Fact-checked by Grok 2 weeks ago

Quadric

In , a quadric is a defined by a . In , it is a second-order given by the general ax^2 + by^2 + cz^2 + dxy + exz + fyz + gx + hy + iz + j = 0, where the coefficients determine the specific shape. These surfaces generalize conic sections—such as ellipses, parabolas, and hyperbolas—from two dimensions to three, encompassing a variety of geometric forms including ellipsoids, hyperboloids, paraboloids, cones, and cylinders. Quadrics are fundamental in and , appearing in applications ranging from and physics to optimization problems, due to their smooth, curved structures that model phenomena like planetary orbits or light reflection. The of quadrics relies on reducing the general to one of 17 standard forms through coordinate transformations, which reveal their symmetries and properties such as ellipticity, hyperbolicity, or degeneracy. Non-degenerate quadrics, like the \frac{x^2}{a^2} + \frac{y^2}{b^2} + \frac{z^2}{c^2} = 1 or the of one sheet \frac{x^2}{a^2} + \frac{y^2}{b^2} - \frac{z^2}{c^2} = 1, are bounded or unbounded depending on the signs of the quadratic terms, while degenerate cases include pairs of planes or single planes. In , quadrics extend to higher dimensions as hypersurfaces defined by homogeneous quadratic polynomials, unifying their study across , affine, and projective spaces. Quadric surfaces exhibit traces—cross-sections parallel to the coordinate planes—that are conic sections, aiding in their and ; for instance, an elliptic z = x^2 + y^2 yields ellipses in horizontal planes and parabolas in vertical planes. Their mathematical significance stems from forms, where the associated symmetric matrix's eigenvalues dictate the surface's type: positive definite for ellipsoids, indefinite for hyperboloids. The mathematical study of quadrics developed in the with the introduction of , evolving into tools for modern fields such as and .

Fundamentals of Quadrics

Definition and Equation

A is defined as the set of points in \mathbb{P}^n where a homogeneous in n+1 variables equals zero. This , of dimension n-1 and degree two, arises as the zero locus X = V(Q) of a single such Q. In affine space \mathbb{A}^n, the general equation of a quadric takes the form \sum_{i=1}^n \sum_{j=1}^n a_{ij} x_i x_j + \sum_{i=1}^n b_i x_i + c = 0, where the quadratic terms are present (i.e., not all a_{ij} = 0), incorporating cross terms like $2a_{12}xy in lower dimensions. To embed this in projective space \mathbb{P}^n, homogenize the equation using an additional variable x_0, yielding the form \sum_{0 \leq i \leq j \leq n} a_{ij} x_i x_j = 0, a homogeneous quadratic polynomial of degree two. This equation corresponds to a quadratic form Q(\mathbf{x}) = \mathbf{x}^T A \mathbf{x}, where \mathbf{x} = (x_0, \dots, x_n)^T is the and A = (a_{ij}) is a symmetric (n+1) \times (n+1) , ensuring the form is well-defined under in projective coordinates. Quadrics are degenerate when the matrix A is singular (i.e., \det(A) = 0), in which case the factors into lower-degree components, such as lines or planes in .

Basic Properties in Euclidean Space

In Euclidean space \mathbb{R}^n, a quadric hypersurface is defined as the level set \{x \in \mathbb{R}^n \mid Q(x) = c\}, where Q(x) = x^T A x + b^T x + d is a with A, linear term b, and constant d. The geometric properties depend on the eigenvalues of A and the value of c. For positive definite A (all eigenvalues positive), the level set for appropriate c > 0 forms an , which is compact and bounded. In contrast, indefinite A (mixed positive and negative eigenvalues) yields hyperboloids, which are unbounded. Elliptic quadrics, such as ellipsoids arising from positive definite s, bound regions, meaning the joining any two points on the surface lies within the bounded region they enclose. Hyperbolic quadrics, like hyperboloids from indefinite forms, are non-, as they feature saddle-like regions where s between points may exit the surface. These convexity distinctions follow from the of the quadratic form, with positive definite cases ensuring the sublevel sets \{x \mid Q(x) \leq c\} are bodies. Non-degenerate quadrics, where A has full rank (determinant nonzero), form smooth hypersurfaces away from singular points, as the gradient \nabla Q(x) = 2Ax + b vanishes only at isolated vertices or along degenerate directions, making the level set a smooth (n-1)-manifold. For surfaces in \mathbb{R}^3, the Gaussian curvature K, defined as the product of principal curvatures, characterizes local geometry: ellipsoids have positive K > 0 everywhere, indicating elliptic points; one-sheet hyperboloids have negative K < 0, marking hyperbolic points; and two-sheet hyperboloids have positive K > 0 on each sheet. Under orthogonal transformations O \in O(n), which preserve the Euclidean inner product \langle x, y \rangle, quadrics transform to quadrics of the same type, as Q(Ox) = x^T (O^T A O) x + (O^T b)^T x + d yields a congruent symmetric matrix O^T A O with unchanged eigenvalues and signature. This invariance ensures that properties like compactness, convexity, and signs are preserved, facilitating forms via in an .

