Cross-ratio

Last updated
Points A, B, C, D and A', B', C', D' are related by a projective transformation so their cross ratios, (A, B; C, D) and (A', B'; C', D') are equal. Projection geometry.svg
Points A, B, C, D and A, B, C, D are related by a projective transformation so their cross ratios, (A, B; C, D) and (A, B; C, D) are equal.

In geometry, the cross-ratio, also called the double ratio and anharmonic ratio, is a number associated with a list of four collinear points, particularly points on a projective line. Given four points A, B, C, D on a line, their cross ratio is defined as

Contents

where an orientation of the line determines the sign of each distance and the distance is measured as projected into Euclidean space. (If one of the four points is the line's point at infinity, then the two distances involving that point are dropped from the formula.) The point D is the harmonic conjugate of C with respect to A and B precisely if the cross-ratio of the quadruple is −1, called the harmonic ratio. The cross-ratio can therefore be regarded as measuring the quadruple's deviation from this ratio; hence the name anharmonic ratio.

The cross-ratio is preserved by linear fractional transformations. It is essentially the only projective invariant of a quadruple of collinear points; this underlies its importance for projective geometry.

The cross-ratio had been defined in deep antiquity, possibly already by Euclid, and was considered by Pappus, who noted its key invariance property. It was extensively studied in the 19th century. [1]

Variants of this concept exist for a quadruple of concurrent lines on the projective plane and a quadruple of points on the Riemann sphere. In the Cayley–Klein model of hyperbolic geometry, the distance between points is expressed in terms of a certain cross-ratio.

Terminology and history

D is the harmonic conjugate of C with respect to A and B, so that the cross-ratio (A, B; C, D) equals -1. Pappusharmonic.svg
D is the harmonic conjugate of C with respect to A and B, so that the cross-ratio (A, B; C, D) equals −1.

Pappus of Alexandria made implicit use of concepts equivalent to the cross-ratio in his Collection: Book VII. Early users of Pappus included Isaac Newton, Michel Chasles, and Robert Simson. In 1986 Alexander Jones made a translation of the original by Pappus, then wrote a commentary on how the lemmas of Pappus relate to modern terminology. [2]

Modern use of the cross ratio in projective geometry began with Lazare Carnot in 1803 with his book Géométrie de Position. [3] [ pages needed ] Chasles coined the French term rapport anharmonique [anharmonic ratio] in 1837. [4] German geometers call it das Doppelverhältnis [double ratio].

Carl von Staudt was unsatisfied with past definitions of the cross-ratio relying on algebraic manipulation of Euclidean distances rather than being based purely on synthetic projective geometry concepts. In 1847, von Staudt demonstrated that the algebraic structure is implicit in projective geometry, by creating an algebra based on construction of the projective harmonic conjugate, which he called a throw (German: Wurf): given three points on a line, the harmonic conjugate is a fourth point that makes the cross ratio equal to −1. His algebra of throws provides an approach to numerical propositions, usually taken as axioms, but proven in projective geometry. [5]

The English term "cross-ratio" was introduced in 1878 by William Kingdon Clifford. [6]

Definition

If A, B, C, and D are four points on an oriented affine line, their cross ratio is:

with the notation defined to mean the signed ratio of the displacement from W to X to the displacement from Y to Z. For colinear displacements this is a dimensionless quantity.

If the displacements themselves are taken to be signed real numbers, then the cross ratio between points can be written

If is the projectively extended real line, the cross-ratio of four distinct numbers in is given by

When one of is the point at infinity (), this reduces to e.g.

The same formulas can be applied to four distinct complex numbers or, more generally, to elements of any field, and can also be projectively extended as above to the case when one of them is

Properties

The cross ratio of the four collinear points A, B, C, and D can be written as

where describes the ratio with which the point C divides the line segment AB, and describes the ratio with which the point D divides that same line segment. The cross ratio then appears as a ratio of ratios, describing how the two points C and D are situated with respect to the line segment AB. As long as the points A, B, C, and D are distinct, the cross ratio (A, B; C, D) will be a non-zero real number. We can easily deduce that

Six cross-ratios

Four points can be ordered in 4! = 4 × 3 × 2 × 1 = 24 ways, but there are only six ways for partitioning them into two unordered pairs. Thus, four points can have only six different cross-ratios, which are related as:

See Anharmonic group below.

