Quadric

From Wikipedia, the free encyclopedia

In mathematics, a quadric or quadric surface (quadric hypersurface in higher dimensions), is a generalization of conic sections (ellipses, parabolas, and hyperbolas). It is a hypersurface (of dimension D) in a (D + 1)-dimensional space, and it is defined as the zero set of an irreducible polynomial of degree two in D + 1 variables (D = 1 in the case of conic sections). When the defining polynomial is not absolutely irreducible, the zero set is generally not considered a quadric, although it is often called a degenerate quadric or a reducible quadric.

In coordinates x1, x2, ..., xD+1, the general quadric is thus defined by the algebraic equation[1]

which may be compactly written in vector and matrix notation as:

where x = (x1, x2, ..., xD+1) is a row vector, xT is the transpose of x (a column vector), Q is a (D + 1) × (D + 1) matrix and P is a (D + 1)-dimensional row vector and R a scalar constant. The values Q, P and R are often taken to be over real numbers or complex numbers, but a quadric may be defined over any field.

A quadric is an affine algebraic variety, or, if it is reducible, an affine algebraic set. Quadrics may also be defined in projective spaces; see § Projective geometry, below.

Euclidean plane[]

As the dimension of a Euclidean plane is two, quadrics in a Euclidean plane have dimension one and are thus plane curves. They are called conic sections, or conics.

Circle (e = 0), ellipse (e = 0.5), parabola (e = 1), and hyperbola (e = 2) with fixed focus F and directrix.

Euclidean space[]

In three-dimensional Euclidean space, quadrics have dimension two, and are known as quadric surfaces. Their quadratic equations have the form

where are real numbers, and at least one of A, B, and C is nonzero.

The quadric surfaces are classified and named by their shape, which corresponds to the orbits under affine transformations. That is, if an affine transformation maps a quadric onto another one, they belong to the same class, and share the same name and many properties.

The principal axis theorem shows that for any (possibly reducible) quadric, a suitable change of Cartesian coordinates or, equivalently, a Euclidean transformation allows putting the equation of the quadric into a unique simple form on which the class of the quadric is immediately visible. This form is called the normal form of the equation, since two quadrics have the same normal form if and only if there is a Euclidean transformation that maps one quadric to the other. The normal forms are as follows:

where the are either 1, –1 or 0, except which takes only the value 0 or 1.

Each of these 17 normal forms[2] corresponds to a single orbit under affine transformations. In three cases there are no real points: (imaginary ellipsoid), (imaginary elliptic cylinder), and (pair of complex conjugate parallel planes, a reducible quadric). In one case, the imaginary cone, there is a single point (). If one has a line (in fact two complex conjugate intersecting planes). For one has two intersecting planes (reducible quadric). For one has a double plane. For one has two parallel planes (reducible quadric).

Thus, among the 17 normal forms, there are nine true quadrics: a cone, three cylinders (often called degenerate quadrics) and five non-degenerate quadrics (ellipsoid, paraboloids and hyperboloids), which are detailed in the following tables. The eight remaining quadrics are the imaginary ellipsoid (no real point), the imaginary cylinder (no real point), the imaginary cone (a single real point), and the reducible quadrics, which are decomposed in two planes; there are five such decomposed quadrics, depending whether the planes are distinct or not, parallel or not, real or complex conjugate.

Non-degenerate real quadric surfaces
    Ellipsoid Ellipsoid Quadric.png
    Elliptic paraboloid Paraboloid Quadric.Png
    Hyperbolic paraboloid Hyperbolic Paraboloid Quadric.png
   Hyperboloid of one sheet
      or
   Hyperbolic hyperboloid
Hyperboloid Of One Sheet Quadric.png
   Hyperboloid of two sheets
      or
   Elliptic hyperboloid
Hyperboloid Of Two Sheets Quadric.png
Degenerate real quadric surfaces
    Elliptic cone
      or
   Conical quadric
Elliptical Cone Quadric.Png
    Elliptic cylinder Elliptic Cylinder Quadric.png
    Hyperbolic cylinder Hyperbolic Cylinder Quadric.png
    Parabolic cylinder Parabolic Cylinder Quadric.png

When two or more of the parameters of the canonical equation are equal, one gets a quadric of revolution, which remains invariant when rotated around an axis (or infinitely many axes, in the case of the sphere).

