Fact-checked by Grok 2 weeks ago

Sphere

A sphere is a three-dimensional geometrical object consisting of all points in space that are equidistant from a fixed point called the center, with the common distance being the radius r. This perfectly symmetrical shape forms the surface of a ball and is analogous to a circle in two dimensions. Unlike polyhedra, a sphere has no edges, vertices, or flat faces, making it one of the simplest curved surfaces in Euclidean geometry. The mathematical properties of a sphere are well-defined and in various fields. Its surface area is given by the , representing the area enclosing the volume. The volume enclosed by a sphere is \frac{4}{3}\pi r^3, derived from integral calculus or Archimedean methods dating back to . Spheres exhibit around any and are invariant under orthogonal transformations, contributing to their ubiquity in physics, such as modeling planetary bodies or atomic orbitals. Historically, the concept of the sphere has roots in ancient , with Greek philosophers like associating it with ideal forms and astronomers like using for celestial models. In modern applications, spheres appear in optimization problems, for rendering rounded objects, and for designing pressure vessels due to their uniform stress distribution. Generalizations include hyperspheres in higher dimensions, known as n-spheres, which extend the definition to n-dimensional .

Definitions and Terminology

Geometric Definition

In three-dimensional , a is defined as the set of all points that are equidistant from a fixed point called , with this distance denoted as the r. This locus-based characterization captures the as a perfectly symmetric surface generated by points maintaining a constant separation from . Historically, formalized the concept in his (Book XI, Definition 14) as the solid figure comprehended when a is rotated about its fixed , emphasizing the inherent to the shape. The sphere specifically refers to the two-dimensional surface bounding the region, distinct from the , which encompasses the three-dimensional solid interior including all points at distances less than or equal to the from the center. This distinction is crucial in , as the sphere constitutes the alone, while the includes the enclosed ; mathematicians rigorously maintain this separation, though colloquial usage sometimes blurs the terms. Intuitively, the sphere appears as a , rounded surface where no point protrudes or recedes relative to others from the center, evoking the shape of a or . The , the maximum straight-line across the sphere passing through the center, measures exactly twice the , providing a fundamental scale for the object's extent. This definition presupposes the axioms of , including the straight-line metric for , without delving into non-Euclidean alternatives. As the three-dimensional counterpart to , the sphere extends planar roundness into spatial .

Key Terms and Notations

In mathematics, the center of a sphere, often denoted by O or C, is the fixed point equidistant from all points on its surface. The radius r represents this constant distance from the center to any point on the surface. The diameter d, defined as d = 2r, is the straight-line distance between two points on the surface passing through the center. These elements form the basis for the sphere's defining equation in vector notation, commonly abbreviated as \|\mathbf{x} - \mathbf{c}\| = r, where \mathbf{x} is a point on the sphere and \mathbf{c} is the center vector. The unit sphere refers to a sphere of radius 1, typically centered at the origin, and is denoted S^n in n+1-dimensional Euclidean space, with S^2 standard for the three-dimensional case as the set of unit vectors. A hollow sphere describes the boundary surface alone, excluding its interior, in contrast to a solid ball (or simply ball), which includes all points within and on the surface up to radius r from the center. For surfaces like the sphere, specifies a consistent choice of normal vector field; a positively oriented surface, such as the standard sphere, uses the outward-pointing unit normal, distinguishing an "inside" from an "outside."/03%3A_Surface_Integrals/3.05%3A_Orientation_of_Surfaces) This convention aligns with the sphere's role as a closed orientable surface in ./03%3A_Surface_Integrals/3.05%3A_Orientation_of_Surfaces) Notation for spheres has evolved from descriptive geometric terms in , such as Euclid's qualitative characterizations without symbolic equations, to modern vector and algebraic forms developed in the alongside coordinate and vector . This progression parallels the extension of terminology into higher dimensions.

Mathematical Equations

Cartesian Equation

The Cartesian equation of a sphere arises from the geometric definition as the set of all points in three-dimensional Euclidean space equidistant from a fixed center point. Consider a sphere with center at (a, b, c) and radius r > 0. The distance from the center to any point (x, y, z) on the sphere's surface is r, given by the Euclidean distance formula: \sqrt{(x - a)^2 + (y - b)^2 + (z - c)^2} = r. Squaring both sides to eliminate the square root yields the standard Cartesian equation: (x - a)^2 + (y - b)^2 + (z - c)^2 = r^2. This implicit equation defines the sphere algebraically, representing all points satisfying the fixed distance condition from the center. For a sphere centered at the origin (0, 0, 0), the equation simplifies to: x^2 + y^2 + z^2 = r^2, known as the unit sphere when r = 1: x^2 + y^2 + z^2 = 1. This form is fundamental in vector analysis and coordinate geometry, where the sphere serves as a model for isotropic distributions in space. Expanding the standard equation produces a general : x^2 + y^2 + z^2 - 2ax - 2by - 2cz + (a^2 + b^2 + c^2 - r^2) = 0. This is a special case of the general surface equation Ax^2 + By^2 + Cz^2 + Dx + Ey + Fz + G = 0, where A = B = C = 1 and there are no cross-product terms (xy, xz, yz), distinguishing the sphere from other quadrics like ellipsoids or hyperboloids. The equal coefficients for the squared terms ensure about the center. This algebraic representation generalizes the equation of a circle in the plane, x^2 + y^2 = r^2, by adding the z^2 term to account for the third dimension.

Parametric Equations

The parametric equations provide an explicit way to describe points on the surface of a sphere, facilitating computations in visualization, computer graphics, and vector calculus such as surface integrals over the sphere. For a sphere of radius r > 0 centered at the origin, the standard parametrization using spherical coordinates is given by \begin{align*} x &= r \sin \theta \cos \phi, \\ y &= r \sin \theta \sin \phi, \\ z &= r \cos \theta, \end{align*} where \theta is the colatitude (polar angle from the positive z-axis), ranging from $0 to \pi, and \phi is the azimuth (longitude angle in the xy-plane), ranging from $0 to $2\pi. This parametrization maps the rectangular domain [0, \pi] \times [0, 2\pi) in the \theta-\phi plane bijectively onto the sphere, excluding the identification of the meridian at \phi = 0 and \phi = 2\pi. These coordinates satisfy the Cartesian equation x^2 + y^2 + z^2 = r^2 as a constraint. One derivation of these equations arises from generating the sphere as a by rotating a in the xz-plane around the z-axis. The of radius r in the xz-plane (with y = 0) is parametrized as x = r \sin \theta, y = 0, z = r \cos \theta for \theta \in [0, \pi]. Applying a rotation by angle \phi around the z-axis transforms the coordinates via the , yielding x' = (r \sin \theta) \cos \phi, y' = (r \sin \theta) \sin \phi, z' = r \cos \theta. Alternatively, the parametrization can be obtained using , where \theta and \phi correspond to the first two rotation angles defining the direction from the to the point on the unit sphere, scaled by r./16:_Vector_Calculus/16.06:_Parametric_Surfaces_and_Their_Areas) In applications involving surface integrals, the Jacobian factor arises from the magnitude of the of the vectors of the position vector \mathbf{r}(\theta, \phi) = (x, y, z). Specifically, \mathbf{r}_\theta = \frac{\partial \mathbf{r}}{\partial \theta} = r (\cos \theta \cos \phi, \cos \theta \sin \phi, -\sin \theta) and \mathbf{r}_\phi = \frac{\partial \mathbf{r}}{\partial \phi} = r \sin \theta (-\sin \phi, \cos \phi, 0), with \| \mathbf{r}_\theta \times \mathbf{r}_\phi \| = r^2 \sin \theta. Thus, the surface element is dS = r^2 \sin \theta \, d\theta \, d\phi, enabling integrals like the surface area $4\pi r^2 = \int_0^{2\pi} \int_0^\pi r^2 \sin \theta \, d\theta \, d\phi. For curves on the sphere, such as meridians (fixed \phi, varying \theta) or parallels (fixed \theta, varying \phi), the velocity vectors are the partial derivatives \mathbf{r}_\theta and \mathbf{r}_\phi, which form an for the at each point (up to scaling), with lengths \| \mathbf{r}_\theta \| = r and \| \mathbf{r}_\phi \| = r \sin \theta. These vectors are to the position vector \mathbf{r}, reflecting the sphere's where tangent directions lie orthogonal to the radial direction.