Quadrics in Low Dimensions

Quadrics in the Euclidean Plane

In the , quadrics are represented by conic sections, which are the curves obtained by intersecting a plane with a right circular . The general equation of a conic section is ax^2 + bxy + cy^2 + dx + ey + f = 0, where a, b, c, d, e, f are real coefficients, and not all of a, b, c are zero. The classification of non-degenerate conics depends on the \Delta = b^2 - 4ac. If \Delta < 0, the conic is an (including a circle as a special case when a = c and b = 0); if \Delta = 0, it is a parabola; and if \Delta > 0, it is a . Standard equations for these conics, assuming centered or vertex-aligned forms without rotation, are as follows. For an centered at the with semi-major axis a and semi-minor axis b (where a > b > 0): \frac{x^2}{a^2} + \frac{y^2}{b^2} = 1 For a parabola with vertex at the origin opening to the right, with focal length p > 0: y^2 = 4px For a hyperbola centered at the origin with transverse axis along x of semi-length a > 0 and conjugate axis semi-length b > 0: \frac{x^2}{a^2} - \frac{y^2}{b^2} = 1 Degenerate conics arise when the factors into linear terms or represents trivial loci, resulting in four types: a single point (e.g., x^2 + y^2 = 0), two intersecting lines (e.g., xy = 0), (e.g., x^2 - 1 = 0), or the (e.g., x^2 + y^2 + 1 = 0). Each non-degenerate conic is characterized by geometric properties involving , directrix, and e, defined as the ratio of the distance from a point on the conic to a over the distance to the corresponding directrix. For an , there are two at (\pm c, 0) where c = \sqrt{a^2 - b^2}, directrices at x = \pm \frac{a}{e}, and $0 \leq e < 1. For a parabola, the single is at (p, 0), the directrix is x = -p, and e = 1. For a , the foci are at (\pm c, 0) where c = \sqrt{a^2 + b^2}, directrices at x = \pm \frac{a}{e}, and e > 1. A line intersects a conic section in at most two points, as substituting the line into the conic yields a in one variable with up to two real roots.

Quadrics in Three-Space

In three-space, quadric surfaces are the loci of points satisfying a second-degree in three variables, generalizing conic sections to three dimensions. These surfaces encompass a variety of shapes, including bounded and unbounded forms, and play a fundamental role in due to their and intersection properties. Unlike their planar counterparts, quadric surfaces exhibit volumetric characteristics and can extend infinitely in certain directions./12%3A_Vectors_in_Space/12.06%3A_Quadric_Surfaces) The classification of non-degenerate quadric surfaces in Euclidean three-space includes ellipsoids, hyperboloids of one or two sheets, elliptic paraboloids, and hyperbolic paraboloids. Degenerate cases comprise cones and cylinders. Ellipsoids are closed, bounded surfaces resembling stretched spheres, while hyperboloids of one sheet form connected, saddle-like structures that extend infinitely. Hyperboloids of two sheets consist of two separate unbounded components. Elliptic paraboloids are bowl-shaped and open in one direction, whereas hyperbolic paraboloids resemble saddles with rulings in two directions. Cones taper to a vertex and extend infinitely, and cylinders are extruded conics along a line, lacking a vertex. Standard equations illustrate these forms; for an aligned with the coordinate , the equation is \frac{x^2}{a^2} + \frac{y^2}{b^2} + \frac{z^2}{c^2} = 1, where a, b, c > 0 determine the semi-axes lengths. For a of one sheet, it is \frac{x^2}{a^2} + \frac{y^2}{b^2} - \frac{z^2}{c^2} = 1, with a, b, c > 0, yielding a surface that narrows at the waist and flares outward. These equations assume principal , but general quadrics involve cross terms resolved via orthogonal transformations./12%3A_Vectors_in_Space/12.06%3A_Quadric_Surfaces) Key properties distinguish these surfaces. Ellipsoids are compact and , enclosing a finite volume. Hyperboloids of and two sheets, along with , exhibit asymptotic behavior: hyperboloids approach a at infinity, known as their asymptotic cone, beyond which the surface does not extend. Hyperboloids of and are ruled surfaces, containing families of straight lines (rulings) lying entirely on ; the former is doubly ruled with two such families. paraboloids are also doubly ruled, facilitating applications in like minimal surfaces. Cylinders inherit the properties of their generating conic, remaining parallel and unbounded. Visualization of quadric surfaces often relies on their cross-sections, or traces, in coordinate planes, which yield conic sections such as ellipses, hyperbolas, parabolas, or degenerate forms. For instance, horizontal slices of an produce ellipses, while vertical slices of a of one sheet yield hyperbolas. These planar intersections provide a direct link to two-dimensional conics, aiding in identification and sketching./12%3A_Vectors_in_Space/12.06%3A_Quadric_Surfaces) Over the real numbers, some quadrics may lack real points or consist partly of imaginary components, such as the imaginary ellipsoid x^2 + y^2 + z^2 = -1, which has no real points but serves as a precursor to where complex extensions unify real and imaginary cases.

