Möbius strip

Last updated

A Mobius strip made with paper and adhesive tape Mobius Strip.jpg
A Möbius strip made with paper and adhesive tape

In mathematics, a Möbius strip, Möbius band, or Möbius loop [lower-alpha 1] is a surface that can be formed by attaching the ends of a strip of paper together with a half-twist. As a mathematical object, it was discovered by Johann Benedict Listing and August Ferdinand Möbius in 1858, but it had already appeared in Roman mosaics from the third century CE. The Möbius strip is a non-orientable surface, meaning that within it one cannot consistently distinguish clockwise from counterclockwise turns. Every non-orientable surface contains a Möbius strip.

Contents

As an abstract topological space, the Möbius strip can be embedded into three-dimensional Euclidean space in many different ways: a clockwise half-twist is different from a counterclockwise half-twist, and it can also be embedded with odd numbers of twists greater than one, or with a knotted centerline. Any two embeddings with the same knot for the centerline and the same number and direction of twists are topologically equivalent. All of these embeddings have only one side, but when embedded in other spaces, the Möbius strip may have two sides. It has only a single boundary curve.

Several geometric constructions of the Möbius strip provide it with additional structure. It can be swept as a ruled surface by a line segment rotating in a rotating plane, with or without self-crossings. A thin paper strip with its ends joined to form a Möbius strip can bend smoothly as a developable surface or be folded flat; the flattened Möbius strips include the trihexaflexagon. The Sudanese Möbius strip is a minimal surface in a hypersphere, and the Meeks Möbius strip is a self-intersecting minimal surface in ordinary Euclidean space. Both the Sudanese Möbius strip and another self-intersecting Möbius strip, the cross-cap, have a circular boundary. A Möbius strip without its boundary, called an open Möbius strip, can form surfaces of constant curvature. Certain highly-symmetric spaces whose points represent lines in the plane have the shape of a Möbius strip.

The many applications of Möbius strips include mechanical belts that wear evenly on both sides, dual-track roller coasters whose carriages alternate between the two tracks, and world maps printed so that antipodes appear opposite each other. Möbius strips appear in molecules and devices with novel electrical and electromechanical properties, and have been used to prove impossibility results in social choice theory. In popular culture, Möbius strips appear in artworks by M. C. Escher, Max Bill, and others, and in the design of the recycling symbol. Many architectural concepts have been inspired by the Möbius strip, including the building design for the NASCAR Hall of Fame. Performers including Harry Blackstone Sr. and Thomas Nelson Downs have based stage magic tricks on the properties of the Möbius strip. The canons of J. S. Bach have been analyzed using Möbius strips. Many works of speculative fiction feature Möbius strips; more generally, a plot structure based on the Möbius strip, of events that repeat with a twist, is common in fiction.

History

Aion mosaic Glyptothek Munich W504.jpg
Mosaic from ancient Sentinum depicting Aion holding a Möbius strip
Al-Jazari Automata 1205.jpg
Chain pump with a Möbius drive chain, by Ismail al-Jazari (1206)

The discovery of the Möbius strip as a mathematical object is attributed independently to the German mathematicians Johann Benedict Listing and August Ferdinand Möbius in 1858. [2] However, it had been known long before, both as a physical object and in artistic depictions; in particular, it can be seen in several Roman mosaics from the third century CE. [3] [4] In many cases these merely depict coiled ribbons as boundaries. When the number of coils is odd, these ribbons are Möbius strips, but for an even number of coils they are topologically equivalent to untwisted rings. Therefore, whether the ribbon is a Möbius strip may be coincidental, rather than a deliberate choice. In at least one case, a ribbon with different colors on different sides was drawn with an odd number of coils, forcing its artist to make a clumsy fix at the point where the colors did not match up. [3] Another mosaic from the town of Sentinum (depicted) shows the zodiac, held by the god Aion, as a band with only a single twist. There is no clear evidence that the one-sidedness of this visual representation of celestial time was intentional; it could have been chosen merely as a way to make all of the signs of the zodiac appear on the visible side of the strip. Some other ancient depictions of the ourobouros or of figure-eight-shaped decorations are also alleged to depict Möbius strips, but whether they were intended to depict flat strips of any type is unclear. [4]

Independently of the mathematical tradition, machinists have long known that mechanical belts wear half as quickly when they form Möbius strips, because they use the entire surface of the belt rather than only the inner surface of an untwisted belt. [3] Additionally, such a belt may be less prone to curling from side to side. An early written description of this technique dates to 1871, which is after the first mathematical publications regarding the Möbius strip. Much earlier, an image of a chain pump in a work of Ismail al-Jazari from 1206 depicts a Möbius strip configuration for its drive chain. [4] Another use of this surface was made by seamstresses in Paris (at an unspecified date): they initiated novices by requiring them to stitch a Möbius strip as a collar onto a garment. [3]

Properties

A 2D object traversing once around the Mobius strip returns in mirrored form Fiddler crab mobius strip.gif
A 2D object traversing once around the Möbius strip returns in mirrored form

The Möbius strip has several curious properties. It is a non-orientable surface: if an asymmetric two-dimensional object slides one time around the strip, it returns to its starting position as its mirror image. In particular, a curved arrow pointing clockwise (↻) would return as an arrow pointing counterclockwise (↺), implying that, within the Möbius strip, it is impossible to consistently define what it means to be clockwise or counterclockwise. It is the simplest non-orientable surface: any other surface is non-orientable if and only if it has a Möbius strip as a subset. [5] Relatedly, when embedded into Euclidean space, the Möbius strip has only one side. A three-dimensional object that slides one time around the surface of the strip is not mirrored, but instead returns to the same point of the strip on what appears locally to be its other side, showing that both positions are really part of a single side. This behavior is different from familiar orientable surfaces in three dimensions such as those modeled by flat sheets of paper, cylindrical drinking straws, or hollow balls, for which one side of the surface is not connected to the other. [6] However, this is a property of its embedding into space rather than an intrinsic property of the Möbius strip itself: there exist other topological spaces in which the Möbius strip can be embedded so that it has two sides. [7] For instance, if the front and back faces of a cube are glued to each other with a left-right mirror reflection, the result is a three-dimensional topological space (the Cartesian product of a Möbius strip with an interval) in which the top and bottom halves of the cube can be separated from each other by a two-sided Möbius strip. [lower-alpha 2] In contrast to disks, spheres, and cylinders, for which it is possible to simultaneously embed an uncountable set of disjoint copies into three-dimensional space, only a countable number of Möbius strips can be simultaneously embedded. [9] [10] [11]

A path along the edge of a Möbius strip, traced until it returns to its starting point on the edge, includes all boundary points of the Möbius strip in a single continuous curve. For a Möbius strip formed by gluing and twisting a rectangle, it has twice the length of the centerline of the strip. In this sense, the Möbius strip is different from an untwisted ring and like a circular disk in having only one boundary. [6] A Möbius strip in Euclidean space cannot be moved or stretched into its mirror image; it is a chiral object with right- or left-handedness. [12] Möbius strips with odd numbers of half-twists greater than one, or that are knotted before gluing, are distinct as embedded subsets of three-dimensional space, even though they are all equivalent as two-dimensional topological surfaces. [13] More precisely, two Möbius strips are equivalently embedded in three-dimensional space when their centerlines determine the same knot and they have the same number of twists as each other. [14] With an even number of twists, however, one obtains a different topological surface, called the annulus. [15]

The Möbius strip can be continuously transformed into its centerline, by making it narrower while fixing the points on the centerline. This transformation is an example of a deformation retraction, and its existence means that the Möbius strip has many of the same properties as its centerline, which is topologically a circle. In particular, its fundamental group is the same as the fundamental group of a circle, an infinite cyclic group. Therefore, paths on the Möbius strip that start and end at the same point can be distinguished topologically (up to homotopy) only by the number of times they loop around the strip. [16]

Moebiusband-1s.svg
Cutting the centerline produces a double-length two-sided (non-Möbius) strip
Moebiusband-2s.svg
A single off-center cut produces a Möbius strip (purple) linked with a double-length two-sided strip

Cutting a Möbius strip along the centerline with a pair of scissors yields one long strip with four half-twists in it (relative to an untwisted annulus or cylinder) rather than two separate strips. Two of the half-twists come from the fact that this thinner strip goes two times through the half-twist in the original Möbius strip, and the other two come from the way the two halves of the thinner strip wrap around each other. The result is not a Möbius strip, but instead is topologically equivalent to a cylinder. Cutting this double-twisted strip again along its centerline produces two linked double-twisted strips. If, instead, a Möbius strip is cut lengthwise, a third of the way across its width, it produces two linked strips. One of the two is a central, thinner, Möbius strip, while the other has two half-twists. [6] These interlinked shapes, formed by lengthwise slices of Möbius strips with varying widths, are sometimes called paradromicrings. [17] [18]