Quadrics of revolution
    Oblate and prolate spheroids (special cases of ellipsoid) Oblate Spheroid Quadric.pngProlate Spheroid Quadric.png
    Sphere (special case of spheroid) Sphere Quadric.png
    Circular paraboloid (special case of elliptic paraboloid) Circular Paraboloid Quadric.png
    Hyperboloid of revolution of one sheet (special case of hyperboloid of one sheet) Circular Hyperboloid Of One Sheet Quadric.png
    Hyperboloid of revolution of two sheets (special case of hyperboloid of two sheets) Circular Hyperboloid of Two Sheets Quadric.png
    Circular cone (special case of elliptic cone) Circular Cone Quadric.png
    Circular cylinder (special case of elliptic cylinder) Circular Cylinder Quadric.png

Definition and basic properties[]

An affine quadric is the set of zeros of a polynomial of degree two. When not specified otherwise, the polynomial is supposed to have real coefficients, and the zeros are points in a Euclidean space. However, most properties remain true when the coefficients belong to any field and the points belong in an affine space. As usually in algebraic geometry, it is often useful to consider points over an algebraically closed field containing the polynomial coefficients, generally the complex numbers, when the coefficients are real.

Many properties becomes easier to state (and to prove) by extending the quadric to the projective space by projective completion, consisting of adding points at infinity. Technically, if

is a polynomial of degree two that defines an affine quadric, then its projective completion is defined by homogenizing p into

(this is a polynomial, because the degree of p is two). The points of the projective completion are the points of the projective space whose projective coordinates are zeros of P.

So, a projective quadric is the set of zeros in a projective space of a homogeneous polynomial of degree two.

As the above process of homogenization can be reverted by setting X0 = 1:

it is often useful to not distinguish an affine quadric from its projective completion, and to talk of the affine equation or the projective equation of a quadric. However, this is not a perfect equivalence; it is generally the case that will include points with , which are not also solutions of because these points in projective space correspond to points "at infinity" in affine space.

Equation[]

A quadric in an affine space of dimension n is the set of zeros of a polynomial of degree 2. That is, it is the set of the points whose coordinates satisfy an equation

where the polynomial p has the form

for a matrix with and running from 0 to . When the characteristic of the field of the coefficients is not two, generally is assumed; equivalently . When the characteristic of the field of the coefficients is two, generally is assumed when ; equivalently is upper triangular.

The equation may be shortened, as the matrix equation

with

The equation of the projective completion is almost identical:

with

These equations define a quadric as an algebraic hypersurface of dimension n – 1 and degree two in a space of dimension n.

The quadric is said to be non-degenerate if the matrix is invertible.

Normal form of projective quadrics[]

In real projective space, by Sylvester's law of inertia, a non-singular quadratic form P(X) may be put into the normal form

by means of a suitable projective transformation (normal forms for singular quadrics can have zeros as well as ±1 as coefficients). For two-dimensional surfaces (dimension D = 2) in three-dimensional space, there are exactly three non-degenerate cases:

The first case is the empty set.

The second case generates the ellipsoid, the elliptic paraboloid or the hyperboloid of two sheets, depending on whether the chosen plane at infinity cuts the quadric in the empty set, in a point, or in a nondegenerate conic respectively. These all have positive Gaussian curvature.

The third case generates the hyperbolic paraboloid or the hyperboloid of one sheet, depending on whether the plane at infinity cuts it in two lines, or in a nondegenerate conic respectively. These are doubly ruled surfaces of negative Gaussian curvature.