Projective Quadrics

Formulation in Projective Space

In , the real projective space \mathbb{RP}^n is defined as the set of lines through the origin in \mathbb{R}^{n+1}, or equivalently, the quotient space \mathbb{R}^{n+1} \setminus \{\mathbf{0}\} / \sim where \mathbf{x} \sim \lambda \mathbf{x} for \lambda \in \mathbb{R} \setminus \{0\}. Points in \mathbb{RP}^n are represented using [x_1 : \cdots : x_{n+1}], where the coordinates are defined up to nonzero , allowing a uniform treatment of affine points and directions at infinity. A projective quadric in \mathbb{RP}^n is the zero set of a homogeneous Q(\mathbf{x}) = \mathbf{x}^T A \mathbf{x}, where A is a symmetric (n+1) \times (n+1) and \mathbf{x} = (x_1, \dots, x_{n+1})^T \in \mathbb{R}^{n+1}. This defines a consisting of all points [ \mathbf{x} ] \in \mathbb{RP}^n such that Q(\mathbf{x}) = 0, with the homogeneity ensuring that the equation is well-defined under scaling of coordinates. The connection to affine quadrics arises through dehomogenization: in the affine chart where x_{n+1} \neq 0, setting x_{n+1} = 1 yields affine coordinates (u_1, \dots, u_n) = (x_1/x_{n+1}, \dots, x_n/x_{n+1}), transforming the projective equation into a nonhomogeneous in \mathbb{R}^n. Conversely, any affine quadric in \mathbb{R}^n, given by a in these coordinates, homogenizes to a projective quadric by introducing the extra variable x_{n+1} and multiplying linear and constant terms appropriately. This projective completion incorporates points at , corresponding to the x_{n+1} = 0 in , which "closes" the affine quadric into a compact . For instance, the affine x^2/a^2 + y^2/b^2 = 1 in \mathbb{R}^2 homogenizes to the projective conic x^2/a^2 + y^2/b^2 - z^2 = 0 in \mathbb{RP}^2, where the points at (z=0) form the line at intersecting the conic at two points. In the \mathbb{RP}^2, provides a basic result on intersections: two curves of degrees m and n with no common components intersect in exactly mn points, counting multiplicities and including points at . For two projective conics (degree 2 curves), this implies they intersect in 4 points, offering a foundational tool for understanding their geometric interactions.

Normal Forms and Classification

In over the real numbers, the classification of quadrics relies on reducing the associated to a normal form through linear transformations, specifically transformations that preserve the projective structure. A Q(\mathbf{x}) = \mathbf{x}^T A \mathbf{x}, where A is a real , defines the quadric as the set \{\mathbf{x} \in \mathbb{RP}^n \mid Q(\mathbf{x}) = 0 \}. By Sylvester's law of inertia, any such form is congruent to a with p entries of +1, q entries of -1, and r = (n+1) - p - q zeros, where the numbers p, q, and r are invariants under transformations A \mapsto P^T A P for invertible P \in \mathrm{GL}(n+1, \mathbb{R}). This yields the normal form Q(\mathbf{x}) = \sum_{i=1}^p x_i^2 - \sum_{j=1}^q x_{p+j}^2 = 0, where the variables corresponding to zero eigenvalues can be set to zero, effectively reducing to a quadric in a projective of p + q - 1. The of the quadric is k = p + q, measuring its degeneracy (degenerate if k < n+1), and the signature is s = p - q, which distinguishes the geometric type. Non-degenerate quadrics have full k = n+1 and r = 0. Quadrics are classified by their rank and signature. Degenerate cases include pairs of hyperplanes (k = 2) or a double hyperplane (k = 1), but non-degenerate quadrics fall into hyperbolic and elliptic types based on the signature. Hyperbolic quadrics have balanced signature |s| \leq 2 (often p = q or p = q+1), featuring real rulings and hyperbolic cross-sections, such as the hyperboloid of one sheet given by x_1^2 + x_2^2 - x_3^2 - x_4^2 = 0 in \mathbb{RP}^3. Elliptic quadrics have definite or nearly definite signatures (e.g., p = n+1, q = 0), often empty over the reals, like the imaginary ellipsoid x_1^2 + x_2^2 + x_3^2 + x_4^2 = 0 in \mathbb{RP}^3, though they may contain real points in lower-dimensional slices. Parabolic types arise in affine views but are unified projectively via the signature. For conics in the real projective plane \mathbb{RP}^2 (dimension n=2, rank 3), non-degenerate cases have signatures (3,0), (0,3) (empty ovals, no real points), or (2,1) (with real points). The projective classification distinguishes ovals (bounded components, like ellipses, intersecting the line at infinity in 0 points), hyperbolas (two components or unbounded, intersecting in 2 points), and parabolas (unbounded with one vertex, tangent to the line at infinity in 1 point). These distinctions emerge when embedding the projective conic into an affine plane by choosing a line at infinity, with the intersection multiplicity determining the type: 0 for oval/ellipse, 1 for parabola, and 2 for hyperbola. All non-degenerate conics with real points are projectively equivalent over \mathbb{R}, but their affine realizations differ topologically.