Projective geometry

Use of cross-ratios in projective geometry to measure real-world dimensions of features depicted in a perspective projection. A, B, C, D and V are points on the image, their separation given in pixels; A', B', C' and D' are in the real world, their separation in metres.
In (1), the width of the side street, W is computed from the known widths of the adjacent shops.
In (2), the width of only one shop is needed because a vanishing point, V is visible. Cross ratio metrology example.svg
Use of cross-ratios in projective geometry to measure real-world dimensions of features depicted in a perspective projection. A, B, C, D and V are points on the image, their separation given in pixels; A', B', C' and D' are in the real world, their separation in metres.
  • In (1), the width of the side street, W is computed from the known widths of the adjacent shops.
  • In (2), the width of only one shop is needed because a vanishing point, V is visible.

The cross-ratio is a projective invariant in the sense that it is preserved by the projective transformations of a projective line.

In particular, if four points lie on a straight line in then their cross-ratio is a well-defined quantity, because any choice of the origin and even of the scale on the line will yield the same value of the cross-ratio.

Furthermore, let be four distinct lines in the plane passing through the same point . Then any line not passing through intersects these lines in four distinct points (if is parallel to then the corresponding intersection point is "at infinity"). It turns out that the cross-ratio of these points (taken in a fixed order) does not depend on the choice of a line , and hence it is an invariant of the 4-tuple of lines

This can be understood as follows: if and are two lines not passing through then the perspective transformation from to with the center is a projective transformation that takes the quadruple of points on into the quadruple of points on .

Therefore, the invariance of the cross-ratio under projective automorphisms of the line implies (in fact, is equivalent to) the independence of the cross-ratio of the four collinear points on the lines from the choice of the line that contains them.

Definition in homogeneous coordinates

If four collinear points are represented in homogeneous coordinates by vectors a, b, c, d such that c = a + b and d = ka + b, then their cross-ratio is k. [7]

Role in non-Euclidean geometry

Arthur Cayley and Felix Klein found an application of the cross-ratio to non-Euclidean geometry. Given a nonsingular conic C in the real projective plane, its stabilizer GC in the projective group G = PGL(3, R) acts transitively on the points in the interior of C. However, there is an invariant for the action of GC on pairs of points. In fact, every such invariant is expressible as a function of the appropriate cross ratio.[ citation needed ]

Hyperbolic geometry

Explicitly, let the conic be the unit circle. For any two points P and Q, inside the unit circle . If the line connecting them intersects the circle in two points, X and Y and the points are, in order, X, P, Q, Y. Then the hyperbolic distance between P and Q in the Cayley–Klein model of the hyperbolic plane can be expressed as

(the factor one half is needed to make the curvature −1). Since the cross-ratio is invariant under projective transformations, it follows that the hyperbolic distance is invariant under the projective transformations that preserve the conic C.

Conversely, the group G acts transitively on the set of pairs of points (p, q) in the unit disk at a fixed hyperbolic distance.

Later, partly through the influence of Henri Poincaré, the cross ratio of four complex numbers on a circle was used for hyperbolic metrics. Being on a circle means the four points are the image of four real points under a Möbius transformation, and hence the cross ratio is a real number. The Poincaré half-plane model and Poincaré disk model are two models of hyperbolic geometry in the complex projective line.

These models are instances of Cayley–Klein metrics.

Anharmonic group and Klein four-group

The cross-ratio may be defined by any of these four expressions:

These differ by the following permutations of the variables (in cycle notation):

We may consider the permutations of the four variables as an action of the symmetric group S4 on functions of the four variables. Since the above four permutations leave the cross ratio unaltered, they form the stabilizer K of the cross-ratio under this action, and this induces an effective action of the quotient group on the orbit of the cross-ratio. The four permutations in K make a realization of the Klein four-group in S4, and the quotient is isomorphic to the symmetric group S3.