Equations in Other Coordinate Systems

In cylindrical coordinates (\rho, \phi, z), where \rho = \sqrt{x^2 + y^2} is the radial distance from the z-axis, \phi is the azimuthal , and z is the along the z-axis, the equation of a sphere of radius r centered at the simplifies to \rho^2 + z^2 = r^2. The azimuthal \phi remains free, ranging from 0 to $2\pi, reflecting the around the z-axis. This form is particularly useful in applications involving , such as or , where the z-axis aligns with the system's principal direction. In spherical coordinates (\rho, \theta, \phi), where \rho is the radial distance from the , \theta is the polar angle from the positive z-axis (0 to \pi), and \phi is the azimuthal angle (0 to $2\pi), the equation of the sphere centered at the becomes simply \rho = r. This representation highlights the sphere as a constant-radius surface, ideal for problems in or that exploit radial symmetry. However, this introduces singularities at the poles (\theta = 0 and \theta = \pi), where the azimuthal angle \phi becomes undefined, leading to coordinate discontinuities that complicate numerical integrations or visualizations near these points. Transformations between these coordinate systems and Cartesian coordinates facilitate deriving sphere equations in alternative bases. The conversion from cylindrical to Cartesian is given by: \begin{pmatrix} x \\ y \\ z \end{pmatrix} = \begin{pmatrix} \rho \cos \phi \\ \rho \sin \phi \\ z \end{pmatrix}, with the inverse: \rho = \sqrt{x^2 + y^2}, \quad \phi = \atan2(y, x), \quad z = z. For spherical to Cartesian: \begin{pmatrix} x \\ y \\ z \end{pmatrix} = \begin{pmatrix} \rho \sin \theta \cos \phi \\ \rho \sin \theta \sin \phi \\ \rho \cos \theta \end{pmatrix}, and the inverse involves: \rho = \sqrt{x^2 + y^2 + z^2}, \quad \theta = \arccos\left(\frac{z}{\rho}\right), \quad \phi = \atan2(y, x). Between cylindrical and spherical, the relations are \rho = r \sin \theta, z = r \cos \theta, and \phi = \phi, where r here denotes the cylindrical radial coordinate to distinguish from the sphere's radius./12:_Vectors_in_Space/12.07:_Cylindrical_and_Spherical_Coordinates) These transformations, often represented via Jacobian matrices for volume elements or gradients, preserve the sphere's equation under substitution into the Cartesian form x^2 + y^2 + z^2 = r^2. In , spheres are represented using [x : y : z : w], where points in three-dimensional correspond to rays through the origin in , with dehomogenization via x/w, y/w, z/w for w \neq 0. The equation of a centered at the origin becomes the : x^2 + y^2 + z^2 - w^2 = 0. This homogeneous quadratic form extends the Euclidean sphere to the projective plane at infinity, enabling unified treatments of conics and quadrics in and , such as intersection computations with planes. For a general sphere of r, the equation scales to x^2 + y^2 + z^2 = r^2 w^2.

Geometric Properties

Volume and Surface Area

The surface area of a sphere of radius r is given by A = 4\pi r^2. This result was first established by Archimedes in his treatise On the Sphere and Cylinder, where Proposition 33 demonstrates that the surface area equals four times the area of the great circle, achieved through a method of exhaustion comparing the sphere to inscribed and circumscribed polyhedra without relying on limits or calculus. In modern terms, the surface area can be derived via surface integration in spherical coordinates, where the parameterization \mathbf{r}(\theta, \phi) = (r \sin\phi \cos\theta, r \sin\phi \sin\theta, r \cos\phi) for $0 \leq \theta \leq 2\pi and $0 \leq \phi \leq \pi yields the surface element dS = r^2 \sin\phi \, d\theta \, d\phi, and integrating \iint dS = \int_0^{2\pi} \int_0^\pi r^2 \sin\phi \, d\phi \, d\theta = 4\pi r^2. The volume enclosed by the sphere, known as the volume of the ball, is V = \frac{4}{3} \pi r^3. derived this in 34 of the same work by comparing the sphere to a circumscribed and inscribed , showing the sphere's is two-thirds that of the (of height and base r) minus the 's , yielding the exactly. A non-calculus approach uses , equating the sphere's cross-sectional areas (circles of \sqrt{r^2 - z^2} at height z) to those of a of r and height $2r with two s of height r and base r removed; the is then \pi r^2 (2r) - 2 \cdot \frac{1}{3} \pi r^2 r = \frac{4}{3} \pi r^3. Using , the follows from the triple in spherical coordinates: V = \int_0^{2\pi} \int_0^\pi \int_0^r \rho^2 \sin\phi \, d\rho \, d\phi \, d\theta = \frac{4}{3} \pi r^3. These formulas apply in three-dimensional , where has units of cubed and surface area length squared, both scaling with powers of the . In higher , the of an n-ball of r generalizes to V_n(r) = \frac{\pi^{n/2} r^n}{\Gamma(n/2 + 1)}, and the surface area of the bounding (n-1)-sphere to A_{n-1}(r) = \frac{2 \pi^{n/2} r^{n-1}}{\Gamma(n/2)}, revealing that as n increases, the unit ball's approaches zero while surface area peaks before declining.

Diameter, Radius, and Circumference

In geometry, the radius of a sphere is defined as the distance from its center to any point on its surface. The diameter is twice the radius and represents the longest straight-line distance between any two points on the sphere, equivalent to the length of a chord passing through the center. A great circle on a sphere is the intersection of the sphere with a plane passing through its center, forming a circle of the same radius as the sphere. The circumference of this great circle is $2\pi r, where r is the sphere's radius, mirroring the circumference formula for a two-dimensional circle of equal radius. This connection highlights how the sphere extends circular geometry into three dimensions, with great circles serving as the sphere's "equators." Jung's theorem provides a key bound relating the of a set of points in to the of the smallest enclosing sphere. In three-dimensional , any set of points with d (the supremum of distances between pairs of points) can be enclosed by a sphere of at most d \sqrt{3/8}, implying the enclosing sphere's is at most d \sqrt{3/2}. This result, originally established by Heinrich Jung in , quantifies how the sphere's limits the spread of contained sets and is tight for certain configurations like the vertices of a regular . In geometric applications, the diameter of a bounding sphere—often the smallest enclosing sphere for a finite set of points—directly constrains the maximum pairwise distances within that set, with equality achieved when the points lie on the sphere's surface antipodally. This relation is fundamental in computational geometry for tasks like clustering and approximation, where the sphere's linear dimensions provide efficient bounds on set diameters without exhaustive distance computations.

Symmetry and Isometry Groups

The rotational symmetries of the sphere in three-dimensional Euclidean space are described by the special orthogonal group SO(3), which consists of all orientation-preserving isometries fixing the origin and thus mapping the unit sphere S^2 to itself. This group has three degrees of freedom, corresponding to the three independent angles required to specify an arbitrary rotation, such as Euler angles for yaw, pitch, and roll. The action of SO(3) on S^2 is transitive, meaning the orbit of any point on the sphere under this group action is the entire sphere. The stabilizer subgroup of a point on the sphere under the SO(3) action is isomorphic to SO(2), consisting of rotations around the axis passing through that point and the origin, which fixes the point (and its antipodal point) while rotating the equator. For instance, rotations about the z-axis fix the north and south poles. The full isometry group of the sphere, including reflections and improper rotations, is the orthogonal group O(3), which comprises SO(3) together with elements of determinant -1 that reverse orientation, such as reflections through planes passing through the origin. These isometries preserve distances and map the sphere to itself but may turn it "inside out." Finite subgroups of SO(3) correspond to the rotational symmetries of Platonic solids that can be inscribed in the sphere, providing discrete approximations to the full . For example, the and its dual, the , each admit 24 rotational symmetries, isomorphic to the S_4, generated by rotations about axes through vertices, face centers, and edge midpoints.

Advanced Geometric Properties

Pencil of Spheres

A of spheres is a one-parameter family of spheres formed by taking linear combinations of the Cartesian equations of two given spheres. The equation of a sphere in three-dimensional is x^2 + y^2 + z^2 + Dx + Ey + Fz + G = 0. For two such spheres with equations S_1(x,y,z) = 0 and S_2(x,y,z) = 0, the pencil consists of all surfaces defined by \lambda S_1 + \mu S_2 = 0, where \lambda and \mu are real scalars not both zero. This family includes all spheres passing through the of of the original two spheres, which is generally a if they intersect along one. The radical plane of the two generating spheres plays a central role in the geometry of the pencil. It is the locus of points having equal power with respect to both spheres and is given by the linear equation S_1 - S_2 = 0. For spheres with centers \mathbf{m} = (m_1, m_2, m_3) and \mathbf{n} = (n_1, n_2, n_3) and radii r_1, r_2, the radical plane equation is $2(m_1 - n_1)x + 2(m_2 - n_2)y + 2(m_3 - n_3)z = |\mathbf{m}|^2 - |\mathbf{n}|^2 + r_2^2 - r_1^2, perpendicular to the line joining the centers. For non-intersecting spheres, this plane separates regions where the powers differ in sign, and it contains the circle of intersection when the spheres do intersect. Spheres orthogonal to one member of the pencil are orthogonal to all, with their centers lying on the radical plane. Pencils of spheres provide a powerful tool for solving classical geometric problems, particularly the three-dimensional Apollonius problem of constructing spheres to four given spheres. Solutions arise as common spheres to pencils generated by pairs of the given spheres; for instance, the pencil from two given spheres yields candidates to the remaining two via conditions on their radical planes. Up to eight real solutions exist in general, analogous to the planar case. Degenerate cases within the pencil occur as limits of the family parameters. When the effective radius squared R^2 = A^2 + B^2 + C^2 - G (in normalized form) equals zero, the surface reduces to a point sphere, representing a degenerate sphere of zero radius at the center (A, B, C). For R^2 < 0, no real sphere exists. Planes appear as degenerate spheres with infinite radius, corresponding to cases where the quadratic terms vanish in the equation, effectively yielding a linear plane equation as a sphere at infinity. These limits facilitate transitions between spherical and planar geometries in problem-solving.