Parametrization and Points

Rational Parametrization

A rational parametrization of a quadric provides a birational map from an affine or projective space to the quadric surface, allowing points on the quadric to be expressed using rational functions of parameters. This approach is particularly useful for non-degenerate quadrics, as it establishes that such varieties are rational over fields like the reals, meaning they are birationally equivalent to projective space of one dimension less. The stereographic projection serves as a fundamental method to construct these parametrizations by projecting from a chosen point on the quadric to a complementary line or plane, yielding explicit rational coordinates. For conics in the plane, such as circles, the stereographic projection from a point on the conic to a line provides a birational equivalence to the projective line \mathbb{P}^1. Consider the unit circle defined by x^2 + y^2 = 1 in the affine plane, which can be homogenized to x_0^2 + x_1^2 = x_2^2 in \mathbb{P}^2. Projecting from the point [1, 0, 1] on the conic to the line x_0 = 0 yields the parametrization: [x : y : z] = [1 - t^2 : 2t : 1 + t^2], where t \in \mathbb{R} parameterizes the line via [0 : 1 : t]. Dehomogenizing by setting z = 1 gives the affine map (x, y) = \left( \frac{1 - t^2}{1 + t^2}, \frac{2t}{1 + t^2} \right), which covers all points except the projection point and is rational over the reals. This construction works for any non-degenerate conic over the reals, as the existence of a real point ensures the projection is defined. In three-dimensional Euclidean space, stereographic projection extends to quadric surfaces, mapping from a point on the surface to the plane tangent at the opposite pole, establishing birational equivalence to \mathbb{P}^2. For the unit sphere x^2 + y^2 + z^2 = 1, projecting from the north pole (0, 0, 1) to the equatorial plane z = 0 gives the inverse map (u, v) = \left( \frac{x}{1 - z}, \frac{y}{1 - z} \right), and the forward parametrization: (x, y, z) = \left( \frac{2u}{1 + u^2 + v^2}, \frac{2v}{1 + u^2 + v^2}, \frac{1 - u^2 - v^2}{1 + u^2 + v^2} \right), which is rational over the reals and bijective except at the pole. This method applies to any non-degenerate quadric surface over the reals, provided it is smooth, as the projection avoids singularities and covers the surface densely. Ellipsoids, as affine transformations of spheres, inherit rational parametrizations by scaling the spherical map. For the ellipsoid \frac{x^2}{a^2} + \frac{y^2}{b^2} + \frac{z^2}{c^2} = 1 with a, b, c > 0, projecting from the pole (0, 0, c) to the plane z = 0 yields: (x, y, z) = \left( \frac{2a^2 b^2 u}{b^2 u^2 + a^2 v^2 + a^2 b^2}, \frac{2a^2 b^2 v}{b^2 u^2 + a^2 v^2 + a^2 b^2}, \frac{c (b^2 u^2 + a^2 v^2 - a^2 b^2)}{b^2 u^2 + a^2 v^2 + a^2 b^2} \right), a rational map over the reals that parametrizes the surface birationally. Over the rationals, such parametrizations exist if the quadric has a rational point, enabling the projection to be defined with rational coefficients; otherwise, they may require extensions.