Tietze-Moebius.svg
Subdivision into six mutually-adjacent regions, bounded by Tietze's graph
3 utilities problem moebius.svg
Solution to the three utilities problem on a Möbius strip

The Möbius strip can be cut into six mutually-adjacent regions, showing that maps on the surface of the Möbius strip can sometimes require six colors, in contrast to the four color theorem for the plane. [19] Six colors are always enough. This result is part of the Ringel–Youngs theorem, which states how many colors each topological surface needs. [20] The edges and vertices of these six regions form Tietze's graph, which is a dual graph on this surface for the six-vertex complete graph but cannot be drawn without crossings on a plane. Another family of graphs that can be embedded on the Möbius strip, but not on the plane, are the Möbius ladders, the boundaries of subdivisions of the Möbius strip into rectangles meeting end-to-end. [21] These include the utility graph, a six-vertex complete bipartite graph whose embedding into the Möbius strip shows that, unlike in the plane, the three utilities problem can be solved on a transparent Möbius strip. [22] The Euler characteristic of the Möbius strip is zero, meaning that for any subdivision of the strip by vertices and edges into regions, the numbers , , and of vertices, edges, and regions satisfy . For instance, Tietze's graph has vertices, edges, and regions; . [19]

Constructions

There are many different ways of defining geometric surfaces with the topology of the Möbius strip, yielding realizations with additional geometric properties.

Sweeping a line segment

Mobius strip.gif
A Möbius strip swept out by a rotating line segment in a rotating plane
Plucker's conoid (n=2).gif
Plücker's conoid swept out by a different motion of a line segment

One way to embed the Möbius strip in three-dimensional Euclidean space is to sweep it out by a line segment rotating in a plane, which in turn rotates around one of its lines. [23] For the swept surface to meet up with itself after a half-twist, the line segment should rotate around its center at half the angular velocity of the plane's rotation. This can be described as a parametric surface defined by equations for the Cartesian coordinates of its points,

for and , where one parameter describes the rotation angle of the plane around its central axis and the other parameter describes the position of a point along the rotating line segment. This produces a Möbius strip of width 1, whose center circle has radius 1, lies in the -plane and is centered at . [24] The same method can produce Möbius strips with any odd number of half-twists, by rotating the segment more quickly in its plane. The rotating segment sweeps out a circular disk in the plane that it rotates within, and the Möbius strip that it generates forms a slice through the solid torus swept out by this disk. Because of the one-sidedness of this slice, the sliced torus remains connected. [25]

A line or line segment swept in a different motion, rotating in a horizontal plane around the origin as it moves up and down, forms Plücker's conoid or cylindroid, an algebraic ruled surface in the form of a self-crossing Möbius strip. [26] It has applications in the design of gears. [27]

Polyhedral surfaces and flat foldings

Trihexaflexagon being flexed Flexagon.gif
Trihexaflexagon being flexed

A strip of paper can form a flattened Möbius strip in the plane by folding it at angles so that its center line lies along an equilateral triangle, and attaching the ends. The shortest strip for which this is possible consists of three equilateral triangles, folded at the edges where two triangles meet. Its aspect ratio  the ratio of the strip's length [lower-alpha 3] to its width is , and the same folding method works for any larger aspect ratio. [28] [29] For a strip of nine equilateral triangles, the result is a trihexaflexagon, which can be flexed to reveal different parts of its surface. [30] For strips too short to apply this method directly, one can first "accordion fold" the strip in its wide direction back and forth using an even number of folds. With two folds, for example, a strip would become a folded strip whose cross section is in the shape of an 'N' and would remain an 'N' after a half-twist. The narrower accordion-folded strip can then be folded and joined in the same way that a longer strip would be. [28] [29]

5-vertex polyhedral Mobius strip.svg
Pentagonal Mobius strip.svg
Five-vertex polyhedral and flat-folded Möbius strips

The Möbius strip can also be embedded as a polyhedral surface in space or flat-folded in the plane, with only five triangular faces sharing five vertices. In this sense, it is simpler than the cylinder, which requires six triangles and six vertices, even when represented more abstractly as a simplicial complex. [31] [lower-alpha 4] A five-triangle Möbius strip can be represented most symmetrically by five of the ten equilateral triangles of a four-dimensional regular simplex. This four-dimensional polyhedral Möbius strip is the only tight Möbius strip, one that is fully four-dimensional and for which all cuts by hyperplanes separate it into two parts that are topologically equivalent to disks or circles. [32]

Other polyhedral embeddings of Möbius strips include one with four convex quadrilaterals as faces, another with three non-convex quadrilateral faces, [33] and one using the vertices and center point of a regular octahedron, with a triangular boundary. [34] Every abstract triangulation of the projective plane can be embedded into 3D as a polyhedral Möbius strip with a triangular boundary after removing one of its faces; [35] an example is the six-vertex projective plane obtained by adding one vertex to the five-vertex Möbius strip, connected by triangles to each of its boundary edges. [31] However, not every abstract triangulation of the Möbius strip can be represented geometrically, as a polyhedral surface. [36] To be realizable, it is necessary and sufficient that there be no two disjoint non-contractible 3-cycles in the triangulation. [37]

Smoothly embedded rectangles

A rectangular Möbius strip, made by attaching the ends of a paper rectangle, can be embedded smoothly into three-dimensional space whenever its aspect ratio is greater than , the same ratio as for the flat-folded equilateral-triangle version of the Möbius strip. [38] This flat triangular embedding can lift to a smooth [lower-alpha 5] embedding in three dimensions, in which the strip lies flat in three parallel planes between three cylindrical rollers, each tangent to two of the planes. [38] Mathematically, a smoothly embedded sheet of paper can be modeled as a developable surface, that can bend but cannot stretch. [39] [40] As its aspect ratio decreases toward , all smooth embeddings seem to approach the same triangular form. [41]

The lengthwise folds of an accordion-folded flat Möbius strip prevent it from forming a three-dimensional embedding in which the layers are separated from each other and bend smoothly without crumpling or stretching away from the folds. [29] Instead, unlike in the flat-folded case, there is a lower limit to the aspect ratio of smooth rectangular Möbius strips. Their aspect ratio cannot be less than , even if self-intersections are allowed. Self-intersecting smooth Möbius strips exist for any aspect ratio above this bound. [29] [42] Without self-intersections, the aspect ratio must be at least [43]

Unsolved problem in mathematics:

Can a paper rectangle be glued end-to-end to form a smooth Möbius strip embedded in space? [lower-alpha 6]

For aspect ratios between this bound and , it has been an open problem whether smooth embeddings, without self-intersection, exist. [29] [42] [43] In 2023, Richard Schwartz announced a proof that they do not exist, but this result still awaits peer review and publication. [44] [45] If the requirement of smoothness is relaxed to allow continuously differentiable surfaces, the Nash–Kuiper theorem implies that any two opposite edges of any rectangle can be glued to form an embedded Möbius strip, no matter how small the aspect ratio becomes. [lower-alpha 7] The limiting case, a surface obtained from an infinite strip of the plane between two parallel lines, glued with the opposite orientation to each other, is called the unbounded Möbius strip or the real tautological line bundle. [46] Although it has no smooth closed embedding into three-dimensional space, it can be embedded smoothly as a closed subset of four-dimensional Euclidean space. [47]

The minimum-energy shape of a smooth Möbius strip glued from a rectangle does not have a known analytic description, but can be calculated numerically, and has been the subject of much study in plate theory since the initial work on this subject in 1930 by Michael Sadowsky. [39] [40] It is also possible to find algebraic surfaces that contain rectangular developable Möbius strips. [48] [49]

Making the boundary circular

Mobius to Klein.gif
Gluing two Möbius strips to form a Klein bottle
MobiusStrip-02.png
A projection of the Sudanese Möbius strip

The edge, or boundary, of a Möbius strip is topologically equivalent to a circle. In common forms of the Möbius strip, it has a different shape from a circle, but it is unknotted, and therefore the whole strip can be stretched without crossing itself to make the edge perfectly circular. [50] One such example is based on the topology of the Klein bottle, a one-sided surface with no boundary that cannot be embedded into three-dimensional space, but can be immersed (allowing the surface to cross itself in certain restricted ways). A Klein bottle is the surface that results when two Möbius strips are glued together edge-to-edge, and reversing that process a Klein bottle can be sliced along a carefully chosen cut to produce two Möbius strips. [51] For a form of the Klein bottle known as Lawson's Klein bottle, the curve along which it is sliced can be made circular, resulting in Möbius strips with circular edges. [52]