The degenerate form

generates the elliptic cylinder, the parabolic cylinder, the hyperbolic cylinder, or the cone, depending on whether the plane at infinity cuts it in a point, a line, two lines, or a nondegenerate conic respectively. These are singly ruled surfaces of zero Gaussian curvature.

We see that projective transformations don't mix Gaussian curvatures of different sign. This is true for general surfaces.[3]

In complex projective space all of the nondegenerate quadrics become indistinguishable from each other.

Integer and rational solutions[]

Each solution of with a vector having rational components yields a vector with integer components that satisfies ; set where the multiplying factor is the smallest positive integer that clears all the denominators of the components of .

Furthermore, when the underlying matrix is invertible, any one solution to for with rational components can be used to find any other solution with rational components, as follows. Let for some values of and , both with integer components, and value . Writing for a non-singular symmetric matrix with integer components, we have that

When

then the two solutions to , when viewed as a quadratic equation in , will be , where the latter is non-zero and rational. In particular, if is a solution of and is the corresponding non-zero solution of then any for which (1) is not orthogonal to and (2) satisfies these three conditions and gives a non-zero rational value for .

In short, if one knows one solution with rational components then one can find many integer solutions where depends upon the choice of . Furthermore, the process is reversible! If both satisfies and satisfies then the choice of will necessarily produce . With this approach one can generate all Pythagorean triples or Heronian triangles.

Projective quadrics over fields[]

The definition of a projective quadric in a real projective space (see above) can be formally adopted defining a projective quadric in an n-dimensional projective space over a field. In order to omit dealing with coordinates a projective quadric is usually defined starting with a quadratic form on a vector space [4]

Quadratic form[]

Let be a field and a vector space over . A mapping from to such that

(Q1) for any and .
(Q2) is a bilinear form.

is called quadratic form. The bilinear form is symmetric.

In case of the bilinear form is , i.e. and are mutually determined in a unique way.
In case of (that means: ) the bilinear form has the property , i.e. is symplectic.

For and ( is a base of ) has the familiar form

and
.

For example:

n-dimensional projective space over a field[]

Let be a field, ,

an (n + 1)-dimensional vector space over the field
the 1-dimensional subspace generated by ,
the set of points ,
the set of lines.
is the n-dimensional projective space over .
The set of points contained in a -dimensional subspace of is a -dimensional subspace of . A 2-dimensional subspace is a plane.
In case of a -dimensional subspace is called hyperplane.

Projective quadric[]

For a quadratic form on a vector space a point is called singular if . The set

of singular points of is called quadric (with respect to the quadratic form ).

Examples in .:
(E1): For one gets a conic.
(E2): For one gets the pair of lines with the equations and , respectively. They intersect at point ;

For the considerations below it is assumed that .

Polar space[]

For point the set

is called polar space of (with respect to ).

If for any , one gets .

If for at least one , the equation is a non trivial linear equation which defines a hyperplane. Hence

is either a hyperplane or .

Intersection with a line[]

For the intersection of a line with a quadric the familiar statement is true:

For an arbitrary line the following cases occur:
a) and is called exterior line or
b) and is called tangent line or
b′) and is called tangent line or
c) and is called secant line.

Proof: Let be a line, which intersects at point and is a second point on . From one gets

I) In case of the equation holds and it is for any . Hence either for any or for any , which proves b) and b').
II) In case of one gets and the equation has exactly one solution . Hence: , which proves c).

Additionally the proof shows:

A line through a point is a tangent line if and only if .

f-radical, q-radical[]

In the classical cases or there exists only one radical, because of and and are closely connected. In case of the quadric is not determined by (see above) and so one has to deal with two radicals:

a) is a projective subspace. is called f-radical of quadric .
b) is called singular radical or -radical of .
c) In case of one has .

A quadric is called non-degenerate if .

Examples in (see above):
(E1): For (conic) the bilinear form is
In case of the polar spaces are never . Hence .
In case of the bilinear form is reduced to and . Hence In this case the f-radical is the common point of all tangents, the so called knot.
In both cases and the quadric (conic) ist non-degenerate.
(E2): For (pair of lines) the bilinear form is and the intersection point.
In this example the quadric is degenerate.