Rational Points on Quadrics

A rational point on a quadric is a point with coordinates in the rational numbers \mathbb{Q} that satisfies the defining quadratic equation. The existence of rational points on a quadric over \mathbb{Q} is governed by the Hasse-Minkowski theorem, which asserts that a quadratic form over \mathbb{Q} represents zero non-trivially if and only if it does so over the real numbers \mathbb{R} and over the p-adic rationals \mathbb{Q}_p for every prime p. This local-global principle provides a complete criterion for the solubility of the corresponding projective quadric over \mathbb{Q}. If a quadric over \mathbb{Q} has at least one , it is birationally equivalent over \mathbb{Q} to , allowing a rational parametrization of all its rational points. This parametrization implies that the set of rational points is infinite. Moreover, since the parameters can be chosen rationally and rationals are dense in the reals, the rational points are dense in the real points of the quadric. A classic example arises on the quadric x^2 + y^2 = z^2 over \mathbb{Q}, whose rational points correspond to Pythagorean triples (x, y, z) with x, y, z \in \mathbb{Q} and \gcd(x, y, z) = 1. These points are generated by the formulas x = m^2 - n^2, y = 2mn, z = m^2 + n^2 for m > n > 0 of opposite . This parametrization produces all primitive Pythagorean triples upon clearing denominators. Hilbert's irreducibility theorem ensures the existence of infinitely many specializations of a general quadric over \mathbb{Q}(t) that retain rational points over \mathbb{Q}, contributing to the abundance of rational points across families of quadrics.

Quadrics over General Fields

Quadratic Forms and Projective Quadrics

A over a F is a Q: V \to F of degree 2 on a finite-dimensional V over F, or equivalently, in coordinates, Q(x_1, \dots, x_n) = \sum_{i,j=1}^n a_{ij} x_i x_j with a_{ij} = a_{ji}. It is associated to a B: V \times V \to F via Q(x) = B(x, x), where B is defined by B(x + y, x + y) - B(x, x) - B(y, y) = 2B(x, y) when \mathrm{char}(F) \neq 2; in characteristic 2, quadratic forms are defined separately without requiring symmetry of B, as the alternation B(x, y) - B(y, x) may not vanish. The associated projective quadric is the subvariety of the \mathbb{P}^n(F) consisting of points = [x_1 : \dots : x_{n+1}] such that Q(x) = 0, where V = F^{n+1}. This defines a projective , and for non-constant Q, it is a quadric variety over F. A quadratic form Q is non-degenerate if the associated B has full rank, meaning the matrix of B has non-zero or, equivalently, for every nonzero v \in V, the linear functional w \mapsto B(v, w) is nonzero. Non-degenerate quadrics correspond to smooth hypersurfaces in projective space, without singular points. Over any F, a non-degenerate (V, Q) admits a Witt decomposition V \cong \bigoplus_{i=1}^r H \oplus A, where each H is a hyperbolic plane (2-dimensional with Witt index 1, spanned by isotropic vectors with B(e, f) = 1), r is the Witt index (dimension of a maximal isotropic ), and A is the anisotropic (the unique anisotropic up to isometry, with no nonzero isotropic vectors). The Witt index measures the maximal dimension of totally isotropic subspaces, and the anisotropic part captures the "indecomposable" core; in characteristic 2, the decomposition holds but hyperbolic planes are defined via alternating forms. Over finite fields F_q with q elements, the Chevalley-Warning implies that any non-degenerate in at least three variables has a nontrivial zero, ensuring the associated projective quadric in \mathbb{P}^n(F_q) with n \geq 2 contains at least one point; moreover, the number of affine points on the cone Q(x) = 0 in F_q^{n+1} is congruent to 0 modulo q, yielding the projective point count as (N - 1)/(q - 1) where N \equiv 0 \pmod{q}. For smooth quadrics in \mathbb{P}^{m-1}(F_q), the exact number of points is \frac{q^{m-1} - 1}{q - 1} if the Witt index is (m-1)/2 (even dimension case) or \frac{q^{m-1} - 1}{q - 1} \pm q^{(m-2)/2} otherwise, depending on the type. This framework generalizes the classification of quadrics over the reals, where anisotropic forms are definite and hyperbolic parts yield indefinite signatures.