Inversion Geometry

Inversion geometry involves a transformation that maps points in three-dimensional space relative to a reference sphere, preserving certain geometric structures and angles. Specifically, given a sphere with center O and radius k, the inversion maps a point P to its inverse P' such that P' lies on the ray from O through P and satisfies the relation OP \cdot OP' = k^2. This defines a one-to-one correspondence between points excluding O, with O mapping to the point at infinity in the extended space. Key properties of sphere inversion include its action on spheres and planes, as well as its conformal nature. Under inversion, any sphere or plane not passing through O maps to another sphere, while a sphere or plane passing through O maps to a plane. Additionally, the transformation preserves angles between curves, making it conformal and useful for studying geometric configurations without altering local orientations. These mappings extend the plane case, where circles and lines transform similarly, to three dimensions. The inversion transformation arises in the broader framework of similarity transformations, as Möbius transformations—which generalize inversions to spheres—are compositions of similarities and inversions. This connection highlights how inversion complements similarities by introducing a radial scaling that inverts distances from the center. A representative application involves the intersection of two spheres, which forms a circle. If the inversion center O lies on this circle, the two spheres map to planes, and their intersection circle—passing through O—maps to a straight line, the intersection of those planes. This simplifies computations for coaxial systems of spheres, relating to pencils where inversions transform families into parallel planes.

Stereographic Projection

Stereographic projection is a perspective mapping that projects points on the surface of a onto a plane, typically from a designated pole on the sphere to an equatorial plane. For the unit sphere centered at the origin in three-dimensional Euclidean space, the standard setup projects from the north pole N = (0, 0, 1) onto the plane z = 0. A point P = (x, y, z) on the sphere, excluding the north pole, is mapped by extending the line from N through P until it intersects the plane, yielding coordinates (x', y') = \left( \frac{x}{1 - z}, \frac{y}{1 - z} \right). This construction establishes a bijection between the sphere minus the north pole and the entire plane. The projection preserves angles locally, making it a conformal map, as the differential of the mapping multiplies the metric by a positive scalar factor without distortion of infinitesimal angles. It also maps circles on the sphere—intersections with planes—to circles or straight lines on the plane, with circles passing through the north pole projecting to straight lines and those not passing through it to circles. The inverse projection from the plane (u, v) back to the sphere is given by (x, y, z) = \left( \frac{2u}{u^2 + v^2 + 1}, \frac{2v}{u^2 + v^2 + 1}, \frac{u^2 + v^2 - 1}{u^2 + v^2 + 1} \right), which covers the entire sphere except the north pole. To include the north pole, the mapping extends to the , where the plane is identified with the complex plane \mathbb{C} via w = u + iv, and the north pole corresponds to the point at infinity, compactifying \mathbb{C} into a topologically equivalent sphere. This projection finds applications in complex analysis, where the Riemann sphere facilitates the study of meromorphic functions and by providing a uniform framework for points at infinity. It also aids in visualizing spherical data on flat surfaces, such as in cartography for conformal world maps or in computer graphics for rendering spherical textures onto planar displays.

Geometry on the Sphere

Spherical Geometry Fundamentals

Spherical geometry is the study of geometric figures on the surface of a sphere, which has constant positive curvature. It is closely related to , the latter being obtained by identifying antipodal points on the sphere to form the . This geometry is finite yet boundless, as one can traverse the entire surface without encountering an edge, contrasting with the infinite expanse of . The sphere's positive curvature implies that distances and angles behave differently from flat space, leading to a closed, compact . The axioms of spherical geometry align with the first four of Euclid's postulates but diverge fundamentally in the parallel postulate. In this system, no parallel lines exist; every pair of great circles, which act as the "straight lines" of the geometry, intersects at two antipodal points. This failure of the parallel postulate—replaced by the assertion that through any point not on a given line, every line intersects it—marks spherical geometry as non-Euclidean, eliminating the possibility of parallel transport over indefinite distances without convergence. Consequently, the geometry satisfies for absolute geometry but requires modification for parallelism, ensuring all lines meet. A hallmark of spherical geometry is the spherical excess observed in triangles, where the sum of interior angles exceeds \pi radians, with the excess proportional to the triangle's area. This phenomenon arises directly from the positive curvature, distinguishing it from Euclidean triangles where the angle sum equals \pi. For instance, an equilateral spherical triangle with small side lengths approximates Euclidean behavior, but as sizes increase, the angle sum grows, reflecting the enclosed area on the curved surface. Basic theorems underscore the metric structure of spherical geometry as a space where two distinct points determine a unique geodesic, realized as the shorter arc of the great circle connecting them, provided the points are not antipodal. This uniqueness ensures that spherical geometry functions as a complete metric space for distances up to half the circumference, supporting congruence criteria like SAS and SSS, while AAA implies full congruence due to the absence of similar non-congruent figures. These properties facilitate rigorous treatments of navigation and astronomy on spherical surfaces.

Great and Small Circles

A great circle on a sphere is formed by the intersection of the sphere with a plane passing through its center, resulting in a circle whose diameter equals that of the sphere. These circles represent the largest possible circles on the sphere and serve as the shortest paths, or geodesics, between any two points on the surface. The full circumference of a great circle is $2\pi r, where r is the radius of the sphere. In contrast, a small circle arises from the intersection of the sphere with a plane that does not pass through the center, producing a circle with a diameter smaller than that of the sphere. The circumference of a small circle is given by $2\pi r \sin(\alpha), where \alpha is the angular radius of the circle relative to the sphere's center (with \alpha < \pi/2 for small circles). Small circles are parallel to great circles in certain orientations, such as lines of latitude on . Great circles possess unique properties, including the division of the sphere into two equal hemispheres, as the plane through the center bisects the surface symmetrically. They play a critical role in navigation, where routes along great circles minimize travel distance over the Earth's surface, as utilized in aviation and maritime applications. Small circles, meanwhile, bound spherical caps, which are the portions of the sphere cut off by the plane, forming regions of varying size depending on the plane's distance from the center. Examples of great circles include the equator, which divides into northern and southern hemispheres. Small circles are exemplified by the Tropic of Cancer and Tropic of Capricorn, located at approximately 23.5° north and south latitudes, respectively, with circumferences shorter than the equatorial great circle due to their offset planes.

Geodesics and Distances

On a sphere, geodesics are the shortest paths between two points and correspond to arcs of great circles, which are the intersections of the sphere with planes passing through its center. The length d of a geodesic arc between two points is given by d = r \theta, where r is the radius of the sphere and \theta is the central angle subtended by the arc at the sphere's center, measured in radians. To compute the geodesic distance between two points specified by latitude and longitude coordinates, the haversine formula provides an exact method based on spherical trigonometry. The formula calculates the central angle \theta as follows: \begin{align*} a &= \sin^2\left(\frac{\Delta\phi}{2}\right) + \cos\phi_1 \cdot \cos\phi_2 \cdot \sin^2\left(\frac{\Delta\lambda}{2}\right), \\ c &= 2 \cdot \atan2\left(\sqrt{a}, \sqrt{1 - a}\right), \\ d &= r \cdot c, \end{align*} where \phi_1, \phi_2 are the latitudes, \Delta\phi = \phi_2 - \phi_1, and \Delta\lambda is the difference in longitudes. For small distances where \theta is much less than \pi radians, an approximate form simplifies to the planar distance formula d \approx r \sqrt{(\Delta\phi)^2 + (\cos\phi_1 \cdot \Delta\lambda)^2}, treating the sphere locally as a tangent plane. In spherical trigonometry, the area A of a spherical triangle formed by three geodesics is determined by its spherical excess E = \alpha + \beta + \gamma - \pi, where \alpha, \beta, \gamma are the interior angles in radians; Girard's theorem states that A = r^2 E. This relation highlights how angular excess on the sphere directly measures enclosed area, unlike in Euclidean geometry where the angle sum is fixed at \pi. For practical navigation, traveling along a geodesic requires continuous adjustment of the bearing, as the initial heading at the starting point differs from subsequent directions along the great circle path due to the sphere's curvature. This convergence or divergence of meridians necessitates computational tools or charts to maintain the geodesic route.