Lawson's Klein bottle is a self-crossing minimal surface in the unit hypersphere of 4-dimensional space, the set of points of the form

for . [53] Half of this Klein bottle, the subset with , gives a Möbius strip embedded in the hypersphere as a minimal surface with a great circle as its boundary. [54] This embedding is sometimes called the "Sudanese Möbius strip" after topologists Sue Goodman and Daniel Asimov, who discovered it in the 1970s. [55] Geometrically Lawson's Klein bottle can be constructed by sweeping a great circle through a great-circular motion in the 3-sphere, and the Sudanese Möbius strip is obtained by sweeping a semicircle instead of a circle, or equivalently by slicing the Klein bottle along a circle that is perpendicular to all of the swept circles. [52] [56] Stereographic projection transforms this shape from a three-dimensional spherical space into three-dimensional Euclidean space, preserving the circularity of its boundary. [52] The most symmetric projection is obtained by using a projection point that lies on that great circle that runs through the midpoint of each of the semicircles, but produces an unbounded embedding with the projection point removed from its centerline. [54] Instead, leaving the Sudanese Möbius strip unprojected, in the 3-sphere, leaves it with an infinite group of symmetries isomorphic to the orthogonal group , the group of symmetries of a circle. [53]

Schematic depiction of a cross-cap with an open bottom, showing its level sets. This surface crosses itself along the vertical line segment. Cross-cap level sets.svg
Schematic depiction of a cross-cap with an open bottom, showing its level sets. This surface crosses itself along the vertical line segment.

The Sudanese Möbius strip extends on all sides of its boundary circle, unavoidably if the surface is to avoid crossing itself. Another form of the Möbius strip, called the cross-cap or crosscap, also has a circular boundary, but otherwise stays on only one side of the plane of this circle, [57] making it more convenient for attaching onto circular holes in other surfaces. In order to do so, it crosses itself. It can be formed by removing a quadrilateral from the top of a hemisphere, orienting the edges of the quadrilateral in alternating directions, and then gluing opposite pairs of these edges consistently with this orientation. [58] The two parts of the surface formed by the two glued pairs of edges cross each other with a pinch point like that of a Whitney umbrella at each end of the crossing segment, [59] the same topological structure seen in Plücker's conoid. [26]

Surfaces of constant curvature

The open Möbius strip is the relative interior of a standard Möbius strip, formed by omitting the points on its boundary edge. It may be given a Riemannian geometry of constant positive, negative, or zero Gaussian curvature. The cases of negative and zero curvature form geodesically complete surfaces, which means that all geodesics ("straight lines" on the surface) may be extended indefinitely in either direction.

Zero curvature
An open strip with zero curvature may be constructed by gluing the opposite sides of a plane strip between two parallel lines, described above as the tautological line bundle. [46] The resulting metric makes the open Möbius strip into a (geodesically) complete flat surface (i.e., having zero Gaussian curvature everywhere). This is the unique metric on the Möbius strip, up to uniform scaling, that is both flat and complete. It is the quotient space of a plane by a glide reflection, and (together with the plane, cylinder, torus, and Klein bottle) is one of only five two-dimensional complete flat manifolds. [60]
Negative curvature
The open Möbius strip also admits complete metrics of constant negative curvature. One way to see this is to begin with the upper half plane (Poincaré) model of the hyperbolic plane, a geometry of constant curvature whose lines are represented in the model by semicircles that meet the -axis at right angles. Take the subset of the upper half-plane between any two nested semicircles, and identify the outer semicircle with the left-right reversal of the inner semicircle. The result is topologically a complete and non-compact Möbius strip with constant negative curvature. It is a "nonstandard" complete hyperbolic surface in the sense that it contains a complete hyperbolic half-plane (actually two, on opposite sides of the axis of glide-reflection), and is one of only 13 nonstandard surfaces. [61] Again, this can be understood as the quotient of the hyperbolic plane by a glide reflection. [62]
Positive curvature
A Möbius strip of constant positive curvature cannot be complete, since it is known that the only complete surfaces of constant positive curvature are the sphere and the projective plane. [60] However, in a sense it is only one point away from being a complete surface, as the open Möbius strip is homeomorphic to the once-punctured projective plane, the surface obtained by removing any one point from the projective plane. [63]

The minimal surfaces are described as having constant zero mean curvature instead of constant Gaussian curvature. The Sudanese Möbius strip was constructed as a minimal surface bounded by a great circle in a 3-sphere, but there is also a unique complete (boundaryless) minimal surface immersed in Euclidean space that has the topology of an open Möbius strip. It is called the Meeks Möbius strip, [64] after its 1982 description by William Hamilton Meeks, III. [65] Although globally unstable as a minimal surface, small patches of it, bounded by non-contractible curves within the surface, can form stable embedded Möbius strips as minimal surfaces. [66] Both the Meeks Möbius strip, and every higher-dimensional minimal surface with the topology of the Möbius strip, can be constructed using solutions to the Björling problem, which defines a minimal surface uniquely from its boundary curve and tangent planes along this curve. [67]

Spaces of lines

The family of lines in the plane can be given the structure of a smooth space, with each line represented as a point in this space. The resulting space of lines is topologically equivalent to the open Möbius strip. [68] One way to see this is to extend the Euclidean plane to the real projective plane by adding one more line, the line at infinity. By projective duality the space of lines in the projective plane is equivalent to its space of points, the projective plane itself. Removing the line at infinity, to produce the space of Euclidean lines, punctures this space of projective lines. [69] Therefore, the space of Euclidean lines is a punctured projective plane, which is one of the forms of the open Möbius strip. [63] The space of lines in the hyperbolic plane can be parameterized by unordered pairs of distinct points on a circle, the pairs of points at infinity of each line. This space, again, has the topology of an open Möbius strip. [70]

These spaces of lines are highly symmetric. The symmetries of Euclidean lines include the affine transformations, and the symmetries of hyperbolic lines include the Möbius transformations. [71] The affine transformations and Möbius transformations both form 6-dimensional Lie groups, topological spaces having a compatible algebraic structure describing the composition of symmetries. [72] [73] Because every line in the plane is symmetric to every other line, the open Möbius strip is a homogeneous space, a space with symmetries that take every point to every other point. Homogeneous spaces of Lie groups are called solvmanifolds, and the Möbius strip can be used as a counterexample, showing that not every solvmanifold is a nilmanifold, and that not every solvmanifold can be factored into a direct product of a compact solvmanifold with . These symmetries also provide another way to construct the Möbius strip itself, as a group model of these Lie groups. A group model consists of a Lie group and a stabilizer subgroup of its action; contracting the cosets of the subgroup to points produces a space with the same topology as the underlying homogenous space. In the case of the symmetries of Euclidean lines, the stabilizer of the -axis consists of all symmetries that take the axis to itself. Each line corresponds to a coset, the set of symmetries that map to the -axis. Therefore, the quotient space, a space that has one point per coset and inherits its topology from the space of symmetries, is the same as the space of lines, and is again an open Möbius strip. [74]

Applications

Electrical flow in a Mobius resistor Mobius resistor.svg
Electrical flow in a Möbius resistor

Beyond the already-discussed applications of Möbius strips to the design of mechanical belts that wear evenly on their entire surface, and of the Plücker conoid to the design of gears, other applications of Möbius strips include:

Scientists have also studied the energetics of soap films shaped as Möbius strips, [88] [89] the chemical synthesis of molecules with a Möbius strip shape, [90] [91] and the formation of larger nanoscale Möbius strips using DNA origami. [92]

Endless Twist, Max Bill, 1956, from the Middelheim Open Air Sculpture Museum Middelheim Max Bill Eindeloze kronkel 1956 03 Cropped.jpg
Endless Twist, Max Bill, 1956, from the Middelheim Open Air Sculpture Museum

Two-dimensional artworks featuring the Möbius strip include an untitled 1947 painting by Corrado Cagli (memorialized in a poem by Charles Olson), [93] [94] and two prints by M. C. Escher: Möbius Band I (1961), depicting three folded flatfish biting each others' tails; and Möbius Band II (1963), depicting ants crawling around a lemniscate-shaped Möbius strip. [95] [96] It is also a popular subject of mathematical sculpture, including works by Max Bill (Endless Ribbon, 1953), José de Rivera ( Infinity , 1967), and Sebastián. [93] A trefoil-knotted Möbius strip was used in John Robinson 's Immortality (1982). [97] Charles O. Perry's Continuum (1976) is one of several pieces by Perry exploring variations of the Möbius strip. [98]