Intersections, Radicals, and Subspaces

In over a general k, the of a quadric Q \subset \mathbb{P}^{n-1}_k, defined by a \phi: V \to k on an n-dimensional V, with a line \ell \cong \mathbb{P}^1_k can consist of 0, 1, or 2 points. This follows from restricting \phi to the 2-dimensional corresponding to \ell, yielding a whose roots determine the points; over algebraically closed fields, secant lines intersect at two distinct points, while lines touch at one point with multiplicity two. In general fields, the number of points depends on the splitting of the restricted form, potentially yielding no rational points if anisotropic. Associated to the \phi is its polar b_\phi(u,v) = \phi(u+v) - \phi(u) - \phi(v). The f- of \phi, denoted \mathrm{rad} \, \phi, is \{v \in V \mid \phi(v) = 0 \textrm{ and } b_\phi(v, w) = 0 \ \forall w \in V\}. The q-, or radical of the , is \mathrm{rad} \, b_\phi = \{v \in V \mid b_\phi(v,w) = 0 \ \forall w \in V\}, the to V under b_\phi. For nondegenerate \phi (i.e., \mathrm{rad} \, b_\phi = 0), \mathrm{rad} \, \phi = \{0\}, a proper of the ; in degenerate cases, \mathrm{rad} \, \phi \subseteq \mathrm{rad} \, b_\phi captures the locus. The index of the quadric Q, also called the Witt index of \phi, is the dimension of a maximal isotropic subspace of V, where an isotropic subspace W \subseteq V satisfies b_\phi|_W \equiv 0 (totally singular for the bilinear form). By Witt's decomposition theorem, any quadratic space decomposes as a direct sum of the radical and hyperbolic planes plus an anisotropic kernel, with the index equal to the number of hyperbolic planes (half the dimension of the hyperbolic part). For a nondegenerate form of even dimension $2m, the index is at most m, achieved when \phi is hyperbolic; in odd dimension $2m+1, it is at most m. A q-subspace of Q is a \mathbb{P}(W) \subset Q for some W \subseteq V with \phi|_W = 0, equivalently a totally isotropic subspace under b_\phi. The maximal such subspaces have dimension equal to the of \phi, and over algebraically closed fields, nonsingular quadrics of $2m-1 contain two families of m-1-dimensional q-subspaces. In general fields, the existence and dimension of rational q-subspaces depend on the of \phi, with the of isotropic subspaces parametrizing lines or higher-dimensional flats on Q. A quadric Q is degenerate if \mathrm{rad} \, b_\phi \neq 0, in which case its is the union of the \mathbb{P}(\mathrm{rad} \, b_\phi) and a lower-dimensional quadric on the V / \mathrm{rad} \, b_\phi. For example, in \mathbb{P}^3_k, degenerate quadrics include pairs of (rank 2 form) or a double plane (rank 1), arising when the matrix of \phi has corank at least 1. More generally, the singular locus has \dim \mathrm{rad} \, \phi - 1, and Q decomposes as a over a quadric of dimension one less than the corank allows.

Symmetries and Polar Spaces

The symmetries of a quadric in projective space are governed by the orthogonal group associated to the underlying quadratic form. For a non-degenerate quadratic form Q on a vector space V over a field F, the orthogonal group O(Q) consists of all linear automorphisms g \in \mathrm{GL}(V) such that Q(gv) = Q(v) for all v \in V. In the projective setting, the quadric is the hypersurface \{ \in \mathbb{P}(V) \mid Q(v) = 0 \}, and its automorphism group as a projective variety is the projective orthogonal group \mathrm{PGO}(Q) = O(Q) / Z(O(Q)), where Z(O(Q)) is the kernel of the action on \mathbb{P}(V), typically \{\pm I\} for characteristic not 2. These symmetries act as isometries preserving the geometric structure of the quadric. A representative example is the sphere in Euclidean three-space, which is a quadric defined by Q(x,y,z) = x^2 + y^2 + z^2 = 1. Its group of orientation-preserving isometries is the special orthogonal group \mathrm{SO}(3), a compact of dimension 3 that acts transitively on the sphere. More generally, over the reals, the full O(p,q) preserves indefinite quadratic forms of (p,q), yielding hyperbolic or elliptic geometries on the corresponding quadrics. The polar structure arises from the B associated to Q, defined by B(u,v) = \frac{1}{2} (Q(u+v) - Q(u) - Q(v)) in characteristic not 2, which is non-degenerate if Q is. The polar map sends a point \in \mathbb{P}(V) to its polar \{ \in \mathbb{P}(V) \mid B(x,y) = 0 \}, consisting of points to x with respect to B. For a U \subseteq V, the space is U^\perp = \{ v \in V \mid B(u,v) = 0 \ \forall u \in U \}, satisfying U^{\perp\perp} = U and \dim U + \dim U^\perp = \dim V. This interchanges points and hyperplanes, inducing a duality on the . Polar spaces emerge as incidence geometries derived from these perpendicular structures on quadrics. In the classical , for a non-degenerate on a of dimension at least 3, the polar space has points as the maximal totally singular (isotropic) 1-dimensional subspaces (i.e., points on the quadric), and lines as the maximal totally singular 2-dimensional subspaces (lines lying on the quadric). Axioms ensure that any two points lie on at most one line, and for any point not on a given line, there is a unique line through it intersecting the given line. The O(Q) acts as the collineation group preserving this . In higher dimensions, such polar spaces of rank r \geq 3 form the points and hyperplanes of spherical buildings associated to the , capturing the combinatorial geometry of maximal isotropic subspaces. This framework generalizes to sets, which are subsets S of a such that S intersects every line in at most two points, and the induced structure satisfies polarity-like properties akin to quadrics. Quadrics serve as prototypical quadratic sets, embedding the polar space geometry into broader combinatorial structures over finite fields, where they yield finite geometries with applications in group theory and .