Differential Geometry of the Sphere

Intrinsic and Extrinsic Curvature

In differential geometry, the extrinsic curvature of a sphere describes its bending when embedded in the ambient Euclidean space \mathbb{R}^3. For a sphere of radius r, the surface is an umbilic at every point, meaning the two principal curvatures \kappa_1 and \kappa_2 are equal and given by \kappa_1 = \kappa_2 = \frac{1}{r}. These principal curvatures represent the maximum and minimum normal curvatures at each point, quantifying the rate of bending relative to the surface normal in the embedding space. In contrast, the intrinsic curvature of the sphere is independent of its embedding and can be determined solely from the Riemannian metric on the surface itself. The Gaussian curvature K, a key intrinsic invariant, is constant and positive for the sphere, with K = \frac{1}{r^2}, computed as the product of the principal curvatures via . This value arises from the second fundamental form in relation to the first fundamental form, reflecting the sphere's inherent geometry as perceived by measurements confined to its surface. For curves embedded on the sphere, the geodesic curvature \kappa_g measures the deviation of the curve from a geodesic within the intrinsic geometry of the surface, relying only on the metric tensor. On the sphere, great circles serve as geodesics with \kappa_g = 0, while other curves exhibit nonzero \kappa_g depending on their path. The sphere's positive Gaussian curvature K = \frac{1}{r^2} distinguishes it from the Euclidean plane, which has K = 0 and admits flat embeddings without distortion, and from the hyperbolic plane, which has constant negative curvature K < 0 and exhibits exponential divergence of geodesics. This positive intrinsic curvature implies that the sphere cannot be isometrically developed onto a plane without tearing or overlapping.

Gauss-Bonnet Theorem Applications

The Gauss-Bonnet theorem provides a profound connection between the geometry and topology of surfaces, stating that for a compact orientable surface M without boundary, the integral of the Gaussian curvature K over the surface equals $2\pi times the Euler characteristic \chi(M): \int_M K \, dA = 2\pi \chi(M). This formulation, originally due to Gauss and later generalized by Bonnet, applies directly to the sphere S^2, which is a closed surface with \chi(S^2) = 2. For the standard 2-sphere of radius R embedded in \mathbb{R}^3, the Gaussian curvature is constant and equal to K = 1/R^2. The surface area of S^2 is $4\pi R^2, so the total curvature integral is \int_{S^2} K \, dA = \frac{1}{R^2} \cdot 4\pi R^2 = 4\pi, which matches $2\pi \chi(S^2) = 4\pi. This result underscores the sphere's positive curvature and topological simplicity, distinguishing it from surfaces of negative curvature like the hyperbolic plane. A key application of the theorem arises in computing the area of regions on the bounded by geodesics, such as spherical polygons. For a geodesic polygon P on the unit sphere (where R=1 and K=1), the boundary consists of great circle arcs with zero geodesic curvature \kappa_g = 0. The local form of the theorem for such a region with Euler characteristic \chi(P) = 1 (topologically a disk) simplifies to \int_P K \, dA + \sum_{i} \theta_i = 2\pi, where \theta_i are the exterior angles at the vertices. Since \int_P K \, dA equals the area A(P) on the unit sphere, and the sum of exterior angles is \sum (\pi - \alpha_i) for interior angles \alpha_i, this yields the area formula A(P) = \sum \alpha_i - (n-2)\pi, with n vertices. For a spherical triangle (n=3), the area is the angular excess A = \alpha + \beta + \gamma - \pi, known as Girard's theorem, which quantifies how the positive curvature causes the angle sum to exceed \pi. This "defect" in spherical triangles contrasts with the zero excess on the Euclidean plane and negative excess on hyperbolic surfaces, providing a direct measure of curvature's global effect. A sketch of the proof for the closed surface case relies on parallel transport. Consider a closed curve on the sphere; transporting a tangent vector parallelly around it results in a rotation by the enclosed Gaussian curvature integral, via the holonomy of the Levi-Civita connection. Triangulating the surface and summing these local rotations over a cell decomposition yields the total turning angle $2\pi \chi(M), equating to \int_M K \, dA. This approach, leveraging the sphere's embedding and the Gauss map to S^2, ties local curvature to global topology through vector field indices.

Riemannian Metric

The Riemannian metric on the sphere S^2 of radius r is the metric tensor induced by its isometric embedding as a hypersurface in the Euclidean space \mathbb{R}^3 equipped with the standard flat metric. This induced metric provides a way to measure lengths of curves, angles between tangent vectors, and volumes on the sphere, forming the foundation for its differential geometry. To derive the explicit form, parametrize the sphere using spherical coordinates (\theta, \phi), where \theta \in (0, \pi) is the polar angle and \phi \in [0, 2\pi) is the azimuthal angle, via the embedding map \mathbf{r}(\theta, \phi) = (r \sin\theta \cos\phi, r \sin\theta \sin\phi, r \cos\theta). The induced metric is the pullback of the Euclidean metric ds^2 = dx^2 + dy^2 + dz^2, computed from the first fundamental form using the partial derivatives: \mathbf{r}_\theta = (r \cos\theta \cos\phi, r \cos\theta \sin\phi, -r \sin\theta), \quad \mathbf{r}_\phi = (-r \sin\theta \sin\phi, r \sin\theta \cos\phi, 0). The metric components are then g_{\theta\theta} = \langle \mathbf{r}_\theta, \mathbf{r}_\theta \rangle = r^2, g_{\phi\phi} = \langle \mathbf{r}_\phi, \mathbf{r}_\phi \rangle = r^2 \sin^2\theta, and g_{\theta\phi} = \langle \mathbf{r}_\theta, \mathbf{r}_\phi \rangle = 0, yielding the line element ds^2 = r^2 (d\theta^2 + \sin^2\theta \, d\phi^2). This metric tensor is diagonal, reflecting the orthogonality of the coordinate basis vectors \mathbf{r}_\theta and \mathbf{r}_\phi. The spherical coordinates (\theta, \phi) form an orthogonal and complete chart on the sphere, covering the entire manifold except for the poles at \theta = 0 and \theta = \pi, where the coordinate system degenerates due to the vanishing of the \phi-direction. To study geodesics and other geometric objects using this metric, the Levi-Civita connection is required, whose Christoffel symbols \Gamma^k_{ij} are computed from the metric and its derivatives. For the sphere, the non-vanishing symbols include, for example, \Gamma^\theta_{\phi\phi} = -\sin\theta \cos\theta (arising from terms like \partial_\theta g_{\phi\phi} = 2r^2 \sin\theta \cos\theta) and \Gamma^\phi_{\theta\phi} = \Gamma^\phi_{\phi\theta} = \cot\theta. These symbols encode the intrinsic geometry of the sphere, from which its constant positive curvature can be derived.

Topology of the Sphere

Topological Invariants

The 2-sphere, denoted S^2, is a compact, connected 2-dimensional manifold without boundary, serving as the prototypical example of a closed orientable surface in topology. As a manifold, S^2 is locally Euclidean, meaning every point has a neighborhood homeomorphic to an open disk in \mathbb{R}^2, and its compactness ensures it is covered by finitely many such charts. This structure distinguishes S^2 from non-compact surfaces like the plane or infinite cylinders. A fundamental topological invariant of S^2 is its Euler characteristic, defined for a cell complex as \chi = V - E + F, where V, E, and F are the numbers of vertices, edges, and faces, respectively. For S^2, \chi(S^2) = 2, a result invariant under homeomorphisms and computable via any triangulation; for instance, the boundary of a tetrahedron yields V=4, E=6, F=4, confirming $4 - 6 + 4 = 2. This value arises equivalently from the homology groups of S^2, where \chi = \sum (-1)^n \rank H_n(S^2; \mathbb{Z}), with H_0(S^2) \cong \mathbb{Z}, H_2(S^2) \cong \mathbb{Z}, and H_n(S^2) = 0 otherwise. The Euler characteristic thus classifies S^2 among closed surfaces, as it decreases with increasing genus. Regarding orientability, S^2 is an orientable surface, meaning it admits a consistent choice of orientation across its entirety and contains no orientation-reversing closed paths, such as those in a ; it is also simply connected and two-sided, allowing a global distinction between "inside" and "outside" topologically. This property is detected by the top-dimensional homology group H_2(S^2; \mathbb{Z}) \cong \mathbb{Z}, which is free abelian of rank one. In the classification of compact connected surfaces, S^2 is unique up to homeomorphism as the only closed orientable 2-manifold with Euler characteristic 2 and genus 0, distinguishing it from higher-genus surfaces like the torus (\chi = 0) or non-orientable ones like the real projective plane (\chi = 1). This uniqueness follows from the theorem that every such surface is homeomorphic to a sphere with handles or cross-caps attached, with S^2 requiring neither.

