchExtension to convex curves with weaker smoothness.
1916Arnold EmchProof for piecewise analytic curves with finite inflections.
1921Konrad ZindlerIndependent proof for convex curves using variational methods.
1929Lev SchnirelmannProof for curves slightly beyond C² using bordism (published 1944).
1950C. M. ChristensenConfirmation for convex curves.
1961Richard P. JerrardProof for analytic curves via complex analysis and fixed points.
1970Roger FennProof for convex curves using the table theorem and homology.
1989Walter StromquistProof for locally monotone curves, avoiding special trapezoids.
1995Mark J. Nielsen and Stephen E. WrightProof for line- or point-symmetric curves via group actions.

Resolved Cases

Analytic and Piecewise Analytic Curves

The existence of inscribed squares in analytic and piecewise analytic curves is established through proofs that exploit the curve's local smoothness and global continuity, enabling the application of intermediate value and degree-like arguments. In 1916, Arnold Emch proved that every piecewise analytic Jordan curve with finitely many inflection points admits at least one inscribed square. The proof partitions the curve into finitely many analytic arcs separated at inflection points or corners. On each analytic arc, pairs of points are selected, and the potential completing sides of a square are constructed geometrically, such as by drawing lines perpendicular to the chord at its midpoint or considering medians of the quadrilateral formed by the points. As the pair of points varies continuously along the arc, the distance between the intersection points of these constructions with the curve varies continuously; by the intermediate value theorem, this distance must vanish at some position, yielding the other two vertices of the square on the curve. For the global curve, these local constructions are combined across arcs, ensuring at least one configuration closes to form a full inscribed square without self-intersections. This method particularly applies to the subcase of polygons, where the arcs are straight line segments; here, an explicit algorithm emerges by applying the intermediate value theorem to the lengths of diagonals between points on opposite edges, guaranteeing a position where the diagonals are equal and perpendicular, forming a square. Emch's result was extended to fully analytic curves by R. P. Jerrard in 1961, who showed that a simple closed real-analytic plane curve—parameterized analytically over the circle—has either an odd finite number of inscribed squares or infinitely many forming a continuous family. Parameterizing the curve via real-analytic functions x(t) and y(t) with period T, Jerrard constructs a multi-valued analytic function f(t) that encodes the parametric shifts needed for the other two vertices of a square given an initial chord. The graph of f over [0, T] consists of continuous branches, with isolated tangent points where branches meet or become double-valued. By analyzing the connectivity and multiplicity of these branches—ensuring at least one component has odd degree in its winding or intersection properties—Jerrard applies continuity and parity arguments to conclude that the equation f(t) = t \pmod{T} has an odd number of solutions, corresponding to inscribed squares. This relies on power series expansions of the analytic parameterization to resolve local behavior near singularities, guaranteeing global solvability through a topological degree analogous to Brouwer's fixed-point theorem. The distinction between C^1-smooth and analytic curves lies in the latter's allowance for precise local approximations via convergent power series, which facilitate holomorphic-like extensions and stronger multiplicity counts in configuration spaces. In contrast, piecewise analytic curves permit discontinuities in higher derivatives at finitely many points, so proofs like Emch's rely on patching local analytic solutions with global topological continuity rather than uniform holomorphy. Recent developments for the piecewise linear subcase, such as polygons, include purely topological proofs using equivariant methods on configuration spaces, avoiding analytic tools altogether; for instance, by considering the action of the on pairs of edges and applying to show fixed points corresponding to squares. These approaches highlight how analyticity enables degree-theoretic guarantees, while piecewise cases bridge to more general smooth curves through approximation.