Logo of Google Drive (2012-2014).svg
Google Drive logo (2012–2014)
Stamp of Brazil - 1967 - Colnect 263101 - Mobius Symbol.jpeg
IMPA logo on stamp

Because of their easily recognized form, Möbius strips are a common element of graphic design. [97] The familiar three-arrow logo for recycling, designed in 1970, is based on the smooth triangular form of the Möbius strip, [99] as was the logo for the environmentally-themed Expo '74. [100] Some variations of the recycling symbol use a different embedding with three half-twists instead of one, [99] and the original version of the Google Drive logo used a flat-folded three-twist Möbius strip, as have other similar designs. [101] The Brazilian Instituto Nacional de Matemática Pura e Aplicada (IMPA) uses a stylized smooth Möbius strip as their logo, and has a matching large sculpture of a Möbius strip on display in their building. [102] The Möbius strip has also featured in the artwork for postage stamps from countries including Brazil, Belgium, the Netherlands, and Switzerland. [103] [104]

NASCAR Hall of Fame entrance NASCAR Hall of Fame (7553589908).jpg
NASCAR Hall of Fame entrance

Möbius strips have been a frequent inspiration for the architectural design of buildings and bridges. However, many of these are projects or conceptual designs rather than constructed objects, or stretch their interpretation of the Möbius strip beyond its recognizability as a mathematical form or a functional part of the architecture. [105] [106] An example is the National Library of Kazakhstan, for which a building was planned in the shape of a thickened Möbius strip but refinished with a different design after the original architects pulled out of the project. [107] One notable building incorporating a Möbius strip is the NASCAR Hall of Fame, which is surrounded by a large twisted ribbon of stainless steel acting as a façade and canopy, and evoking the curved shapes of racing tracks. [108] On a smaller scale, Moebius Chair (2006) by Pedro Reyes is a courting bench whose base and sides have the form of a Möbius strip. [109] As a form of mathematics and fiber arts, scarves have been knit into Möbius strips since the work of Elizabeth Zimmermann in the early 1980s. [110] In food styling, Möbius strips have been used for slicing bagels, [111] making loops out of bacon, [112] and creating new shapes for pasta. [113]

Although mathematically the Möbius strip and the fourth dimension are both purely spatial concepts, they have often been invoked in speculative fiction as the basis for a time loop into which unwary victims may become trapped. Examples of this trope include Martin Gardner 's "No-Sided Professor" (1946), Armin Joseph Deutsch 's "A Subway Named Mobius" (1950) and the film Moebius (1996) based on it. An entire world shaped like a Möbius strip is the setting of Arthur C. Clarke's "The Wall of Darkness" (1946), while conventional Möbius strips are used as clever inventions in multiple stories of William Hazlett Upson from the 1940s. [114] Other works of fiction have been analyzed as having a Möbius strip–like structure, in which elements of the plot repeat with a twist; these include Marcel Proust 's In Search of Lost Time (1913–1927), Luigi Pirandello 's Six Characters in Search of an Author (1921), Frank Capra 's It's a Wonderful Life (1946), John Barth 's Lost in the Funhouse (1968), Samuel R. Delany 's Dhalgren (1975) and the film Donnie Darko (2001). [115]

One of the musical canons by J. S. Bach, the fifth of 14 canons (BWV 1087) discovered in 1974 in Bach's copy of the Goldberg Variations , features a glide-reflect symmetry in which each voice in the canon repeats, with inverted notes, the same motif from two measures earlier. Because of this symmetry, this canon can be thought of as having its score written on a Möbius strip. [116] [lower-alpha 8] In music theory, tones that differ by an octave are generally considered to be equivalent notes, and the space of possible notes forms a circle, the chromatic circle. Because the Möbius strip is the configuration space of two unordered points on a circle, the space of all two-note chords takes the shape of a Möbius strip. This conception, and generalizations to more points, is a significant application of orbifolds to music theory. [117] [118] Modern musical groups taking their name from the Möbius strip include American electronic rock trio Mobius Band [119] and Norwegian progressive rock band Ring Van Möbius. [120]

Möbius strips and their properties have been used in the design of stage magic. One such trick, known as the Afghan bands, uses the fact that the Möbius strip remains a single strip when cut lengthwise. It originated in the 1880s, and was very popular in the first half of the twentieth century. Many versions of this trick exist and have been performed by famous illusionists such as Harry Blackstone Sr. and Thomas Nelson Downs. [121] [122]

See also

Notes

  1. Pronounced US: /ˈmbiəs,ˈm-/ MOH-bee-əs, MAY-, UK: /ˈmɜːbiəs/ ; [1] German: [ˈmøːbi̯ʊs] . As is common for words containing an umlaut, it is also often spelled Mobius or Moebius.
  2. Essentially this example, but for a Klein bottle rather than a Möbius strip, is given by Blackett (1982). [8]
  3. The length of a strip can be measured at its centerline, or by cutting the resulting Möbius strip perpendicularly to its boundary so that it forms a rectangle
  4. The flat-folded Möbius strip formed from three equilateral triangles does not come from an abstract simplicial complex, because all three triangles share the same three vertices, while abstract simplicial complexes require each triangle to have a different set of vertices.
  5. This piecewise planar and cylindrical embedding has smoothness class , and can be approximated arbitrarily accurately by infinitely differentiable (class ) embeddings. [39]
  6. 12/7 is the simplest rational number in the range of aspect ratios, between 1.695 and 1.73, for which the existence of a smooth embedding is unknown.
  7. These surfaces have smoothness class . For a more fine-grained analysis of the smoothness assumptions that force an embedding to be developable versus the assumptions under which the Nash–Kuiper theorem allows arbitrarily flexible embeddings, see remarks by Bartels & Hornung (2015), p. 116, following Theorem 2.2. [39]
  8. Möbius strips have also been used to analyze many other canons by Bach and others, but in most of these cases other looping surfaces such as a cylinder could have been used equally well. [116]

Related Research Articles

<span class="mw-page-title-main">Differential topology</span> Branch of mathematics

In mathematics, differential topology is the field dealing with the topological properties and smooth properties of smooth manifolds. In this sense differential topology is distinct from the closely related field of differential geometry, which concerns the geometric properties of smooth manifolds, including notions of size, distance, and rigid shape. By comparison differential topology is concerned with coarser properties, such as the number of holes in a manifold, its homotopy type, or the structure of its diffeomorphism group. Because many of these coarser properties may be captured algebraically, differential topology has strong links to algebraic topology.

<span class="mw-page-title-main">Differential geometry</span> Branch of mathematics dealing with functions and geometric structures on differentiable manifolds

Differential geometry is a mathematical discipline that studies the geometry of smooth shapes and smooth spaces, otherwise known as smooth manifolds. It uses the techniques of differential calculus, integral calculus, linear algebra and multilinear algebra. The field has its origins in the study of spherical geometry as far back as antiquity. It also relates to astronomy, the geodesy of the Earth, and later the study of hyperbolic geometry by Lobachevsky. The simplest examples of smooth spaces are the plane and space curves and surfaces in the three-dimensional Euclidean space, and the study of these shapes formed the basis for development of modern differential geometry during the 18th and 19th centuries.

<span class="mw-page-title-main">Klein bottle</span> Non-orientable mathematical surface

In mathematics, the Klein bottle is an example of a non-orientable surface; that is, informally, a one-sided surface which, if traveled upon, could be followed back to the point of origin while flipping the traveler upside down. More formally, the Klein bottle is a two-dimensional manifold on which one cannot define a normal vector at each point that varies continuously over the whole manifold. Other related non-orientable surfaces include the Möbius strip and the real projective plane. While a Möbius strip is a surface with a boundary, a Klein bottle has no boundary. For comparison, a sphere is an orientable surface with no boundary.

<span class="mw-page-title-main">Surface (topology)</span> Two-dimensional manifold

In the part of mathematics referred to as topology, a surface is a two-dimensional manifold. Some surfaces arise as the boundaries of three-dimensional solid figures; for example, the sphere is the boundary of the solid ball. Other surfaces arise as graphs of functions of two variables; see the figure at right. However, surfaces can also be defined abstractly, without reference to any ambient space. For example, the Klein bottle is a surface that cannot be embedded in three-dimensional Euclidean space.