Compactness and Manifold Properties

The n-sphere S^n is equipped with a smooth manifold structure via a standard atlas consisting of two charts obtained through stereographic projections from the north and south poles. Specifically, the projection from the north pole maps S^n minus that point to \mathbb{R}^n, and similarly for the south pole, providing a covering of the sphere. The transition maps between these charts are given by the composition of inverse projections, which are rational functions that turn out to be smooth diffeomorphisms and, moreover, conformal, preserving angles locally. This atlas establishes S^n as a smooth manifold for all n \geq 1. As a topological space, S^n is Hausdorff, meaning distinct points can be separated by disjoint open neighborhoods, and second-countable, possessing a countable basis for its topology. These properties ensure that S^n is , a condition that guarantees the existence of subordinate to any open cover. are essential tools in , enabling the construction of smooth bump functions and approximations on the manifold. The compactness of S^n further reinforces these features, as compact spaces are automatically normal and . The sphere embeds smoothly into Euclidean space in a minimal dimension via the Whitney embedding theorem, which states that any smooth n-manifold embeds in \mathbb{R}^{2n}, but S^n achieves the tighter bound of \mathbb{R}^{n+1} through its canonical embedding as the boundary of the unit ball. This embedding is smooth and realizes the sphere as a hypersurface, highlighting its global embeddability without self-intersections. For S^2 in particular, this is the standard round embedding in \mathbb{R}^3. The tangent bundle TS^n over the sphere is orientable but non-trivial in general. It is trivializable—meaning isomorphic to the product bundle S^n \times \mathbb{R}^n—only for n=1,3,7, corresponding to the existence of nowhere-vanishing parallelizable vector fields in those dimensions. For even dimensions, such as S^2, the tangent bundle is non-trivial; this follows from the , which asserts the impossibility of a continuous nowhere-zero vector field on S^2, precluding triviality. The Euler characteristic of S^2, equal to 2, underscores this non-triviality as a topological invariant.

Homotopy and Fundamental Group

The 2-sphere S^2 is simply connected, meaning it is path-connected and every closed loop on S^2 is continuously deformable, or nullhomotopic, to a constant loop at any basepoint. This property implies that the fundamental group \pi_1(S^2) is the trivial group \{0\}. To see this, S^2 can be expressed as the union of two open hemispheres, each contractible and thus with trivial fundamental group, whose intersection is an open equatorial band homeomorphic to S^1 \times \mathbb{R}, with \pi_1(S^1 \times \mathbb{R}) \cong \mathbb{Z}. However, the inclusion maps from the intersection into each hemisphere induce the trivial homomorphism on fundamental groups, since the hemispheres are contractible. Consequently, by the , \pi_1(S^2) is the amalgamated free product of two trivial groups over \mathbb{Z}, which is trivial, establishing \pi_1(S^2) = 0. The higher homotopy groups of S^2 reveal more about its topological structure, with \pi_n(S^2) = 0 for n < 2, reflecting the absence of lower-dimensional holes beyond path-connectedness. For n = 2, \pi_2(S^2) \cong \mathbb{Z}, generated by the identity map S^2 \to S^2, which classifies maps up to by their degree, an integer invariant measuring winding around the sphere. Higher groups are nontrivial and more intricate: \pi_3(S^2) \cong \mathbb{Z}, generated by the Hopf map, a fibration S^1 \to S^3 \to S^2 that links fibers topologically and demonstrates infinite-order elements in \pi_3(S^2). For n > 3, the groups are finite except in specific cases, such as \pi_4(S^2) \cong \mathbb{Z}/2\mathbb{Z} and \pi_5(S^2) \cong \mathbb{Z}/2\mathbb{Z}, arising from suspensions of the Hopf map and stable homotopy phenomena. The triviality of \pi_1(S^2) has significant applications in . Since S^2 is simply connected, it admits no nontrivial connected spaces; its cover is S^2 itself, as any covering would correspond to a of the trivial . This simply connectedness also underpins proofs of the Brouwer fixed-point theorem in two dimensions, where the absence of a continuous retraction from the closed 2-disk to its boundary S^1 (which would induce a nontrivial element in \pi_1(S^1)) ensures every continuous self-map of the disk has a fixed point, with S^2 providing the in higher-dimensional analogs.

Curves on the Sphere

Loxodromes and Rhumb Lines

A loxodrome, also known as a , is a on the surface of a sphere that maintains a constant , or bearing, relative to the north direction, crossing all meridians of longitude at the same angle. These curves spiral toward the poles without ever reaching them, forming a type of spherical spiral distinct from the shortest paths known as geodesics. The mathematical relation between longitude \phi and latitude \theta for a loxodrome with constant bearing angle \alpha (the angle between the curve and the meridians) can be expressed as \Delta\phi = \tan \alpha \cdot \ln \left( \frac{\tan(\pi/4 + \theta_2/2)}{\tan(\pi/4 + \theta_1/2)} \right), where \Delta\phi is the change in longitude and \theta_1, \theta_2 are the initial and final latitudes. This integrated form arises from the differential equation d\phi = \tan \alpha \cdot \sec \theta \, d\theta, reflecting the constant angular intersection with meridians. Key properties of loxodromes include their intersection with every at the fixed \alpha, which enables constant compass headings in . However, due to this spiraling nature, loxodromes have infinite length as they approach the poles, making them longer than the corresponding distances between points. For example, the distance between two points exceeds the , with the ratio increasing for larger separations in latitude and longitude. Historically, loxodromes were first described by Portuguese mathematician in 1537, who recognized their importance for maintaining steady directions at sea. The concept gained practical significance through Gerardus Mercator's 1569 world map , which transforms loxodromes into straight lines, facilitating compass-based by allowing sailors to plot constant-bearing courses as simple linear paths on the chart. This 's conformal property preserves angles, ensuring the constant meridian-crossing angle is maintained.

Clelia Curves

Clelia curves, also known as Clélie curves, are a family of spirals on the surface of a sphere defined by the linear relationship between the azimuthal angle () \phi and the polar angle () \theta, specifically \phi = k \theta for a constant ratio k > 0. In form using spherical coordinates on the unit sphere, a point on the curve can be expressed as: \mathbf{r}(t) = (\sin(k t) \sin t, \cos(k t) \sin t, \cos t), where t \in [0, \pi] parameterizes the \theta = t and \phi = k t. This parameterization arises from the constant ratio, producing closed curves when k is rational and open spirals otherwise. These curves spiral continuously from one pole (\theta = 0, where \phi = 0) to the opposite pole (\theta = \pi), making k/2 complete turns around the polar axis along the way, with the number of windings determined by the value of k. For integer k, the curve exhibits rotational symmetry and traces distinct loops or petals resembling floral patterns when projected onto a plane, such as in polar orthographic views. When k = 1, the Clelia curve forms a specific case known as if intersected appropriately, but in general, Clelia curves maintain a uniform angular progression orthogonal to meridians in their proportional advancement. They differ from loxodromes, which maintain a constant angle with meridians, by instead enforcing proportional angular coordinates. Geometrically, a Clelia curve can be constructed as the intersection of the sphere with a generalized Plücker conoid, a defined by lines joining points on a circle in the equatorial plane to points on the polar axis in a linearly twisted manner. The Plücker conoid for parameter n = k generates the linear dependency in spherical coordinates, ensuring the curve lies on both surfaces. This construction highlights the Clelia curve's algebraic nature when k is rational, with degree $2(p + q) where k = p/q in lowest terms. In applications, Clelia curves model the ground tracks of satellites in circular polar orbits, approximating the paths traced on Earth's surface as the satellite completes multiple revolutions while the Earth rotates. For an orbital period T hours, the ratio k = 24/T (Earth's sidereal day) yields the curve's winding, closing after rational multiples of orbits and rotations, useful for analyzing coverage in remote sensing or reconnaissance missions.

Spherical Conics and Other Curves

Spherical conics generalize planar conic sections to the surface of a sphere, defined as the intersection of the sphere with a quadratic cone sharing the same center as the sphere. This intersection yields a quartic curve on the sphere, which can manifest as a spherical ellipse or hyperbola depending on the cone's parameters. A spherical ellipse specifically arises when the cone produces a closed curve analogous to a planar ellipse, characterized by two foci on the sphere. The defining property of a spherical ellipse mirrors that of its planar counterpart but uses geodesic distances: it is the locus of points on the sphere where the sum of the spherical distances to two fixed foci is constant. Equivalently, it can be constructed as the intersection of the sphere with an elliptic , where the cone's is at the sphere's and the ellipse's foci correspond to points where tangent cones from the foci touch the sphere. An adapted focus-directrix property holds, where the ratio of the sine of the geodesic distance to a directrix (a small , the polar of the ) over the sine of the distance to the focus equals the , a constant less than 1 for ellipses. Spherical ellipses are bounded regions on the sphere, often lying between two small circles parallel to the in aligned coordinate systems, reflecting their closed and compact . Beyond ellipses, spherical conics include hyperbolas, defined similarly by a constant difference of distances to foci, forming unbounded branches on the sphere. In the of high , certain spherical conics approach Clelia curves, which are spherical roses generated by rotating a point around an at constant angular speed. Other notable curves on the sphere include , the intersection of the sphere with a of half the sphere's whose passes through the center of the sphere. This produces a figure-eight-shaped space , symmetric about the cylinder's . The spherical , a of constant tangent length analogous to the planar , arises as the polar curve of a loxodrome under absolute polarity on the sphere, exhibiting pseudospherical properties when revolved. Parametrizing these curves poses challenges in spherical coordinates (\theta, \phi), as the resulting equations are generally transcendental due to the trigonometric nature of the coordinates, despite the curves being algebraic (degree 4) in Cartesian coordinates. For instance, the spherical ellipse equation in Cartesian form is x^2/a^2 + y^2/b^2 - z^2/c^2 = 0 intersected with x^2 + y^2 + z^2 = 1, but converting to spherical coordinates yields non-algebraic relations involving sines and cosines of multiple angles. Numerical or methods, such as using the bifocal property, are often employed for computation.