Monotone and Lipschitz Curves

A locally monotone curve is defined as a Jordan curve where, at every point, there exists a neighborhood such that the tangent angle is non-decreasing along the curve in that neighborhood. This condition captures curves with controlled directional variation, including convex curves and piecewise linear curves. In 1989, Walter Stromquist established that every locally monotone Jordan curve admits an inscribed square. His proof proceeds topologically by considering the configuration space of inscribed rectangles on the curve and showing a continuous deformation path that transforms any such rectangle into a square, leveraging the monotonicity to ensure the existence of a fixed point under this variation. Lipschitz graphs provide another class of curves with bounded slope variation, where a function f: [a, b] \to \mathbb{R} satisfies |f(x) - f(y)| \leq K |x - y| for some constant K \geq 0, and the graph is the set \{(x, f(x)) \mid x \in [a, b]\}. For a simple closed curve formed by the union of two such graphs sharing endpoints, early progress came from in 1929, who proved the existence of an inscribed square for curves with bounded continuous curvature, a condition implying a Lipschitz constant K < 1 in suitable parameterizations. In 2017, Terence Tao extended this to Lipschitz constants K < 1, using an integration argument over potential square positions. The 2024 improvement by Joshua Greene and Andrew Lobb further raised the threshold to K < 1 + \sqrt{2}, establishing that such curves always contain an inscribed square. Their approach employs symplectic geometry on the configuration space of quadrilaterals, focusing on flux integrals of the form \int_{\gamma} x \, dy - y \, dx along candidate square boundaries \gamma. By demonstrating that a non-zero flux corresponds to a non-trivial winding number in the symplectic form, they show that the zero set of this flux must intersect the space of potential squares, guaranteeing existence. Another resolution arises for Jordan curves avoiding special trapezoids, defined as inscribed isosceles trapezoids with equal-length parallel sides that are not squares. Stromquist's 1989 work also covers this case: any such curve without special trapezoids of slope 1 admits an inscribed square. The proof relies on a contradiction argument in the space of chord families, where the absence of these trapezoids implies an odd parity in the number of inscribed rhombi that must resolve into a square via topological invariance under deformation.

Symmetric and Annular Configurations

In symmetric configurations, the inscribed square problem benefits from geometric invariances that enforce the existence of inscribed quadrilaterals, often via fixed-point principles or pairing arguments. For Jordan curves exhibiting bilateral symmetry across a line (reflection symmetry), the reflection maps points on one side to the other, allowing the construction of inscribed rectangles by pairing corresponding points and ensuring perpendicularity through the symmetry axis. This result extends to continua symmetric across a hyperplane, where the symmetry guarantees at least one inscribed rectangle. Similarly, for curves invariant under central symmetry (180-degree rotational invariance about a point), the antipodal mapping pairs opposite points on the curve, yielding inscribed rectangles that include squares as special cases when the aspect ratio is 1:1. These symmetries align potential square vertices along axes of invariance, leveraging the curve's self-mapping properties to avoid degenerate configurations. Piecewise symmetric cases, where the curve possesses bilateral symmetry over distinct segments, further apply reflection pairings to localize inscribed rectangles within symmetric portions, then extend globally via continuity. For instance, if a Jordan curve is composed of arcs each symmetric across parallel lines, reflections pair endpoints to form candidate sides, and intersection arguments ensure a closed quadrilateral. This approach handles non-smooth transitions by restricting to compact subsets where symmetry holds locally. In annular configurations, Jordan curves embedded in an annulus—specifically, those separating two concentric circles—admit inscribed squares due to the region's topological constraints. A seminal extension shows that any such curve in an annulus with outer radius at most $1 + \sqrt{2} times the inner radius contains an inscribed square, often with sides oriented radially and tangentially to exploit the annular geometry. The proof relies on the annulus's fundamental group \pi_1(A) \cong \mathbb{Z}, which ensures the curve winds non-trivially around the origin, preventing evasion of square inscriptions. If the annulus is sufficiently thin, the space of potential inscribed squares (parameterized by position, size, and orientation) decomposes into two connected components separated by the curve; topological invariance then forces an odd number of intersections, guaranteeing at least one square. Variants of the normalize the annular embedding, ensuring the separating curve behaves like a standard circle under homeomorphisms that preserve radial and angular coordinates, thereby facilitating chord constructions that form squares via crossing perpendicular bisectors. A representative example is a Jordan curve in the annulus $1 \leq \|x\| \leq 1 + \sqrt{2} that spirals monotonically from the inner to outer boundary. An explicit inscribed square can be constructed by identifying minimal and maximal radial extents at orthogonal angles (e.g., 0° and 90°), then pairing tangential points at equal arc lengths to form sides: let r_{\min}(\theta) and r_{\max}(\theta) denote the curve's radial profile; the square vertices occur at angles \theta, \theta + 90^\circ, \theta + 180^\circ, \theta + 270^\circ with radii interpolated as (r_{\min}(\theta) + r_{\max}(\theta))/2, adjusted for perpendicularity, yielding side length at least \sqrt{2} by the thinness condition. This construction aligns two opposite sides tangentially (constant radius) and the others radially (constant angle), confirmed to lie on the curve via the separation property.