<span class="mw-page-title-main">Topology</span> Branch of mathematics

Topology is the mathematical field of study concerned with the properties of a geometric object that are preserved under continuous deformations, such as stretching, twisting, crumpling, and bending; that is, without closing holes, opening holes, tearing, gluing, or passing through itself.

<span class="mw-page-title-main">Torus</span> Doughnut-shaped surface of revolution

In geometry, a torus is a surface of revolution generated by revolving a circle in three-dimensional space one full revolution about an axis that is coplanar with the circle. The main types of toruses include ring toruses, horn toruses, and spindle toruses. A ring torus is sometimes colloquially referred to as a donut or doughnut.

<span class="mw-page-title-main">Orientability</span> Possibility of a consistent definition of "clockwise" in a mathematical space

In mathematics, orientability is a property of some topological spaces such as real vector spaces, Euclidean spaces, surfaces, and more generally manifolds that allows a consistent definition of "clockwise" and "anticlockwise". A space is orientable if such a consistent definition exists. In this case, there are two possible definitions, and a choice between them is an orientation of the space. Real vector spaces, Euclidean spaces, and spheres are orientable. A space is non-orientable if "clockwise" is changed into "counterclockwise" after running through some loops in it, and coming back to the starting point. This means that a geometric shape, such as , that moves continuously along such a loop is changed into its own mirror image . A Möbius strip is an example of a non-orientable space.

In mathematics, the Chern theorem states that the Euler–Poincaré characteristic of a closed even-dimensional Riemannian manifold is equal to the integral of a certain polynomial of its curvature form.

<span class="mw-page-title-main">Real projective plane</span> Compact non-orientable two-dimensional manifold

In mathematics, the real projective plane is an example of a compact non-orientable two-dimensional manifold; in other words, a one-sided surface. It cannot be embedded in standard three-dimensional space without intersecting itself. It has basic applications to geometry, since the common construction of the real projective plane is as the space of lines in R3 passing through the origin. The real projective plane is then an extension of the (ordinary) plane — to every point (v1,v2) of the ordinary plane, the line spanned by (v1,v2,1) is associated (i.e., the real projective plane is the projective completion of the ordinary plane, cf. also the homogeneous coordinates below) while there are also some “points in the infinity”.

<span class="mw-page-title-main">Geometric topology</span> Branch of mathematics studying (smooth) functions of manifolds

In mathematics, geometric topology is the study of manifolds and maps between them, particularly embeddings of one manifold into another.

<span class="mw-page-title-main">Low-dimensional topology</span> Branch of topology

In mathematics, low-dimensional topology is the branch of topology that studies manifolds, or more generally topological spaces, of four or fewer dimensions. Representative topics are the structure theory of 3-manifolds and 4-manifolds, knot theory, and braid groups. This can be regarded as a part of geometric topology. It may also be used to refer to the study of topological spaces of dimension 1, though this is more typically considered part of continuum theory.

<span class="mw-page-title-main">3-manifold</span> Mathematical space

In mathematics, a 3-manifold is a topological space that locally looks like a three-dimensional Euclidean space. A 3-manifold can be thought of as a possible shape of the universe. Just as a sphere looks like a plane to a small and close enough observer, all 3-manifolds look like our universe does to a small enough observer. This is made more precise in the definition below.

<span class="mw-page-title-main">Willmore energy</span>

In differential geometry, the Willmore energy is a quantitative measure of how much a given surface deviates from a round sphere. Mathematically, the Willmore energy of a smooth closed surface embedded in three-dimensional Euclidean space is defined to be the integral of the square of the mean curvature minus the Gaussian curvature. It is named after the English geometer Thomas Willmore.

<span class="mw-page-title-main">Manifold</span> Topological space that locally resembles Euclidean space

In mathematics, a manifold is a topological space that locally resembles Euclidean space near each point. More precisely, an -dimensional manifold, or -manifold for short, is a topological space with the property that each point has a neighborhood that is homeomorphic to an open subset of -dimensional Euclidean space.

In topology, a branch of mathematics, a topological manifold is a topological space that locally resembles real n-dimensional Euclidean space. Topological manifolds are an important class of topological spaces, with applications throughout mathematics. All manifolds are topological manifolds by definition. Other types of manifolds are formed by adding structure to a topological manifold. Every manifold has an "underlying" topological manifold, obtained by simply "forgetting" the added structure. However, not every topological manifold can be endowed with a particular additional structure. For example, the E8 manifold is a topological manifold which cannot be endowed with a differentiable structure.

<span class="mw-page-title-main">Surface (mathematics)</span> Mathematical idealization of the surface of a body

In mathematics, a surface is a mathematical model of the common concept of a surface. It is a generalization of a plane, but, unlike a plane, it may be curved; this is analogous to a curve generalizing a straight line.

<span class="mw-page-title-main">Immersion (mathematics)</span> Differentiable function whose derivative is everywhere injective

In mathematics, an immersion is a differentiable function between differentiable manifolds whose differential pushforward is everywhere injective. Explicitly, f : MN is an immersion if

In mathematics, a Riemannian manifold is said to be flat if its Riemann curvature tensor is everywhere zero. Intuitively, a flat manifold is one that "locally looks like" Euclidean space in terms of distances and angles, e.g. the interior angles of a triangle add up to 180°.

<span class="mw-page-title-main">Differential geometry of surfaces</span> The mathematics of smooth surfaces

In mathematics, the differential geometry of surfaces deals with the differential geometry of smooth surfaces with various additional structures, most often, a Riemannian metric. Surfaces have been extensively studied from various perspectives: extrinsically, relating to their embedding in Euclidean space and intrinsically, reflecting their properties determined solely by the distance within the surface as measured along curves on the surface. One of the fundamental concepts investigated is the Gaussian curvature, first studied in depth by Carl Friedrich Gauss, who showed that curvature was an intrinsic property of a surface, independent of its isometric embedding in Euclidean space.

In mathematics, a plane is a two-dimensional space or flat surface that extends indefinitely. A plane is the two-dimensional analogue of a point, a line and three-dimensional space.