Sphere Intersections with Planes and Lines

The intersection of a sphere of radius r centered at point O with a plane depends on the perpendicular distance d from O to the plane. If d < r, the intersection is a circle in the plane with radius \sqrt{r^2 - d^2}. If d = r, the intersection degenerates to a single point of tangency. If d > r, the plane does not intersect the sphere, yielding the empty set. The of a sphere with a straight line is determined by parameterizing the line as \mathbf{P} = \mathbf{P_1} + t (\mathbf{P_2} - \mathbf{P_1}) and substituting into the sphere equation |\mathbf{X} - \mathbf{O}|^2 = r^2, which produces a at^2 + bt + c = 0 in the parameter t. The nature of the intersection follows from the \Delta = b^2 - 4ac: two distinct points if \Delta > 0, a single point (tangency) if \Delta = 0, and no real intersection if \Delta < 0. The coefficients are a = |\mathbf{P_2} - \mathbf{P_1}|^2, b = 2(\mathbf{P_1} - \mathbf{O}) \cdot (\mathbf{P_2} - \mathbf{P_1}), and c = |\mathbf{P_1} - \mathbf{O}|^2 - r^2. Tangency occurs when the perpendicular distance from the sphere's center O to the line equals r, corresponding to \Delta = 0. A related concept is the power of a point P with respect to the sphere, defined as |OP|^2 - r^2. For any line through P intersecting the sphere at points A and B, the product of the directed distances PA \cdot PB equals this power (positive if P is outside, negative if inside, and zero if on the sphere). This property, analogous to the planar case for , aids in analyzing intersections without solving the quadratic explicitly. For example, consider a sphere centered at the origin. The plane z = 0 (distance d = 0) intersects it in the equator, a circle of radius r. The line along the z-axis intersects the sphere at the north and south poles, (0,0,r) and (0,0,-r). Such intersections yield small circles for $0 < d < r and great circles when d = 0.

Intersections with Quadric Surfaces

The intersection of a sphere with another quadric surface, both defined by quadratic equations in three-dimensional space, generally yields a space curve of degree 4. This arises because the resultant of two quadratic polynomials eliminates one variable, producing a degree-4 equation in the remaining coordinates, describing an algebraic curve that may consist of one or two branches, potentially non-planar. In projective space \mathbb{P}^3, Bézout's theorem confirms that the intersection of two quadric hypersurfaces has degree 4, counting multiplicities, and a transversal line through the space intersects this curve in up to 4 points. Such curves are fundamental in algebraic geometry and computer-aided design, where robust parametrization is essential to handle their topology. A prominent example is the intersection of a sphere with a cylinder, both quadric surfaces. When the cylinder of radius a is internally tangent to a sphere of radius $2a centered at the origin, with the cylinder's axis along the x-axis and offset such that it passes through the sphere's diameter, the intersection forms Viviani's curve—a figure-eight-shaped space curve symmetric about the xy-plane. This curve, studied by Vincenzo Viviani in 1690, has parametric equations x = a(1 + \cos t), y = a \sin t, z = 2a \sin(t/2) for t \in [-\pi, \pi], and lies on the cylinder (x - a)^2 + y^2 = a^2 and sphere x^2 + y^2 + z^2 = (2a)^2. In other configurations, such as when the cylinder's axis is perpendicular to the line joining the centers and the distances allow disjoint or nested positions, the intersection degenerates into two circles, each a conic section of degree 2, totaling the degree-4 curve. Intersections with cones or hyperboloids also produce notable curves, particularly spherical conics when the quadric's vertex aligns with the sphere's center. A quadratic cone with vertex at the origin intersecting a unit sphere yields a spherical conic, such as a spherical ellipse, defined as the set of points on the sphere where the sum of geodesic distances to two foci equals a constant $2b < \pi. These curves are the spherical analogs of plane conics, projected onto the sphere via the cone's rulings, and appear as closed loops or branches depending on the cone's aperture; for instance, a right circular cone generates symmetric spherical ellipses. Hyperboloid intersections follow similarly, often resulting in hyperbolic spherical branches when the hyperboloid is one-sheeted and coaxial with the sphere.

General Surface Intersections

The intersection of a sphere with an arbitrary algebraic surface of degree n produces a space curve of degree up to $2n, as the sphere itself is a quadric surface of degree 2, and the degree of the intersection curve for two algebraic surfaces is bounded by the product of their degrees. This algebraic structure allows for exact representations in computational geometry, though singularities or multiple components may reduce the effective degree in specific configurations. For non-algebraic or complex surfaces, numerical methods are essential to approximate and visualize intersections. The , originally developed for isosurface extraction from volumetric data, can be adapted to compute intersection curves by sampling the region between surfaces and triangulating the boundary where they meet, enabling high-resolution 3D reconstructions. In computer graphics, iteratively advances rays through implicit surface representations to find intersection points, providing robust handling of arbitrary topologies without explicit meshing. A classic analytic case within this framework is the , solved via a quadratic equation to yield up to two real roots representing entry and exit points, which serves as a foundational primitive for broader ray-surface algorithms. Topologically, the intersection curve inherits properties from the parent surfaces, with its genus determined by algebraic invariants for smooth complete intersections. For a (degree 2) intersecting a modeled as a degree-4 surface, the resulting curve is a complete intersection of type (2,4) with genus g = \frac{2 \cdot 4 \cdot (2 + 4 - 4)}{2} + 1 = 9, reflecting a highly complex embedding that can feature multiple loops and self-intersections akin to higher-genus structures. This genus formula, g = \frac{d_1 d_2 (d_1 + d_2 - 4)}{2} + 1 for degrees d_1 and d_2 in \mathbb{P}^3, underscores how sphere intersections can yield curves of arbitrary topological complexity depending on the intersecting surface. Such intersections find critical applications in computer-aided design (CAD), where algorithms compute boundary curves between trimmed surfaces to construct solid models and ensure watertight assemblies. In ray tracing for realistic rendering, sphere-surface intersections enable efficient visibility computations and shadow generation, with post-2020 advancements in GPU hardware—such as NVIDIA's RT Cores and extended ray-traversal units—accelerating these operations by orders of magnitude through dedicated intersection pipelines and hierarchical bounding. These techniques, exemplified in quadric-sphere overlaps, extend seamlessly to general surfaces for interactive simulations.

Generalizations of the Sphere

Ellipsoids and Superquadrics

An ellipsoid is a quadric surface defined by the equation \mathbf{x}^T A \mathbf{x} = 1, where \mathbf{x} = (x, y, z)^T is a point in three-dimensional space and A is a symmetric positive definite matrix. The positive definiteness of A ensures that the surface is bounded and closed, forming an oval-shaped solid without self-intersections. The principal axes of the ellipsoid align with the eigenvectors of A, which correspond to the directions of the semi-axes lengths determined by the reciprocals of the square roots of the eigenvalues. Unlike the sphere, an ellipsoid lacks constant Gaussian curvature, with the curvature varying across its surface depending on the position relative to the principal axes. For an ellipsoid with semi-axes lengths a, b, and c along the principal directions, the volume enclosed by the surface is \frac{4}{3} \pi a b c. The sphere emerges as a special isotropic case of the ellipsoid when a = b = c. Superquadrics generalize ellipsoids by raising the terms in the defining equation to a power p \neq 2, yielding the implicit form \left( \left| \frac{x}{a} \right|^p + \left| \frac{y}{b} \right|^p + \left| \frac{z}{c} \right|^p \right)^{2/p} = 1. This family of surfaces, introduced by , allows for a continuous range of shapes from pinched forms (when p > 2) to more concave or star-like appearances (when $0 < p < 2). Like ellipsoids, superquadrics do not possess constant curvature, with the surface geometry distorting based on the value of p and the scaling parameters a, b, and c. Setting p = 2 recovers the standard ellipsoid equation, providing a smooth transition between these generalized shapes.

n-Dimensional Spheres

In higher-dimensional Euclidean space, the , denoted S^n, is defined as the set of points \mathbf{x} \in \mathbb{R}^{n+1} such that \|\mathbf{x}\| = r, where r > 0 is the radius and \|\cdot\| denotes the Euclidean norm. This generalizes the familiar 2-sphere in three dimensions, which corresponds to the ordinary sphere. The surface measure (or "surface area") of the of radius r is given by \sigma_n(r) = \frac{2 \pi^{(n+1)/2} r^n}{\Gamma\left(\frac{n+1}{2}\right)}, where \Gamma is the . This formula arises from integrating over the in \mathbb{R}^{n+1}, scaling with r^n due to the dimensionality of the manifold. The n-ball, denoted B^n, is the solid region enclosed by S^{n-1}, consisting of all points in \mathbb{R}^n with \|\mathbf{x}\| \leq r. Its volume is V_n(r) = \frac{\pi^{n/2} r^n}{\Gamma\left(\frac{n}{2} + 1\right)}. This volume can be derived recursively through integration: the volume of the n-ball of radius R satisfies V_n(R) = \int_0^R \sigma_{n-1}(r) \, dr, leading to the relation V_n(r) = \frac{\sigma_{n-1}(r) r}{n}, which connects the interior volume to the boundary surface measure. The n-sphere equipped with the induced Riemannian metric from \mathbb{R}^{n+1} has constant K = 1/r^2, independent of the dimension n; this follows from the fact that all 2-planes in the at any point are congruent under the , yielding the same as in the 2-sphere case. For example, the 1-sphere S^1 is the circle in the , while the 3-sphere S^3 in \mathbb{R}^4 admits a natural parametrization using unit quaternions, which form a under multiplication and double-cover the rotation group SO(3).