Thus, the other permutations of the four variables alter the cross-ratio to give the following six values, which are the orbit of the six-element group :

The stabilizer of {0, 1, [?]} is isomorphic to the rotation group of the trigonal dihedron, the dihedral group D3. It is convenient to visualize this by a Mobius transformation M mapping the real axis to the complex unit circle (the equator of the Riemann sphere), with 0, 1, [?] equally spaced.
Considering {0, 1, [?]} as the vertices of the dihedron, the other fixed points of the 2-cycles are the points {2, -1, 1/2}, which under M are opposite each vertex on the Riemann sphere, at the midpoint of the opposite edge. Each 2-cycles is a half-turn rotation of the Riemann sphere exchanging the hemispheres (the interior and exterior of the circle in the diagram).
The fixed points of the 3-cycles are exp(+-ip/3), corresponding under M to the poles of the sphere: exp(ip/3) is the origin and exp(-ip/3) is the point at infinity. Each 3-cycle is a 1/3 turn rotation about their axis, and they are exchanged by the 2-cycles. Symmetries of the anharmonic group.png

The stabilizer of {0, 1, ∞} is isomorphic to the rotation group of the trigonal dihedron, the dihedral group D3. It is convenient to visualize this by a Möbius transformation M mapping the real axis to the complex unit circle (the equator of the Riemann sphere), with 0, 1, ∞ equally spaced.

Considering {0, 1, ∞} as the vertices of the dihedron, the other fixed points of the 2-cycles are the points {2, −1, 1/2}, which under M are opposite each vertex on the Riemann sphere, at the midpoint of the opposite edge. Each 2-cycles is a half-turn rotation of the Riemann sphere exchanging the hemispheres (the interior and exterior of the circle in the diagram).

The fixed points of the 3-cycles are exp(±/3), corresponding under M to the poles of the sphere: exp(/3) is the origin and exp(−/3) is the point at infinity. Each 3-cycle is a 1/3 turn rotation about their axis, and they are exchanged by the 2-cycles.

As functions of these are examples of Möbius transformations, which under composition of functions form the Mobius group PGL(2, Z). The six transformations form a subgroup known as the anharmonic group, again isomorphic to S3. They are the torsion elements (elliptic transforms) in PGL(2, Z). Namely, , , and are of order 2 with respective fixed points and (namely, the orbit of the harmonic cross-ratio). Meanwhile, the elements and are of order 3 in PGL(2, Z), and each fixes both values of the "most symmetric" cross-ratio (the solutions to , the primitive sixth roots of unity). The order 2 elements exchange these two elements (as they do any pair other than their fixed points), and thus the action of the anharmonic group on gives the quotient map of symmetric groups .

Further, the fixed points of the individual 2-cycles are, respectively, and and this set is also preserved and permuted by the 3-cycles. Geometrically, this can be visualized as the rotation group of the trigonal dihedron, which is isomorphic to the dihedral group of the triangle D3, as illustrated at right. Algebraically, this corresponds to the action of S3 on the 2-cycles (its Sylow 2-subgroups) by conjugation and realizes the isomorphism with the group of inner automorphisms,

The anharmonic group is generated by and Its action on gives an isomorphism with S3. It may also be realised as the six Möbius transformations mentioned, [8] which yields a projective representation of S3 over any field (since it is defined with integer entries), and is always faithful/injective (since no two terms differ only by 1/−1). Over the field with two elements, the projective line only has three points, so this representation is an isomorphism, and is the exceptional isomorphism . In characteristic 3, this stabilizes the point , which corresponds to the orbit of the harmonic cross-ratio being only a single point, since . Over the field with three elements, the projective line has only 4 points and , and thus the representation is exactly the stabilizer of the harmonic cross-ratio, yielding an embedding equals the stabilizer of the point .

Exceptional orbits

For certain values of there will be greater symmetry and therefore fewer than six possible values for the cross-ratio. These values of correspond to fixed points of the action of S3 on the Riemann sphere (given by the above six functions); or, equivalently, those points with a non-trivial stabilizer in this permutation group.

The first set of fixed points is However, the cross-ratio can never take on these values if the points A, B, C, and D are all distinct. These values are limit values as one pair of coordinates approach each other:

The second set of fixed points is This situation is what is classically called the harmonic cross-ratio, and arises in projective harmonic conjugates. In the real case, there are no other exceptional orbits.