Symmetries[]

A quadric is a rather homogeneous object:

For any point there exists an involutorial central collineation with center and .

Proof: Due to the polar space is a hyperplane.

The linear mapping

induces an involutorial central collineation with axis and centre which leaves invariant.
In case of mapping gets the familiar shape with and for any .

Remark:

a) An exterior line, a tangent line or a secant line is mapped by the involution on an exterior, tangent and secant line, respectively.
b) is pointwise fixed by .

q-subspaces and index of a quadric[]

A subspace of is called -subspace if

For example: points on a sphere or lines on a hyperboloid (s. below).

Any two maximal -subspaces have the same dimension .[5]

Let be the dimension of the maximal -subspaces of then

The integer is called index of .

Theorem: (BUEKENHOUT)[6]

For the index of a non-degenerate quadric in the following is true:
.

Let be a non-degenerate quadric in , and its index.

In case of quadric is called sphere (or oval conic if ).
In case of quadric is called hyperboloid (of one sheet).

Examples:

a) Quadric in with form is non-degenerate with index 1.
b) If polynomial is irreducible over the quadratic form gives rise to a non-degenerate quadric in of index 1 (sphere). For example: is irreducible over (but not over  !).
c) In the quadratic form generates a hyperboloid.

Generalization of quadrics: quadratic sets[]

It is not reasonable to formally extend the definition of quadrics to spaces over genuine skew fields (division rings). Because one would get secants bearing more than 2 points of the quadric which is totally different from usual quadrics.[7][8][9] The reason is the following statement.

A division ring is commutative if and only if any equation , has at most two solutions.

There are generalizations of quadrics: quadratic sets.[10] A quadratic set is a set of points of a projective space with the same geometric properties as a quadric: every line intersects a quadratic set in at most two points or is contained in the set.

See also[]

References[]

  1. ^ Silvio Levy Quadrics in "Geometry Formulas and Facts", excerpted from 30th Edition of CRC Standard Mathematical Tables and Formulas, CRC Press, from The Geometry Center at University of Minnesota
  2. ^ Stewart Venit and Wayne Bishop, Elementary Linear Algebra (fourth edition), International Thompson Publishing, 1996.
  3. ^ S. Lazebnik and J. Ponce, "The Local Projective Shape of Smooth Surfaces and Their Outlines" (PDF)., Proposition 1
  4. ^ Beutelspacher/Rosenbaum: p. 158
  5. ^ Beutelpacher/Rosenbaum, p.139
  6. ^ F. Buekenhout: Ensembles Quadratiques des Espace Projective, Math. Teitschr. 110 (1969), p. 306-318.
  7. ^ R. Artzy: The Conic in Moufang Planes, Aequat.Mathem. 6 (1971), p. 31-35
  8. ^ E. Berz: Kegelschnitte in Desarguesschen Ebenen, Math. Zeitschr. 78 (1962), p. 55-8
  9. ^ external link E. Hartmann: Planar Circle Geometries, p. 123
  10. ^ Beutelspacher/Rosenbaum: p. 135

Bibliography[]

  • M. Audin: Geometry, Springer, Berlin, 2002, ISBN 978-3-540-43498-6, p. 200.
  • M. Berger: Problem Books in Mathematics, ISSN 0941-3502, Springer New York, pp 79–84.
  • A. Beutelspacher, U. Rosenbaum: Projektive Geometrie, Vieweg + Teubner, Braunschweig u. a. 1992, ISBN 3-528-07241-5, p. 159.
  • P. Dembowski: Finite Geometries, Springer, 1968, ISBN 978-3-540-61786-0, p. 43.
  • Iskovskikh, V.A. (2001) [1994], "Quadric", Encyclopedia of Mathematics, EMS Press
  • Weisstein, Eric W. "Quadric". MathWorld.

External links[]

Retrieved from ""