Spheres in Metric and Normed Spaces

In a X equipped with a \|\cdot\|, the is defined as the set S_X = \{ x \in X \mid \|x\| = 1 \}. This set generalizes the but lacks the and smoothness characteristic of the \ell_2 case, as the depends on the specific ; for instance, the of the may have flat faces or sharp corners. In a general (M, d), a sphere centered at c \in M with radius r > 0 is the set \{ x \in M \mid d(c, x) = r \}. Unlike spheres in spaces, these sets need not be compact, connected, or even closed, nor do they possess differentiable structure, as the metric may induce pathologies such as non-Hausdorff topologies or unbounded curvature in discrete spaces. In Hilbert spaces, which are complete normed spaces arising from inner products, the unit sphere becomes infinite-dimensional when the ambient space is, enabling applications in such as the representation of solutions to partial differential equations (PDEs). For example, in Sobolev spaces over domains, the unit sphere facilitates the study of weak solutions to elliptic PDEs via trace operators and embedding theorems, ensuring in appropriate topologies for existence and regularity results. Prominent examples arise in sequence spaces like \ell_p for $1 \leq p < \infty, where the unit sphere distorts relative to the round Euclidean sphere (p=2); for p=1, the taxicab or Manhattan metric yields a unit sphere in \mathbb{R}^2 that forms a diamond (square rotated 45 degrees) with vertices at (\pm 1, 0) and (0, \pm 1), highlighting how non- norms alter geometric intuition.

Historical Development

Ancient and Classical Contributions

The concept of the sphere emerged in thought as a symbol of perfection and cosmic order, with early contributions rooted in philosophical and astronomical speculations. Pre-Socratic philosopher (ca. 570–495 BCE) is credited with introducing the "harmony of the spheres," positing that the celestial bodies—envisioned as spheres—moved in harmonious ratios, producing an inaudible music reflective of mathematical proportions underlying the . This idea integrated with cosmology, viewing the sphere as the most harmonious form due to its uniformity and . Basic properties of the sphere, such as its roundness and equidistance from a center, were intuitively recognized by early observers through natural phenomena like celestial bodies and rolling objects. In the Classical period, (ca. 428–348 BCE) elevated the sphere to an ideal geometric form in his dialogue Timaeus, describing the as a living being shaped like a sphere by the , the most perfect and self-sufficient figure enclosing all directions equally from its center. 's cosmology emphasized the sphere's perfection, associating it with the elemental body of fire and the eternal motion of the heavens. (384–322 BCE), building on Platonic ideas but critiquing their details, developed a in works like , where the universe consists of nested concentric spheres made of , with the outermost sphere of imparting uniform circular motion to the . argued that the sphere's natural motion is rotation, distinguishing the sublunary realm of change from the immutable , thus grounding in physical principles. Hellenistic mathematicians advanced rigorous geometric treatments of the sphere. Archimedes of Syracuse (ca. 287–212 BCE) provided the first precise calculations of the sphere's volume and surface area in his treatise On the Sphere and Cylinder, employing the method of exhaustion—a precursor to integral calculus—to prove that the surface area of a sphere equals four times that of its great circle and its volume is two-thirds that of the circumscribing cylinder. These results, derived by approximating the sphere with inscribed and circumscribed polyhedra and taking limits, demonstrated the sphere's proportional relations to other solids and were so significant that Archimedes requested a sphere and cylinder be depicted on his tombstone.

Medieval Islamic Contributions

During the (8th–14th centuries), mathematicians built upon Greek foundations, making substantial advances in essential for astronomy and navigation. Abu Abd Allah ibn Muʿādh al-Jayyānī (989–1079 ) authored The Book of Unknown Arcs of a Sphere, the first comprehensive treatise on , providing solutions for right-angled spherical triangles and introducing formulas for spherical excess. Al-Bīrūnī (973–1048 ) further developed methods for measuring the Earth's using spherical models and contributed to the understanding of great circles and geodesics on spheres. These works preserved and extended Hellenistic knowledge, influencing later European through translations.

Renaissance to Modern Mathematics

The Renaissance marked a pivotal shift in the mathematical treatment of the sphere, transitioning from synthetic geometry to analytic methods. In 1637, René Descartes introduced analytic geometry in his appendix La Géométrie to Discours de la méthode, providing a framework to represent geometric objects algebraically. This approach enabled the equation of a sphere of radius r centered at the origin as x^2 + y^2 + z^2 = r^2, extending coordinate-based descriptions from plane curves to three-dimensional surfaces. The development of calculus by Isaac Newton and Gottfried Wilhelm Leibniz in the late 17th century further advanced the quantitative analysis of spheres. Newton employed his method of fluxions to compute the volume of a sphere by integrating cross-sectional areas, deriving V = \frac{4}{3} \pi r^3 through limits of cylindrical approximations. Similarly, Leibniz's infinitesimal calculus facilitated parametric integrals for surface area, yielding A = 4 \pi r^2 via summation of ring-like elements along meridians, emphasizing the sphere's rotational symmetry. In the early 19th century, revolutionized the study of spheres through . In his 1827 Disquisitiones generales circa superficies curvas, Gauss introduced the K, defined for a surface as the product of principal curvatures. For a , K = 1, constant and positive, reflecting its intrinsic geometry independent of embedding. His proved that this curvature is an intrinsic invariant, measurable solely from the surface's metric without reference to ambient space, as it depends only on the . The 20th century integrated spheres into and physics. , in his 1904 paper "Cinquièmme complément à l'analyse situs," constructed the , a with the of the but a non-trivial , highlighting non-trivial topological structures mimicking spherical properties. In , Albert Einstein's 1915 field equations described , where spatial hyperspheres (n-spheres for n \geq 3) model closed universes, with the Friedmann-Lemaître-Robertson-Walker metric incorporating spherical symmetry for homogeneous cosmologies.