References

  1. [1]
    Quadratic Surface -- from Wolfram MathWorld
    A second-order algebraic surface given by the general equation (1) Quadratic surfaces are also called quadrics, and there are 17 standard-form types.
  2. [2]
    Calculus III - Quadric Surfaces - Pauls Online Math Notes
    Nov 16, 2022 · In this section we are going to be looking at quadric surfaces. Quadric surfaces are the graphs of any equation that can be put into the ...
  3. [3]
    11.6: Quadric Surfaces - Mathematics LibreTexts
    Sep 5, 2018 · Definition: traces. The traces of a surface are the cross-sections created when the surface intersects a plane parallel to one of the coordinate ...
  4. [4]
    Quadric Surfaces - GeeksforGeeks
    Jul 23, 2025 · Quadric surfaces are three-dimensional shapes like ellipsoids, hyperboloids, or paraboloids, described by second-degree equations in three variables.
  5. [5]
    [PDF] math 137 notes: undergraduate algebraic geometry
    Jan 25, 2016 · Definition 9.15. A quadric X ⊂ Pn = PV with V ∼= kn+1 is the zero locus of a single homogeneous quadric polynomial Q. Remark 9.16. We have ...
  6. [6]
    [PDF] 2. Quadrics.
    Definition 1.1 A quadric is a surface in E3 determined, in affine coordinates, by a quadratic equation. In this way the general equation of a quadric will be:.
  7. [7]
    [PDF] Quadratic forms beyond arithmetic - UCLA Mathematics
    We introduce the quadric hypersurface (quadric) Xq associated with a ... Thus q is given by a quadratic homogeneous polynomial over F. We say that q ...
  8. [8]
    [PDF] Affine Maps, Euclidean Motions and Quadrics
    The standard example of an affine space is given by A = E, that is, the “points” of this affine space are the elements of the vector space. The action is. A ...
  9. [9]
  10. [10]
    None
    Below is a merged and comprehensive summary of the content on quadric surfaces (e.g., ellipsoids, hyperboloids) and Gaussian curvature from Do Carmo’s *Differential Geometry of Curves and Surfaces*, based on the provided segments. To retain all information in a dense and organized manner, I will use a combination of narrative text and tables in CSV format where appropriate. The source document is consistently referenced as http://home.ustc.edu.cn/~huangty0407/DGNotes/Do%20Carmo%20-%20Differential%20Geometry%20of%20Curves%20and%20Surfaces.pdf, and no additional URLs are provided in the content.
  11. [11]
    Conic Section -- from Wolfram MathWorld
    The pedal curve of a conic section with pedal point at a focus is either a circle or a line. In particular the ellipse pedal curve and hyperbola pedal curve are ...
  12. [12]
    Ellipse -- from Wolfram MathWorld
    ### Summary of Ellipse Properties (Euclidean Plane)
  13. [13]
    Parabola -- from Wolfram MathWorld
    ### Summary of Parabola Properties (Euclidean Plane)
  14. [14]
    Hyperbola -- from Wolfram MathWorld
    ### Summary of Hyperbola Properties (Euclidean Plane)
  15. [15]
    Index: Ruled Surfaces - Arizona Math
    Some quadratic surfaces are ruled: hyperboloids of one sheet, hyperbolic paraboloids, and quadratic cones and cylinders. The first two are doubly ruled ...
  16. [16]
    A model of a hyperboloid of one sheet and its asymptotic cone
    Oct 31, 2008 · In this article is described the construction of a thread model of a hyperboloid of one sheet (H) and its asymptotic cone (C).
  17. [17]
    [PDF] Gallery of Surfaces Math 131 Multivariate Calculus
    Like the hyperboloid of one sheet, the hyperbolic paraboloid is a doubly ruled surface. Through each its points there are two lines that lie on the surface. ...
  18. [18]
    Chapter 6 - Quadric hypersurfaces - EMS Press
    Apr 6, 2023 · If p is an imaginary point of a quadric Q of a real projective space, then its complex-conjugate .p/ is also an imaginary point of Q. The easy ...<|control11|><|separator|>
  19. [19]
    [PDF] Basics of Projective Geometry - UPenn CIS
    With respect to this projective frame, note that an+2 has homogeneous coordinates (1,...,1), and that ai has homogeneous coordinates (0,...,1,...,0), where the ...
  20. [20]
    [PDF] 3 Quadrics - People
    We can add symmetric bilinear forms: (B+C)(v, w) = B(v, w)+C(v, w) and multiply by a scalar (λB)(v, w) = λB(v, w) so they form a vector space isomorphic to ...Missing: hypersurface | Show results with:hypersurface<|control11|><|separator|>
  21. [21]
    [PDF] Bézout's Theorem for curves - UChicago Math
    Aug 26, 2011 · Bézout's Theorem describes the number of times two algebraic curves intersect in projective space, up to multi- plicity.
  22. [22]
    Sylvester's Inertia Law -- from Wolfram MathWorld