Approximations to Smooth Curves

Recent advances in the inscribed square problem have addressed Jordan curves that approximate smooth ones, providing a pathway toward the general case through perturbation and stability arguments. In particular, results establish the existence of inscribed squares for curves sufficiently close to those with higher regularity, leveraging the known solvability for C^2 Jordan curves. A key contribution is the work showing that if a Jordan curve \gamma in \mathbb{R}^2 is sufficiently close to a C^2 Jordan curve \beta—specifically, if there exists a map f such that \|f(x) - x\| < 1/(10\kappa) for all x, where \kappa is the maximum unsigned curvature of \beta, and f \circ \gamma has winding number 1—then \gamma contains an inscribed square. This theorem, proved by in 2022, relies on C^0 convergence of \gamma to \beta and the stability of square configurations under small perturbations. The proof constructs a continuous homotopy between \beta and approximations \alpha_i of \gamma, using curvature bounds to ensure that inscribed squares in \beta (with segment lengths at least \pi/\kappa) persist without collapsing into degenerate configurations during the deformation. Intermediate-sized squares are ruled out by Proposition 3.3, which exploits the geometry of close C^2 curves to maintain non-degeneracy. This approach builds on earlier resolutions for C^2 curves, which were partially addressed using integration methods to track potential square vertices via conserved quantities like signed areas \int_\gamma y \, dx. Tao's 2017 analysis provided a framework for such smooth cases by applying the contraction mapping theorem and homological invariants, though it remained partial for general C^2 curves without additional Lipschitz constraints. The perturbation result thus inherits existence from these smooth bases through continuous dependence, bridging to less regular curves. An extension by Greene and Lobb in 2021 further supports this direction by proving that every smooth Jordan curve admits inscribed cyclic quadrilaterals of arbitrary aspect ratios, with squares emerging as the special case of equal sides and right angles. Their proof, while focused on rectangles, implies the abundance of square-like configurations in smooth settings, reinforcing the stability arguments for nearby curves. However, these approximation results are limited: the perturbation parameter \varepsilon (or equivalent distance) must be sufficiently small relative to the curvature of the approximating C^2 curve, and the method does not extend to arbitrary continuous without such proximity to smoothness.

Variants and Generalizations

Rectangular and Cyclic Quadrilaterals

The rectangular peg problem extends the inscribed square problem by asking whether every Jordan curve admits an inscribed rectangle similar to any given one with prescribed aspect ratio. In 2020, Joshua Greene and Andrew Lobb resolved this affirmatively for smooth Jordan curves, proving that for every smooth Jordan curve \gamma and any rectangle R in the Euclidean plane, there exists a rectangle similar to R whose vertices lie on \gamma. Their approach embeds the curve \gamma into a contact manifold via Seidel's representation and leverages embedded contact homology (ECH) capacities to detect the existence of symplectic fillings corresponding to the desired rectangle. This framework naturally generalizes to cyclic quadrilaterals. In 2021, Greene and Lobb extended their methods to show that for every smooth Jordan curve \gamma and any cyclic quadrilateral Q in the Euclidean plane, there exists a cyclic quadrilateral inscribed in \gamma that is similar to Q, using symplectic geometry and embedded contact homology capacities to detect existence via properties of Lagrangian tori in \mathbb{C}^2. Squares represent special instances of both cases: as rectangles with aspect ratio 1 or as cyclic quadrilaterals with all angles equal to $90^\circ. These results provide partial progress toward the inscribed square problem by confirming the existence of such configurations for smooth curves, though the full square peg problem remains unresolved for general non-smooth Jordan curves. For non-smooth Jordan curves, the rectangular peg problem with prescribed aspect ratio lacks a general resolution.