References

  1. Wells, John C. (2008). Longman Pronunciation Dictionary (3rd ed.). Longman. ISBN   978-1-4058-8118-0.
  2. Pickover, Clifford A. (2005). The Möbius Strip: Dr. August Möbius's Marvelous Band in Mathematics, Games, Literature, Art, Technology, and Cosmology. Thunder's Mouth Press. pp. 28–29. ISBN   978-1-56025-826-1.
  3. 1 2 3 4 Larison, Lorraine L. (1973). "The Möbius band in Roman mosaics". American Scientist . 61 (5): 544–547. Bibcode:1973AmSci..61..544L. JSTOR   27843983.
  4. 1 2 3 Cartwright, Julyan H. E.; González, Diego L. (2016). "Möbius strips before Möbius: topological hints in ancient representations". The Mathematical Intelligencer . 38 (2): 69–76. arXiv: 1609.07779 . Bibcode:2016arXiv160907779C. doi:10.1007/s00283-016-9631-8. MR   3507121. S2CID   119587191.
  5. Flapan, Erica (2000). When Topology Meets Chemistry: A Topological Look at Molecular Chirality. Outlooks. Washington, DC: Mathematical Association of America. pp.  82–83. doi:10.1017/CBO9780511626272. ISBN   0-521-66254-0. MR   1781912.
  6. 1 2 3 Pickover (2005), pp. 8–9.
  7. Woll, John W. Jr. (Spring 1971). "One-sided surfaces and orientability". The Two-Year College Mathematics Journal . 2 (1): 5–18. doi:10.2307/3026946. JSTOR   3026946.
  8. Blackett, Donald W. (1982). Elementary Topology: A Combinatorial and Algebraic Approach. Academic Press. p. 195. ISBN   9781483262536.
  9. Frolkina, Olga D. (2018). "Pairwise disjoint Moebius bands in space". Journal of Knot Theory and Its Ramifications . 27 (9): 1842005, 9. arXiv: 2212.02983 . doi:10.1142/S0218216518420051. MR   3848635. S2CID   126421578.
  10. Lamb, Evelyn (February 20, 2019). "Möbius strips defy a link with infinity". Quanta Magazine .
  11. Melikhov, Sergey A. (2019). "A note on O. Frolkina's paper "Pairwise disjoint Moebius bands in space"". Journal of Knot Theory and Its Ramifications . 28 (7): 1971001, 3. arXiv: 1810.04089 . doi:10.1142/s0218216519710019. MR   3975576. S2CID   119179202.
  12. Pickover (2005), p. 52.
  13. Pickover (2005), p. 12.
  14. Kyle, R. H. (1955). "Embeddings of Möbius bands in 3-dimensional space". Proceedings of the Royal Irish Academy, Section A. 57: 131–136. JSTOR   20488581. MR   0091480.
  15. Pickover (2005), p. 11.
  16. Massey, William S. (1991). A Basic Course in Algebraic Topology. Graduate Texts in Mathematics. Vol. 127. New York: Springer-Verlag. p. 49. ISBN   0-387-97430-X. MR   1095046.
  17. Rouse Ball, W. W. (1892). "Paradromic rings". Mathematical Recreations and Problems of Past and Present Times (2nd ed.). London & New York: Macmillan and co. pp. 53–54. ISBN   9780608377803.
  18. Bennett, G. T. (June 1923). "Paradromic rings". Nature . 111 (2800): 882. Bibcode:1923Natur.111R.882B. doi: 10.1038/111882b0 . S2CID   4099647.
  19. 1 2 Tietze, Heinrich (1910). "Einige Bemerkungen zum Problem des Kartenfärbens auf einseitigen Flächen" (PDF). Jahresbericht der Deutschen Mathematiker-Vereinigung. 19: 155–159.
  20. Ringel, G.; Youngs, J. W. T. (1968). "Solution of the Heawood map-coloring problem". Proceedings of the National Academy of Sciences of the United States of America . 60 (2): 438–445. Bibcode:1968PNAS...60..438R. doi: 10.1073/pnas.60.2.438 . MR   0228378. PMC   225066 . PMID   16591648.
  21. Jablan, Slavik; Radović, Ljiljana; Sazdanović, Radmila (2011). "Nonplanar graphs derived from Gauss codes of virtual knots and links". Journal of Mathematical Chemistry. 49 (10): 2250–2267. doi:10.1007/s10910-011-9884-6. MR   2846715. S2CID   121332704.
  22. Larsen, Mogens Esrom (1994). "Misunderstanding my mazy mazes may make me miserable". In Guy, Richard K.; Woodrow, Robert E. (eds.). Proceedings of the Eugène Strens Memorial Conference on Recreational Mathematics and its History held at the University of Calgary, Calgary, Alberta, August 1986. MAA Spectrum. Washington, DC: Mathematical Association of America. pp. 289–293. ISBN   0-88385-516-X. MR   1303141.. See Figure 7, p. 292.
  23. Maschke, Heinrich (1900). "Note on the unilateral surface of Moebius". Transactions of the American Mathematical Society . 1 (1): 39. doi: 10.2307/1986401 . JSTOR   1986401. MR   1500522.
  24. Junghenn, Hugo D. (2015). A Course in Real Analysis. Boca Raton, Florida: CRC Press. p. 430. ISBN   978-1-4822-1927-2. MR   3309241.
  25. Séquin, Carlo H. (2005). "Splitting tori, knots, and Moebius bands". In Sarhangi, Reza; Moody, Robert V. (eds.). Renaissance Banff: Mathematics, Music, Art, Culture. Southwestern College, Winfield, Kansas: Bridges Conference. pp. 211–218. ISBN   0-9665201-6-5.
  26. 1 2 Francis, George K. (1987). "Plücker conoid". A Topological Picturebook. Springer-Verlag, New York. pp. 81–83. ISBN   0-387-96426-6. MR   0880519.
  27. Dooner, David B.; Seireg, Ali (1995). "3.4.2 The cylindroid". The Kinematic Geometry of Gearing: A Concurrent Engineering Approach. Wiley Series in Design Engineering. Vol. 3. John Wiley & Sons. pp. 135–137. ISBN   9780471045977.
  28. 1 2 Barr, Stephen (1964). Experiments in Topology. New York: Thomas Y. Crowell Company. pp. 40–49, 200–201. ISBN   9780690278620.
  29. 1 2 3 4 5 Fuchs, Dmitry; Tabachnikov, Serge (2007). "Lecture 14: Paper Möbius band". Mathematical Omnibus: Thirty Lectures on Classic Mathematics (PDF). Providence, Rhode Island: American Mathematical Society. pp. 199–206. doi:10.1090/mbk/046. ISBN   978-0-8218-4316-1. MR   2350979. Archived from the original (PDF) on 2016-04-24.
  30. Pook, Les (2003). "4.2: The trihexaflexagon revisited". Flexagons Inside Out. Cambridge, UK: Cambridge University Press. pp. 33–36. doi:10.1017/CBO9780511543302. ISBN   0-521-81970-9. MR   2008500.
  31. 1 2 Kühnel, W.; Banchoff, T. F. (1983). "The 9-vertex complex projective plane" (PDF). The Mathematical Intelligencer . 5 (3): 11–22. doi:10.1007/BF03026567. MR   0737686. S2CID   120926324.
  32. Kuiper, Nicolaas H. (1972). "Tight topological embeddings of the Moebius band". Journal of Differential Geometry . 6 (3): 271–283. doi: 10.4310/jdg/1214430493 . MR   0314057.
  33. Szilassi, Lajos (2008). "A polyhedral model in Euclidean 3-space of the six-pentagon map of the projective plane". Discrete & Computational Geometry . 40 (3): 395–400. doi: 10.1007/s00454-007-9033-y . MR   2443291. S2CID   38606607.
  34. Tuckerman, Bryant (1948). "A non-singular polyhedral Möbius band whose boundary is a triangle". American Mathematical Monthly. 55 (5): 309–311. doi:10.2307/2305482. JSTOR   2305482. MR   0024138.
  35. Bonnington, C. Paul; Nakamoto, Atsuhiro (2008). "Geometric realization of a triangulation on the projective plane with one face removed". Discrete & Computational Geometry . 40 (1): 141–157. doi: 10.1007/s00454-007-9035-9 . MR   2429652. S2CID   10887519.
  36. Brehm, Ulrich (1983). "A nonpolyhedral triangulated Möbius strip". Proceedings of the American Mathematical Society . 89 (3): 519–522. doi:10.2307/2045508. JSTOR   2045508. MR   0715878.
  37. Nakamoto, Atsuhiro; Tsuchiya, Shoichi (2012). "On geometrically realizable Möbius triangulations". Discrete Mathematics . 312 (14): 2135–2139. doi: 10.1016/j.disc.2011.06.007 . MR   2921579.
  38. 1 2 Hinz, Denis F.; Fried, Eliot (2015). "Translation of Michael Sadowsky's paper "An elementary proof for the existence of a developable Möbius band and the attribution of the geometric problem to a variational problem"". Journal of Elasticity. 119 (1–2): 3–6. arXiv: 1408.3034 . doi:10.1007/s10659-014-9490-5. MR   3326180. S2CID   119733903. Reprinted in Fosdick, Roger; Fried, Eliot (2016). The Mechanics of Ribbons and Möbius Bands (PDF). Springer, Dordrecht. pp. 3–6. doi:10.1007/978-94-017-7300-3. ISBN   978-94-017-7299-0. MR   3381564.