    Sylvester's Inertia Law: The numbers of eigenvalues that are positive, negative, or 0 do not change under a congruence transformation.Missing: reference | Show results with:reference
  23. [23]
    [PDF] plane conics in algebraic geometry
    Aug 29, 2014 · After classifying projective conics, we can classify affine conics by paying atten- tion to the line at infinity in projective spaces. Examining ...
  24. [24]
    [PDF] Quadric and Cubic Hypersurface Parameterization - Purdue e-Pubs
    While the rational parametric representation of a curve or surface allows greater ease for transformation and shape control, the implicit representation is.
  25. [25]
    [PDF] Stereographic Projections on Some Quadric Surfaces - arXiv
    Jun 9, 2025 · We begin by constructing the stereographic projections for both quadric surfaces separately, analyzing the geometric particularities of each ...
  26. [26]
    [PDF] Hasse's Theorem and Rational Points on the General Conic
    In this paper we will detail the theory for conics. We define a point in the plane to be rational if its coordinates are rational. We define a curve to be ...
  27. [27]
    [PDF] THE LOCAL-GLOBAL PRINCIPLE 1. Introduction Hensel created p ...
    Then we will state Hasse's version of Minkowski's theorem on quadratic forms over Qp as an example where the local–global principle works. Next we will see.
  28. [28]
    [PDF] Pythagorean triples: Rational points on the circle - Niles Johnson
    Jan 28, 2019 · A Pythagorean triple is a triple of whole numbers (a, b, c) such that a2 +b2 = c2. It is easy to make a right triangle where the two legs ...
  29. [29]
    [PDF] Hilbert's irreducibility theorem and the larger sieve
    In this paper, we are interested in quantitative versions of Hilbert's irreducibility theorem (HIT). ... The larger sieve for rational points. Theorem 3.1 (Larger ...
  30. [30]
    [PDF] 9.1 Quadratic forms
    Oct 3, 2013 · This defines an equivalence relation on the set of all quadratic forms of the same dimension over the field k.
  31. [31]
    [PDF] Quadratic forms Chapter I: Witt's theory
    In these notes we give a detailed treatment of the foundations of the algebraic theory of quadratic forms, starting from scratch and ending with Witt Cancella-.
  32. [32]
    [PDF] Quadratic forms and their invariants.
    Jun 1, 2007 · To each quadratic form q one can asign the respective projective quadric. Q ⊂ P(Vq) given by the equation q = 0. If q is nondegenerate, the ...<|separator|>
  33. [33]
    Quadratic forms over finite fields - linear algebra - MathOverflow
    Mar 4, 2010 · Over a finite field, any quadratic form in three or more variables admits a non-trivial zero. This is a consequence of the Chevalley-Warning Theorem.Proofs of the Chevalley-Warning Theorem - MathOverflowChevalley–Warning theorem for rational field Q - MathOverflowMore results from mathoverflow.net
  34. [34]
    [PDF] 1 Projective space 2 Chevalley-Warning
    Below we shall need that a quadric on a projective plane over a finite field F has at least one point. This is a very special case of the following. Theorem 2.1 ...
  35. [35]
    Number of points of a quadric hypersurface over a finite field
    Feb 3, 2022 · The number of points of a smooth quadric hypersurface is qn−1q−1 if n is even or qn−1q−1±qn−12 if n is odd. This can be proven in a number of ...Forms over finite fields and Chevalley's theorem - MathOverflowProofs of the Chevalley-Warning Theorem - MathOverflowMore results from mathoverflow.net
  36. [36]
    [PDF] Algebraic and Geometric Theory of Quadratic Forms
    Historically, the powerful approach using algebraic geometry has been the last to be developed. ... radical of b and denoted by rad b. The form b is non ...
  37. [37]
    [PDF] Chapter 3 Quadratic curves, quadric surfaces
    12 Degenerate quadrics. As with quadratic curves, some quadric surfaces are more or less 'degenerate'. Here are typical examples. Page 22. 78. Quadratic curves ...
  38. [38]
    [PDF] 13 Orthogonal groups - UC Berkeley math
    Orthogonal groups are the groups preserving a non-degenerate quadratic form on a vector space. Over the complex numbers there is essentially only one such.
  39. [39]
    On the foundations of polar geometry
    FRANCIS BUEKENHOUT AND ERNEST SHULT. ON THE FOUNDATIONS OF POLAR GEOMETRY. 1. INTRODUCTION. 1.0. Brief Summary of Results. The problem of characterizing the ...
  40. [40]
    [PDF] 6 Polar spaces
    Such geometries play much the same rôle in the theory of polar spaces as projective planes do in the theory of projective spaces. We will return to them later.