Higher-Dimensional and Polygonal Extensions

The inscribed square problem extends to the existence of inscribed regular n-gons in Jordan curves for n > 4. For odd n ≥ 3, every Jordan curve admits an inscribed regular n-gon. This was resolved in the using arguments from topological theory, building on Meyerson's proof for equilateral (n=), which showed that any continuous simple closed curve in the contains the vertices of infinitely many equilateral triangles via a degree computation on the configuration space of triangle vertices. The generalization to higher odd n relies on similar equivariant , ensuring the existence map has nonzero degree due to the odd preventing fixed-point-free maps. For even n > 4, the problem remains open in general, though counterexamples exist for certain curves, such as a connected by its , which admits no inscribed regular n-gon for even n ≥ . Variants of the problem involving non-regular quadrilaterals, such as and , have been resolved for curves using affine invariance. Any is the affine image of , and affine transformations preserve both the smoothness of the curve and the inscription ( inscribed squares to inscribed parallelograms). Since curves admit inscribed squares by classical results (e.g., Emch's 1916 theorem), they therefore admit inscribed of any prescribed shape up to affine equivalence. , as special , follow similarly, with existence guaranteed for curves via the same mechanism; recent work by Greene and Lobb (2021) extends this to confirm that every curve inscribes every possible (a special or ) up to orientation-preserving similarity. In higher dimensions, the problem generalizes to inscribing regular polytopes or their analogues in knotted curves or . A key open question is whether every knotted curve in ℝ³ admits an inscribed regular tetrahedron. Partial results exist for the boundaries of convex bodies: for instance, every centrally symmetric convex body in ℝ³ admits an inscribed , proven using equivariant on the of the cube. Similar techniques yield existence for spheres (the of the unit ), but the knotted case remains unresolved, with no known counterexamples. The rectangular peg problem also extends to higher dimensions, asking whether every smooth in ℝᵈ admits an inscribed of prescribed s; this is open beyond d=2, though existence holds without aspect ratio constraints for smooth cases via degree-theoretic arguments analogous to the planar version. Open problems in these extensions include whether every smooth knotted curve in ℝ³ admits an inscribed . Partial affirmative results exist for specific classes of space curves, such as those with prescribed , where (2004) showed existence of inscribed regular polygons approximating the curve via variational methods. Connections arise to Steinitz's theorem, which characterizes graphs realizable as boundaries of convex polyhedra in ℝ³; inscribed regular polytopes on knotted curves or spheres relate to realizing such graphs with straight-line embeddings on the curve, providing a topological obstruction or existence criterion for polyhedral approximations in higher dimensions.