In the complex case, the most symmetric cross-ratio occurs when . These are then the only two values of the cross-ratio, and these are acted on according to the sign of the permutation.

Transformational approach

The cross-ratio is invariant under the projective transformations of the line. In the case of a complex projective line, or the Riemann sphere, these transformations are known as Möbius transformations. A general Möbius transformation has the form

These transformations form a group acting on the Riemann sphere, the Möbius group.

The projective invariance of the cross-ratio means that

The cross-ratio is real if and only if the four points are either collinear or concyclic, reflecting the fact that every Möbius transformation maps generalized circles to generalized circles.

The action of the Möbius group is simply transitive on the set of triples of distinct points of the Riemann sphere: given any ordered triple of distinct points, (z2, z3, z4), there is a unique Möbius transformation f(z) that maps it to the triple (1, 0, ∞). This transformation can be conveniently described using the cross-ratio: since (z, z2; z3, z4) must equal (f(z), 1; 0, ∞), which in turn equals f(z), we obtain

An alternative explanation for the invariance of the cross-ratio is based on the fact that the group of projective transformations of a line is generated by the translations, the homotheties, and the multiplicative inversion. The differences zjzk are invariant under the translations

where a is a constant in the ground field F. Furthermore, the division ratios are invariant under a homothety

for a non-zero constant b in F. Therefore, the cross-ratio is invariant under the affine transformations.

In order to obtain a well-defined inversion mapping

the affine line needs to be augmented by the point at infinity, denoted , forming the projective line P1(F). Each affine mapping f : FF can be uniquely extended to a mapping of P1(F) into itself that fixes the point at infinity. The map T swaps 0 and . The projective group is generated by T and the affine mappings extended to P1(F). In the case F = C, the complex plane, this results in the Möbius group. Since the cross-ratio is also invariant under T, it is invariant under any projective mapping of P1(F) into itself.

Co-ordinate description

If we write the complex points as vectors and define , and let be the dot product of with , then the real part of the cross ratio is given by:

This is an invariant of the 2-dimensional special conformal transformation such as inversion .

The imaginary part must make use of the 2-dimensional cross product

Ring homography

The concept of cross ratio only depends on the ring operations of addition, multiplication, and inversion (though inversion of a given element is not certain in a ring). One approach to cross ratio interprets it as a homography that takes three designated points to 0, 1, and . Under restrictions having to do with inverses, it is possible to generate such a mapping with ring operations in the projective line over a ring. The cross ratio of four points is the evaluation of this homography at the fourth point.

Differential-geometric point of view

The theory takes on a differential calculus aspect as the four points are brought into proximity. This leads to the theory of the Schwarzian derivative, and more generally of projective connections.

Higher-dimensional generalizations

The cross-ratio does not generalize in a simple manner to higher dimensions, due to other geometric properties of configurations of points, notably collinearity – configuration spaces are more complicated, and distinct k-tuples of points are not in general position.

While the projective linear group of the projective line is 3-transitive (any three distinct points can be mapped to any other three points), and indeed simply 3-transitive (there is a unique projective map taking any triple to another triple), with the cross ratio thus being the unique projective invariant of a set of four points, there are basic geometric invariants in higher dimension. The projective linear group of n-space has (n + 1)2 − 1 dimensions (because it is projectivization removing one dimension), but in other dimensions the projective linear group is only 2-transitive – because three collinear points must be mapped to three collinear points (which is not a restriction in the projective line) – and thus there is not a "generalized cross ratio" providing the unique invariant of n2 points.

Collinearity is not the only geometric property of configurations of points that must be maintained – for example, five points determine a conic, but six general points do not lie on a conic, so whether any 6-tuple of points lies on a conic is also a projective invariant. One can study orbits of points in general position – in the line "general position" is equivalent to being distinct, while in higher dimensions it requires geometric considerations, as discussed – but, as the above indicates, this is more complicated and less informative.

However, a generalization to Riemann surfaces of positive genus exists, using the Abel–Jacobi map and theta functions.