  39. 1 2 3 4 Bartels, Sören; Hornung, Peter (2015). "Bending paper and the Möbius strip". Journal of Elasticity. 119 (1–2): 113–136. doi:10.1007/s10659-014-9501-6. MR   3326187. S2CID   119782792. Reprinted in Fosdick & Fried (2016), pp. 113–136. See in particular Section 5.2, pp. 129–130.
  40. 1 2 Starostin, E. L.; van der Heijden, G. H. M. (2015). "Equilibrium shapes with stress localisation for inextensible elastic Möbius and other strips". Journal of Elasticity. 119 (1–2): 67–112. doi: 10.1007/s10659-014-9495-0 . MR   3326186. S2CID   53462568. Reprinted in Fosdick & Fried (2016), pp. 67–112.
  41. Schwarz, Gideon E. (1990). "The dark side of the Moebius strip". The American Mathematical Monthly . 97 (10): 890–897. doi:10.1080/00029890.1990.11995680. JSTOR   2324325. MR   1079975.
  42. 1 2 Halpern, B.; Weaver, C. (1977). "Inverting a cylinder through isometric immersions and isometric embeddings". Transactions of the American Mathematical Society. 230: 41–70. doi: 10.2307/1997711 . JSTOR   1997711. MR   0474388.
  43. 1 2 Schwartz, Richard Evan (2021). "An improved bound on the optimal paper Moebius band". Geometriae Dedicata . 215: 255–267. arXiv: 2008.11610 . doi:10.1007/s10711-021-00648-5. MR   4330341. S2CID   220279013.
  44. Schwartz, Richard (2023). "The optimal paper Moebius band". arXiv: 2308.12641 [math.MG].
  45. Crowell, Rachel (September 12, 2023). "Mathematicians solve 50-year-old Möbius strip puzzle". Scientific American.
  46. 1 2 Dundas, Bjørn Ian (2018). "Example 5.1.3: The unbounded Möbius band". A Short Course in Differential Topology. Cambridge Mathematical Textbooks. Cambridge University Press, Cambridge. p.  https://books.google.com/books?id=7a1eDwAAQBAJ&pg=PA101. doi:10.1017/9781108349130. ISBN   978-1-108-42579-7. MR   3793640. S2CID   125997451.
  47. Blanuša, Danilo (1954). "Le plongement isométrique de la bande de Möbius infiniment large euclidienne dans un espace sphérique, parabolique ou hyperbolique à quatre dimensions". Bulletin International de l'Académie Yougoslave des Sciences et des Beaux-Arts. 12: 19–23. MR   0071060.
  48. Wunderlich, W. (1962). "Über ein abwickelbares Möbiusband". Monatshefte für Mathematik . 66 (3): 276–289. doi:10.1007/BF01299052. MR   0143115. S2CID   122215321.
  49. Schwarz, Gideon (1990). "A pretender to the title 'canonical Moebius strip'". Pacific Journal of Mathematics . 143 (1): 195–200. doi: 10.2140/pjm.1990.143.195 . MR   1047406.
  50. Hilbert, David; Cohn-Vossen, Stephan (1952). Geometry and the Imagination (2nd ed.). Chelsea. pp. 315–316. ISBN   978-0-8284-1087-8.
  51. Spivak, Michael (1979). A Comprehensive Introduction to Differential Geometry, Volume I (2nd ed.). Wilmington, Delaware: Publish or Perish. p. 591.
  52. 1 2 3 Knöppel, Felix (Summer 2019). "Tutorial 3: Lawson's Minimal Surfaces and the Sudanese Möbius Band". DDG2019: Visualization course at TU Berlin.
  53. 1 2 Lawson, H. Blaine Jr. (1970). "Complete minimal surfaces in ". Annals of Mathematics . Second Series. 92 (3): 335–374. doi:10.2307/1970625. JSTOR   1970625. MR   0270280. See Section 7, pp. 350–353, where the Klein bottle is denoted .
  54. 1 2 Schleimer, Saul; Segerman, Henry (2012). "Sculptures in S3". In Bosch, Robert; McKenna, Douglas; Sarhangi, Reza (eds.). Proceedings of Bridges 2012: Mathematics, Music, Art, Architecture, Culture. Phoenix, Arizona: Tessellations Publishing. pp. 103–110. arXiv: 1204.4952 . ISBN   978-1-938664-00-7.
  55. Gunn, Charles (August 23, 2018). "Sudanese Möbius Band". Vimeo. Retrieved 2022-03-17.
  56. Franzoni, Gregorio (2012). "The Klein bottle: variations on a theme". Notices of the American Mathematical Society. 59 (8): 1076–1082. doi: 10.1090/noti880 . MR   2985809.
  57. Huggett, Stephen; Jordan, David (2009). A Topological Aperitif (Revised ed.). Springer-Verlag. p. 57. ISBN   978-1-84800-912-7. MR   2483686.
  58. Flapan, Erica (2016). Knots, Molecules, and the Universe: An Introduction to Topology. Providence, Rhode Island: American Mathematical Society. pp. 99–100. doi:10.1090/mbk/096. ISBN   978-1-4704-2535-7. MR   3443369.
  59. Richeson, David S. (2008). Euler's Gem: The Polyhedron Formula and the Birth of Topology. Princeton, New Jersey: Princeton University Press. p.  171. ISBN   978-0-691-12677-7. MR   2440945.
  60. 1 2 Godinho, Leonor; Natário, José (2014). An Introduction to Riemannian Geometry: With Applications to Mechanics and Relativity. Universitext. Springer, Cham. pp. 152–153. doi:10.1007/978-3-319-08666-8. ISBN   978-3-319-08665-1. MR   3289090.
  61. Cantwell, John; Conlon, Lawrence (2015). "Hyperbolic geometry and homotopic homeomorphisms of surfaces". Geometriae Dedicata . 177: 27–42. arXiv: 1305.1379 . doi:10.1007/s10711-014-9975-1. MR   3370020. S2CID   119640200.
  62. Stillwell, John (1992). "4.6 Classification of isometries". Geometry of Surfaces. Universitext. Cham: Springer. pp. 96–98. doi:10.1007/978-1-4612-0929-4. ISBN   0-387-97743-0. MR   1171453.
  63. 1 2 Seifert, Herbert; Threlfall, William (1980). A Textbook of Topology. Pure and Applied Mathematics. Vol. 89. Translated by Goldman, Michael A. New York & London: Academic Press. p. 12. ISBN   0-12-634850-2. MR   0575168.
  64. López, Francisco J.; Martín, Francisco (1997). "Complete nonorientable minimal surfaces with the highest symmetry group". American Journal of Mathematics . 119 (1): 55–81. doi:10.1353/ajm.1997.0004. MR   1428058. S2CID   121366986.
  65. Meeks, William H. III (1981). "The classification of complete minimal surfaces in with total curvature greater than ". Duke Mathematical Journal . 48 (3): 523–535. doi:10.1215/S0012-7094-81-04829-8. MR   0630583.
  66. Pesci, Adriana I.; Goldstein, Raymond E.; Alexander, Gareth P.; Moffatt, H. Keith (2015). "Instability of a Möbius strip minimal surface and a link with systolic geometry" (PDF). Physical Review Letters . 114 (12): 127801. Bibcode:2015PhRvL.114l7801P. doi:10.1103/PhysRevLett.114.127801. MR   3447638. PMID   25860771.
  67. Mira, Pablo (2006). "Complete minimal Möbius strips in and the Björling problem". Journal of Geometry and Physics. 56 (9): 1506–1515. Bibcode:2006JGP....56.1506M. doi:10.1016/j.geomphys.2005.08.001. MR   2240407.
  68. Parker, Phillip E. (1993). "Spaces of geodesics". In Del Riego, L. (ed.). Differential Geometry Workshop on Spaces of Geometry (Guanajuato, 1992). Aportaciones Mat. Notas Investigación. Vol. 8. Soc. Mat. Mexicana, México. pp. 67–79. MR   1304924.
  69. Bickel, Holger (1999). "Duality in stable planes and related closure and kernel operations". Journal of Geometry. 64 (1–2): 8–15. doi:10.1007/BF01229209. MR   1675956. S2CID   122209943.
  70. Mangahas, Johanna (July 2017). "Office Hour Five: The Ping-Pong Lemma". In Clay, Matt; Margalit, Dan (eds.). Office Hours with a Geometric Group Theorist. Princeton University Press. pp. 85–105. doi:10.1515/9781400885398. ISBN   9781400885398. See in particular Project 7, pp. 104–105.
  71. Ramírez Galarza, Ana Irene; Seade, José (2007). Introduction to Classical Geometries. Basel: Birkhäuser Verlag. pp. 83–88, 157–163. ISBN   978-3-7643-7517-1. MR   2305055.
  72. Fomenko, Anatolij T.; Kunii, Tosiyasu L. (2013). Topological Modeling for Visualization. Springer. p. 269. ISBN   9784431669562.
  73. Isham, Chris J. (1999). Modern Differential Geometry for Physicists. World Scientific lecture notes in physics. Vol. 61 (2nd ed.). World Scientific. p. 269. ISBN   981-02-3555-0. MR   1698234.