References

  1. [1]
    [PDF] arXiv:2103.13848v1 [math.DG] 25 Mar 2021
    Mar 25, 2021 · A BRIEF HISTORY OF THE SQUARE-PEG PROBLEM. Let us recall the square-peg problem, originally due to O. Toeplitz [38]. Conjecture 13. Let γ: S1 ...
  2. [2]
    [PDF] A survey on the Square Peg Problem
    Mar 14, 2014 · This is a short survey article on the 102 years old Square Peg Problem of Toeplitz, which is also called the Inscribed Square Problem. It ...
  3. [3]
    [PDF] THE SQUARE PEG PROBLEM
    The Square Peg Problem, also known as the Inscribed Square Problem or Toeplitz' Con- jecture, is a still open problem which numerous authors have tried to ...
  4. [4]
    An integration approach to the Toeplitz square peg problem
    Nov 22, 2016 · Inscribed Square Conjecture is true! The mapping all four points of the Jordan curve will eventually converge to the four corners or vertices, ...
  5. [5]
    Two Mathematicians Just Solved a Century-Old Geometry Problem
    Jun 29, 2020 · In 1911, German mathematician Otto Toeplitz first posed the inscribed square problem, in which he predicted that "any closed curve contains ...
  6. [6]
    Jordan Curve -- from Wolfram MathWorld
    A Jordan curve is a plane curve which is topologically equivalent to (a homeomorphic image of) the unit circle, i.e., it is simple and closed.
  7. [7]
    A Note on Toeplitz' Conjecture | Discrete & Computational Geometry
    Feb 19, 2014 · Inscribed square · Jordan curve · Toeplitz' conjecture · Use our pre-submission checklist. Avoid common mistakes on your manuscript.
  8. [8]
    An integration approach to the Toeplitz square peg problem - arXiv
    Nov 22, 2016 · The "square peg problem" or "inscribed square problem" of Toeplitz asks if every simple closed curve in the plane inscribes a (non-degenerate) ...
  9. [9]
    Square -- from Wolfram MathWorld
    ### Summary: Inscribed Square in Circle and Parametric Equations for Vertices
  10. [10]
    Geometric constructions: circle-inscribed square - Khan Academy
    Aug 9, 2016 · Sal constructs a square that is inscribed inside a given circle using compass and straightedge. Created by Sal Khan.
  11. [11]
    Squares on curves | Dongryul Kim - Stanford University
    Jan 18, 2017 · Conjecture 1 (Toeplitz, 1911). Every Jordan curve in has an inscribed square, i.e., a square that has vertices on the curve.
  12. [12]
    Triangle Square Inscribing -- from Wolfram MathWorld
    Calabi's triangle is the only triangle (besides the equilateral triangle) for which the largest inscribed square can be inscribed in three different ways.
  13. [13]
    [PDF] arXiv:1712.10205v2 [math.MG] 4 Jun 2018
    Jun 4, 2018 · Makeev conjectured that any cyclic quadrilateral can be inscribed in any Jordan curve. He proved the conjecture for the case of star-shaped C2- ...
  14. [14]
  15. [15]
    A Survey on the Square Peg Problem
    Benjamin Matschke, Equivariant topology methods in discrete geometry, Ph.D. thesis, Freie Universität Berlin,. 2011. 27. Mark D. Meyerson, Equilateral ...
  16. [16]
    On the Medians of a Closed Convex Polygon - jstor
    In what follows I shall show that at least one square may be inscribed in any convex rectilinear polygon, and generally in any closed convex curve formed by ...<|control11|><|separator|>
  17. [17]
    None
    ### Summary of Main Theorem and Proof Outline for Inscribed Squares in Analytic Plane Curves
  18. [18]
    [PDF] arXiv:0804.0657v1 [math.MG] 4 Apr 2008
    Apr 4, 2008 · Figure 1. Jordan curve C and an inscribed square. The problem goes back to Toeplitz (1911), and over almost a century has been.
  19. [19]
    [2407.07798] Square pegs between two graphs - arXiv
    Jul 10, 2024 · Authors:Joshua Evan Greene, Andrew Lobb. View a PDF of the paper titled Square pegs between two graphs, by Joshua Evan Greene and Andrew Lobb.
  20. [20]
    Rectangles inscribed in symmetric continua | Geometriae Dedicata
    LetQ be any rectangle and letK ⊂ R d (d ≥ 2) be a continuum which is either symmetric across a hyperplane or symmetric through a pointz ∉K.
  21. [21]
    [2203.02613] On the square peg problem - arXiv
    Mar 4, 2022 · We show that if \gamma is a Jordan curve in \mathbb{R}^2 which is close to a C^2 Jordan curve \beta in \mathbb{R}^2, then \gamma contains an inscribed square.Missing: GR 2025
  22. [22]
    [2005.09193] The Rectangular Peg Problem - arXiv
    May 19, 2020 · Authors:Joshua Evan Greene, Andrew Lobb. View a PDF of the paper titled The Rectangular Peg Problem, by Joshua Evan Greene and Andrew Lobb.
  23. [23]
    [2011.05216] Cyclic quadrilaterals and smooth Jordan curves - arXiv
    Nov 10, 2020 · View a PDF of the paper titled Cyclic quadrilaterals and smooth Jordan curves, by Joshua Evan Greene and Andrew Lobb. View PDF. Abstract:For ...Missing: inscribed 2021