See also

Notes

  1. A theorem on the anharmonic ratio of lines appeared in the work of Pappus, but Michel Chasles, who devoted considerable efforts to reconstructing lost works of Euclid, asserted that it had earlier appeared in his book Porisms.
  2. Alexander Jones (1986) Book 7 of the Collection, part 1: introduction, text, translation ISBN   0-387-96257-3, part 2: commentary, index, figures ISBN   3-540-96257-3, Springer-Verlag
  3. Carnot, Lazare (1803). Géométrie de Position. Crapelet.
  4. Chasles, Michel (1837). Aperçu historique sur l'origine et le développement des méthodes en géométrie. Hayez. p. 35. (Link is to the reprinted second edition, Gauthier-Villars: 1875.)
  5. Howard Eves (1972) A Survey of Geometry, Revised Edition, page 73, Allyn and Bacon
  6. W.K. Clifford (1878) Elements of Dynamic, books I,II,III, page 42, London: MacMillan & Co; on-line presentation by Cornell University Historical Mathematical Monographs.
  7. Irving Kaplansky (1969). Linear Algebra and Geometry: A Second Course . ISBN   0-486-43233-5.
  8. Chandrasekharan, K. (1985). Elliptic Functions. Grundlehren der mathematischen Wissenschaften. Vol. 281. Springer-Verlag. p. 120. ISBN   3-540-15295-4. Zbl   0575.33001.

Related Research Articles

<span class="mw-page-title-main">Lorentz transformation</span> Family of linear transformations

In physics, the Lorentz transformations are a six-parameter family of linear transformations from a coordinate frame in spacetime to another frame that moves at a constant velocity relative to the former. The respective inverse transformation is then parameterized by the negative of this velocity. The transformations are named after the Dutch physicist Hendrik Lorentz.

<span class="mw-page-title-main">Affine transformation</span> Geometric transformation that preserves lines but not angles nor the origin

In Euclidean geometry, an affine transformation or affinity is a geometric transformation that preserves lines and parallelism, but not necessarily Euclidean distances and angles.

<span class="mw-page-title-main">Optical depth</span> Physics concept

In physics, optical depth or optical thickness is the natural logarithm of the ratio of incident to transmitted radiant power through a material. Thus, the larger the optical depth, the smaller the amount of transmitted radiant power through the material. Spectral optical depth or spectral optical thickness is the natural logarithm of the ratio of incident to transmitted spectral radiant power through a material. Optical depth is dimensionless, and in particular is not a length, though it is a monotonically increasing function of optical path length, and approaches zero as the path length approaches zero. The use of the term "optical density" for optical depth is discouraged.

<span class="mw-page-title-main">Ellipsoid</span> Quadric surface that looks like a deformed sphere

An ellipsoid is a surface that can be obtained from a sphere by deforming it by means of directional scalings, or more generally, of an affine transformation.

<span class="mw-page-title-main">Poisson's ratio</span> Measure of material deformation perpendicular to loading

In materials science and solid mechanics, Poisson's ratio (nu) is a measure of the Poisson effect, the deformation of a material in directions perpendicular to the specific direction of loading. The value of Poisson's ratio is the negative of the ratio of transverse strain to axial strain. For small values of these changes, is the amount of transversal elongation divided by the amount of axial compression. Most materials have Poisson's ratio values ranging between 0.0 and 0.5. For soft materials, such as rubber, where the bulk modulus is much higher than the shear modulus, Poisson's ratio is near 0.5. For open-cell polymer foams, Poisson's ratio is near zero, since the cells tend to collapse in compression. Many typical solids have Poisson's ratios in the range of 0.2–0.3. The ratio is named after the French mathematician and physicist Siméon Poisson.

In geometry, inversive geometry is the study of inversion, a transformation of the Euclidean plane that maps circles or lines to other circles or lines and that preserves the angles between crossing curves. Many difficult problems in geometry become much more tractable when an inversion is applied. Inversion seems to have been discovered by a number of people contemporaneously, including Steiner (1824), Quetelet (1825), Bellavitis (1836), Stubbs and Ingram (1842-3) and Kelvin (1845).

In geometry and complex analysis, a Möbius transformation of the complex plane is a rational function of the form

<span class="mw-page-title-main">Modular group</span> Orientation-preserving mapping class group of the torus

In mathematics, the modular group is the projective special linear group of 2 × 2 matrices with integer coefficients and determinant 1. The matrices A and A are identified. The modular group acts on the upper-half of the complex plane by fractional linear transformations, and the name "modular group" comes from the relation to moduli spaces and not from modular arithmetic.