  74. Gorbatsevich, V. V.; Onishchik, A. L.; Vinberg, È. B. (1993). Lie groups and Lie algebras I: Foundations of Lie Theory; Lie Transformation Groups. Encyclopaedia of Mathematical Sciences. Vol. 20. Springer-Verlag, Berlin. pp. 164–166. doi:10.1007/978-3-642-57999-8. ISBN   3-540-18697-2. MR   1306737.
  75. Yamashiro, Atsushi; Shimoi, Yukihiro; Harigaya, Kikuo; Wakabayashi, Katsunori (2004). "Novel electronic states in graphene ribbons: competing spin and charge orders". Physica E. 22 (1–3): 688–691. arXiv: cond-mat/0309636 . Bibcode:2004PhyE...22..688Y. doi:10.1016/j.physe.2003.12.100. S2CID   17102453.
  76. Rzepa, Henry S. (September 2005). "Möbius aromaticity and delocalization". Chemical Reviews. 105 (10): 3697–3715. doi:10.1021/cr030092l. PMID   16218564.
  77. Yoon, Zin Seok; Osuka, Atsuhiro; Kim, Dongho (May 2009). "Möbius aromaticity and antiaromaticity in expanded porphyrins". Nature Chemistry. 1 (2): 113–122. Bibcode:2009NatCh...1..113Y. doi:10.1038/nchem.172. PMID   21378823.
  78. "Making resistors with math". Time . Vol. 84, no. 13. September 25, 1964.
  79. Pickover (2005), pp. 45–46.
  80. Pond, J. M. (2000). "Mobius dual-mode resonators and bandpass filters". IEEE Transactions on Microwave Theory and Techniques. 48 (12): 2465–2471. Bibcode:2000ITMTT..48.2465P. doi:10.1109/22.898999.
  81. Rohde, Ulrich L.; Poddar, Ajay; Sundararajan, D. (November 2013). "Printed resonators: Möbius strip theory and applications" (PDF). Microwave Journal. 56 (11).
  82. Bauer, Thomas; Banzer, Peter; Karimi, Ebrahim; Orlov, Sergej; Rubano, Andrea; Marrucci, Lorenzo; Santamato, Enrico; Boyd, Robert W.; Leuchs, Gerd (February 2015). "Observation of optical polarization Möbius strips". Science . 347 (6225): 964–966. Bibcode:2015Sci...347..964B. doi:10.1126/science.1260635. PMID   25636796. S2CID   206562350.
  83. Candeal, Juan Carlos; Induráin, Esteban (January 1994). "The Moebius strip and a social choice paradox". Economics Letters . 45 (3): 407–412. doi:10.1016/0165-1765(94)90045-0.
  84. Easdown, Martin (2012). Amusement Park Rides. Bloomsbury Publishing. p. 43. ISBN   9781782001522.
  85. Hook, Patrick (2019). Ticket To Ride: The Essential Guide to the World's Greatest Roller Coasters and Thrill Rides. Chartwell Books. p. 20. ISBN   9780785835776.
  86. Tobler, Waldo R. (1961). "A world map on a Möbius strip". Surveying & Mapping. 21: 486.
  87. Kumler, Mark P.; Tobler, Waldo R. (January 1991). "Three world maps on a Moebius strip". Cartography and Geographic Information Systems. 18 (4): 275–276. doi:10.1559/152304091783786781.
  88. Courant, Richard (1940). "Soap film experiments with minimal surfaces". The American Mathematical Monthly . 47 (3): 167–174. doi:10.1080/00029890.1940.11990957. JSTOR   2304225. MR   0001622.
  89. Goldstein, Raymond E.; Moffatt, H. Keith; Pesci, Adriana I.; Ricca, Renzo L. (December 2010). "Soap-film Möbius strip changes topology with a twist singularity". Proceedings of the National Academy of Sciences . 107 (51): 21979–21984. Bibcode:2010PNAS..10721979G. doi: 10.1073/pnas.1015997107 . PMC   3009808 .
  90. Walba, David M.; Richards, Rodney M.; Haltiwanger, R. Curtis (June 1982). "Total synthesis of the first molecular Moebius strip". Journal of the American Chemical Society . 104 (11): 3219–3221. doi:10.1021/ja00375a051.
  91. Pickover (2005), pp. 52–58.
  92. Gitig, Diana (October 18, 2010). "Chemical origami used to create a DNA Möbius strip". Ars Technica . Retrieved 2022-03-28.
  93. 1 2 Emmer, Michele (Spring 1980). "Visual art and mathematics: the Moebius band". Leonardo . 13 (2): 108–111. doi:10.2307/1577979. JSTOR   1577979. S2CID   123908555.
  94. Byers, Mark (2018). Charles Olson and American Modernism: The Practice of the Self. Oxford University Press. pp. 77–78. ISBN   9780198813255.
  95. Crato, Nuno (2010). "Escher and the Möbius strip". Figuring It Out: Entertaining Encounters with Everyday Math. Springer. pp. 123–126. doi:10.1007/978-3-642-04833-3_29.
  96. Kersten, Erik (March 13, 2017). "Möbius Strip I". Escher in the Palace . Retrieved 2022-04-17.
  97. 1 2 Pickover (2005), p. 13.
  98. Brecher, Kenneth (2017). "Art of infinity". In Swart, David; Séquin, Carlo H.; Fenyvesi, Kristóf (eds.). Proceedings of Bridges 2017: Mathematics, Art, Music, Architecture, Education, Culture. Phoenix, Arizona: Tessellations Publishing. pp. 153–158. ISBN   978-1-938664-22-9.
  99. 1 2 Peterson, Ivars (2002). "Recycling topology". Mathematical Treks: From Surreal Numbers to Magic Circles. MAA Spectrum. Mathematical Association of America, Washington, DC. pp. 31–35. ISBN   0-88385-537-2. MR   1874198.
  100. "Expo '74 symbol selected". The Spokesman-Review . March 12, 1972. p. 1.
  101. Millward, Steven (April 30, 2012). "Did Google Drive Copy its Icon From a Chinese App?". Tech in Asia. Retrieved 2022-03-27 via Yahoo! News.
  102. "Símbolo do IMPA". Para quem é fã do IMPA, dez curiosidades sobre o instituto. IMPA. May 7, 2020. Retrieved 2022-03-27.
  103. Pickover (2005), pp. 156–157.
  104. Decker, Heinz; Stark, Eberhard (1983). "Möbius-Bänder: ...und natürlich auch auf Briefmarken". Praxis der Mathematik. 25 (7): 207–215. MR   0720681.
  105. Thulaseedas, Jolly; Krawczyk, Robert J. (2003). "Möbius concepts in architecture". In Barrallo, Javier; Friedman, Nathaniel; Maldonado, Juan Antonio; Mart\'\inez-Aroza, José; Sarhangi, Reza; Séquin, Carlo (eds.). Meeting Alhambra, ISAMA-BRIDGES Conference Proceedings. Granada, Spain: University of Granada. pp. 353–360. ISBN   84-930669-1-5.
  106. Séquin, Carlo H. (January 2018). "Möbius bridges". Journal of Mathematics and the Arts . 12 (2–3): 181–194. doi:10.1080/17513472.2017.1419331. S2CID   216116708.
  107. Wainwright, Oliver (October 17, 2017). "'Norman said the president wants a pyramid': how starchitects built Astana". The Guardian .
  108. Muret, Don (May 17, 2010). "NASCAR Hall of Fame 'looks fast sitting still'". Sports Business Journal.
  109. Gopnik, Blake (October 17, 2014). "Pedro Reyes Makes an Infinite Love Seat". Artnet News.
  110. Thomas, Nancy J. (October 4, 1998). "Making a Mobius a matter of mathematics". The Times (Trenton) . p. aa3 via NewsBank.
  111. Pashman, Dan (August 6, 2015). "Cut Your Bagel The Mathematically Correct Way". The Salt. NPR.
  112. Miller, Ross (September 5, 2014). "How to make a mathematically-endless strip of bacon". The Verge.
  113. Chang, Kenneth (January 9, 2012). "Pasta Graduates From Alphabet Soup to Advanced Geometry". The New York Times .
  114. Pickover (2005), pp. 174–177.
  115. Pickover (2005), pp. 179–187.
  116. 1 2 Phillips, Tony (November 25, 2016). "Bach and the musical Möbius strip". Plus Magazine . Reprinted from an American Mathematical Society Feature Column.
  117. Moskowitz, Clara (May 6, 2008). "Music reduced to beautiful math". Live Science . Retrieved 2022-03-21.
  118. Tymoczko, Dmitri (July 7, 2006). "The geometry of musical chords" (PDF). Science . 313 (5783): 72–4. Bibcode:2006Sci...313...72T. doi:10.1126/science.1126287. JSTOR   3846592. PMID   16825563. S2CID   2877171.
  119. Parks, Andrew (August 30, 2007). "Mobius Band: Friendly Fire". Magnet .
  120. Lawson, Dom (February 9, 2021). "Ring Van Möbius". Prog .
  121. Prevos, Peter (2018). The Möbius Strip in Magic: A Treatise on the Afghan Bands. Kangaroo Flat: Third Hemisphere.
  122. Gardner, Martin (1956). "The Afghan Bands". Mathematics, Magic and Mystery. New York: Dover Books. pp. 70–73.