<span class="mw-page-title-main">Projective linear group</span> Construction in group theory

In mathematics, especially in the group theoretic area of algebra, the projective linear group (also known as the projective general linear group or PGL) is the induced action of the general linear group of a vector space V on the associated projective space P(V). Explicitly, the projective linear group is the quotient group

In mathematics, the Schwarzian derivative is an operator similar to the derivative which is invariant under Möbius transformations. Thus, it occurs in the theory of the complex projective line, and in particular, in the theory of modular forms and hypergeometric functions. It plays an important role in the theory of univalent functions, conformal mapping and Teichmüller spaces. It is named after the German mathematician Hermann Schwarz.

<span class="mw-page-title-main">Linear time-invariant system</span> Mathematical model which is both linear and time-invariant

In system analysis, among other fields of study, a linear time-invariant (LTI) system is a system that produces an output signal from any input signal subject to the constraints of linearity and time-invariance; these terms are briefly defined below. These properties apply (exactly or approximately) to many important physical systems, in which case the response y(t) of the system to an arbitrary input x(t) can be found directly using convolution: y(t) = (xh)(t) where h(t) is called the system's impulse response and ∗ represents convolution (not to be confused with multiplication). What's more, there are systematic methods for solving any such system (determining h(t)), whereas systems not meeting both properties are generally more difficult (or impossible) to solve analytically. A good example of an LTI system is any electrical circuit consisting of resistors, capacitors, inductors and linear amplifiers.

<span class="mw-page-title-main">Fundamental pair of periods</span> Way of defining a lattice in the complex plane

In mathematics, a fundamental pair of periods is an ordered pair of complex numbers that defines a lattice in the complex plane. This type of lattice is the underlying object with which elliptic functions and modular forms are defined.

<span class="mw-page-title-main">Lemniscate elliptic functions</span> Mathematical functions

In mathematics, the lemniscate elliptic functions are elliptic functions related to the arc length of the lemniscate of Bernoulli. They were first studied by Giulio Fagnano in 1718 and later by Leonhard Euler and Carl Friedrich Gauss, among others.

In 3-dimensional topology, a part of the mathematical field of geometric topology, the Casson invariant is an integer-valued invariant of oriented integral homology 3-spheres, introduced by Andrew Casson.

A ratio distribution is a probability distribution constructed as the distribution of the ratio of random variables having two other known distributions. Given two random variables X and Y, the distribution of the random variable Z that is formed as the ratio Z = X/Y is a ratio distribution.

SL<sub>2</sub>(<b>R</b>) Group of real 2×2 matrices with unit determinant

In mathematics, the special linear group SL(2, R) or SL2(R) is the group of 2 × 2 real matrices with determinant one:

<span class="mw-page-title-main">Projective harmonic conjugate</span> Point found separated from another, given a point pair

In projective geometry, the harmonic conjugate point of a point on the real projective line with respect to two other points is defined by the following construction:

<span class="mw-page-title-main">Poisson distribution</span> Discrete probability distribution

In probability theory and statistics, the Poisson distribution is a discrete probability distribution that expresses the probability of a given number of events occurring in a fixed interval of time if these events occur with a known constant mean rate and independently of the time since the last event. It can also be used for the number of events in other types of intervals than time, and in dimension greater than 1.

<span class="mw-page-title-main">Riemann sphere</span> Model of the extended complex plane plus a point at infinity

In mathematics, the Riemann sphere, named after Bernhard Riemann, is a model of the extended complex plane: the complex plane plus one point at infinity. This extended plane represents the extended complex numbers, that is, the complex numbers plus a value for infinity. With the Riemann model, the point is near to very large numbers, just as the point is near to very small numbers.

<span class="mw-page-title-main">Poisson point process</span> Type of random mathematical object

In probability, statistics and related fields, a Poisson point process is a type of random mathematical object that consists of points randomly located on a mathematical space with the essential feature that the points occur independently of one another. The Poisson point process is also called a Poisson random measure, Poisson random point field or Poisson point field. When the process is defined on the real line, it is often called simply the Poisson process.

References