References

  1. [1]
    Sphere -- from Wolfram MathWorld
    A sphere is defined as the set of all points in three-dimensional Euclidean space R^3 that are located at a distance r (the "radius") from a given point ...
  2. [2]
    SPHERE Definition & Meaning - Merriam-Webster
    1. a (1) : the apparent surface of the heavens of which half forms the dome of the visible sky (2) : any of the concentric and eccentric revolving spherical ...
  3. [3]
    Sphere - BYJU'S
    A sphere is a three-dimensional object that is round in shape. The sphere is defined in three axes, i.e., x-axis, y-axis and z-axis.Equation Of Sphere · Area & Volume Of Sphere · Volume of Hemisphere
  4. [4]
    Sphere - Shape, Definition, Formulas, Properties, Examples - Cuemath
    A sphere is a three-dimensional round-shaped object. Unlike other three-dimensional shapes, a sphere does not have any vertices or edges. All the points on its ...
  5. [5]
    Sphere Definition | GIS Dictionary - Esri Support
    A three-dimensional solid in which all points on the surface are the same distance from the center; used to approximate the true shape and size of the earth.
  6. [6]
    Euclid's Elements, Book XI, Definitions 14 through 17
    Definition 17. A diameter of the sphere is any straight line drawn through the center and terminated in both directions by the surface of the sphere.
  7. [7]
    Ball -- from Wolfram MathWorld
    (Although physicists often use the term "sphere" to mean the solid ball, mathematicians definitely do not!) ... Sphere Packings, Lattices, and Groups, 2nd ed.
  8. [8]
    sphere in nLab
    Sep 30, 2025 · Definition 1.1. The n n -dimensional unit sphere , or simply n n -sphere, is the topological space given by the subset of the ( n + 1 ) ...Missing: notation | Show results with:notation
  9. [9]
    Mathematical Notation: Past and Future (2000) - Stephen Wolfram
    Stephen Wolfram on mathematical notation's development from antiquity through Leibniz, Euler, Peano, & modern times, & how it is like human language.
  10. [10]
    Spheres and Ellipsoids
    from the origin. Suppose we want to know what the equation of a sphere is. We can use the distance formula above to help. If the sphere is centered at ...
  11. [11]
    14.1 The Coordinate System
    Now we can get the similar equation r2=(x−h)2+(y−k)2+(z−l)2, which describes all points (x,y,z) at distance r from (h,k,l), namely, the sphere with radius r and ...
  12. [12]
    Calculus III - Parametric Surfaces - Pauls Online Math Notes
    Mar 25, 2024 · In spherical coordinates we know that the equation of a sphere of radius a a is given by,. ρ=a ρ = a. and so the equation of this sphere (in ...
  13. [13]
    [PDF] 7.2 | Calculus of Parametric Curves
    Figure 7.26 A semicircle generated by parametric equations. When this curve is revolved around the x-axis, it generates a sphere of radius r. To calculate the ...
  14. [14]
    16.6 Parametric Surfaces - Courses
    A parametric sphere surface with the geogebra commands to create the plot displayed on the left ... surface area is found with the integral. A(S)=∬D|ru×rv ...
  15. [15]
    Properties of Velocity and Speed - Ximera - The Ohio State University
    Thus, for any parametrization of a curve which lies on a sphere, the velocity vector will always be perpendicular to the position vector.
  16. [16]
    2.7 Cylindrical and Spherical Coordinates - Calculus Volume 3
    Mar 30, 2016 · A sphere that has Cartesian equation x2+y2+z2=c2 has the simple equation ρ=c in spherical coordinates. In geography, latitude and longitude are ...
  17. [17]
    Coordinate Transformation on a Sphere Using Conformal Mapping in
    A problem with the spherical coordinate system is related to the singularities at the North and South Poles.
  18. [18]
    Quadric Surface - an overview | ScienceDirect Topics
    A quadric surface is defined by a homogeneous quadratic equation F(x, y, z, w) = 0, represented in matrix form as X^T M X = 0, where X is the column vector ...
  19. [19]
    [PDF] Quadric Surfaces
    The general implicit form for a 3D quadric surface can be written in homoge- neous x, y, z, w coordinates as: ax2 +2bxy +2cxz +2dxw +ey2 +2fyz +2gyw +hz2 ...
  20. [20]
    [PDF] Archimedes' Determination of Circular Area - Mathematics
    Mar 24, 2006 · ▫ Can restate proof as the surface area of a sphere is equal to ... ○ Volume (sphere) = 4(1/3πr3) = 4/3πr3. Page 36. Archimedes ...<|control11|><|separator|>
  21. [21]
    Calculus III - Surface Integrals - Pauls Online Math Notes
    Nov 28, 2022 · In this section we introduce the idea of a surface integral. With surface integrals we will be integrating over the surface of a solid.
  22. [22]
    [PDF] Cavalieri's determination of the volume of a sphere - People
    Now we give the diagram that yields a formula for the volume of the sphere. The figure on the left is the sphere. The figure on the right is a cylinder with two ...
  23. [23]
    [PDF] Triple Integrals for Volumes of Some Classic Shapes
    A Sphere. The equation for the outer edge of a sphere of radius a is given by x2 + y2 + z2 = a2. If we want to consider the volume inside, then we are ...
  24. [24]
    [PDF] VOLUMES OF n-DIMENSIONAL SPHERES AND ELLIPSOIDS
    This paper explores the volume of n-dimensional spheres, under different p-norms, and uses linear transformation to find the volume of an n-dimensional ellipse.
  25. [25]
    Great Circle -- from Wolfram MathWorld
    A great circle is a section of a sphere that contains a diameter of the sphere ( ... sphere, also known as an orthodrome, is a segment of a great circle ...Missing: definition | Show results with:definition
  26. [26]
    Jung's Theorem -- from Wolfram MathWorld
    Jung's theorem states that the generalized diameter D of a compact set X in R^n satisfies D>=Rsqrt((2(n+1))/n), where R is the circumradius of X.Missing: sphere | Show results with:sphere<|control11|><|separator|>
  27. [27]
    (PDF) A generalization of Jung's theorem - ResearchGate
    Aug 6, 2025 · Jung's theorem establishes a relation between circumradius and diameter of a convex body. Half of the diameter can be interpreted as the ...Missing: original | Show results with:original
  28. [28]
    Combinatorial Generalizations of Jung's Theorem
    Mar 1, 2013 · The famous theorem of Jung states that any set with diameter in can be covered by the ball of radius (see [3]). The proof of this theorem is ...
  29. [29]
    Why study matrix groups?
    SO(3) = all positions of a globe on a fixed stand. Three elements of SO(3) ... For example, the symmetry group of the sphere Sn ⊂ Rn+1 equals the group ...
  30. [30]
    [PDF] Notes 2 Lecture Notes on the Differential Geometry of Lie Groups
    For example, in the action of SO(3) on 3 , the isotropy subgroup of any nonzero vector x is the SO(2) subgroup of rotations about the axis defined by x, whereas ...
  31. [31]
    [PDF] Some notes on group theory.
    8.9 The standard irreps of SO(3). SO(3) is the symmetry group of the sphere in three dimensions. Points at the sphere are characterized by the polar angle ϑ ...
  32. [32]
    [PDF] Symmetry Groups of Platonic Solids - Brown Math Department
    Sep 17, 2007 · The purpose of this handout is to discuss the symmetry groups of Platonic solids. Let R3 denote 3-dimensional space. A rotation of R3 is any ...
  33. [33]
    [PDF] Pencils of Planes and Spheres in Problem Solving
    In addition, we discuss possible approach how to use the pencils of planes or pencil of spheres in solving some problems in the analytic geometry. ... If the ...
  34. [34]
  35. [35]
    Inversion -- from Wolfram MathWorld
    Inversion is the process of transforming points P to a corresponding set of points P^' known as their inverse points. Two points P and P^' are said to be ...<|control11|><|separator|>
  36. [36]
    [PDF] Euclidean Plane and its Relatives; a minimalist introduction.
    The inversion of the space has many properties of the inversion of the plane. Most important for us are the analogs of theorems 9.6, 9.7,. 9.25 which can be ...
  37. [37]
    [PDF] Geometry II
    Proposition 1.2. 5. Every Mِbius transformation is a composition of similarity trans- formations and inversions in the unit sphere.
  38. [38]
    [PDF] Stereographic Projection
    A geometric construction known as stereographic projection gives rise to a one-to-one cor- respondence between the complement of a chosen point A on the sphere ...
  39. [39]
    [PDF] STEREOGRAPHIC PROJECTION IS CONFORMAL Let S2 = {(x, y, z ...
    The proof that stereographic projection is conformal tacitly assumed that t and t meet. Must they? What happens to the proof if they don't? • Show that ...Missing: property | Show results with:property
  40. [40]
    stereographic projection in nLab
    Feb 15, 2024 · For p ∈ S n p \in S^n one of the corresponding poles, the stereographic projection is the map which sends a point x ∈ S n \ { p } x \in S^{n}\ ...Idea · Definition · Generalizations · Properties
  41. [41]
    [PDF] 4. The Riemann sphere and stereographic projection - People
    Stereographic projection gives us a more concrete way of identifying ˜C with a sphere, one which, moreover, yields a lot of geometrical insight. b b b. N. ˜z z.
  42. [42]
    [PDF] Stereographic Projection (the Riemann sphere)
    Riemann sphere or or conformal. The images. Curves on two stereographic projection is angle-preserving the surface of the sphere of have the same angles ...
  43. [43]
    Spherical Geometry -- from Wolfram MathWorld
    In spherical geometry, straight lines are great circles, so any two lines meet in two points. There are also no parallel lines. The angle between two lines in ...
  44. [44]
    [PDF] A Brief Survey of Elliptic Geometry - University of West Florida
    There are three fundamental branches of geometry: Euclidean, hyperbolic and elliptic, each characterized by its postulate concerning parallelism.Missing: geodesic | Show results with:geodesic
  45. [45]
    Non-Euclidean Geometry -- from Wolfram MathWorld
    Spherical geometry is a non-Euclidean two-dimensional geometry. It was not until 1868 that Beltrami proved that non-Euclidean geometries were as logically ...
  46. [46]
    How to Use History to Clarify Common COnfusions in Geometry
    However, the existence of parallel lines is false in spherical geometry and thus it must be one of the other postulates that fails on the sphere. Thus it is ...
  47. [47]
    Small Circle -- from Wolfram MathWorld
    A small circle is a spheric section that does not contain a diameter of the sphere (Kern and Bland 1948, p. 87; Tietze 1965, p. 25).
  48. [48]
    [PDF] Homework 2 Solutions
    ... sphere of radius R. Show that if you draw a circle of radius r, the circle's circumference will be. C = 2πR sin r. R . (1). Idealize the Earth as a perfect ...
  49. [49]
    Math 497
    A great circle divides a sphere into 2 hemispheres. One special property of a great circle is that any two great circles intersect in two points (on opposite ...
  50. [50]
    4. Great circle sailing
    The calculation of the great circle track between two points A and B with given latitude and longitude is an exercise in spherical trigonometry.
  51. [51]
    [PDF] Finding Geodesics on Surfaces - Stanford Computer Science
    Consider the case of a sphere. The geodesic between two points on a sphere is always a segment of a great circle. Recall that a great circle on a sphere is ...
  52. [52]
    Chapter 3: Section 7: Part 4
    Moreover, the shortest distance s from P to Q on the surface of the sphere is. s = Ra. since this is the shortest distance from P to Q along the great circle g( ...
  53. [53]
    [PDF] Distance between Points on the Earth's Surface - KSU Math
    This is an interesting exercise in spherical coordinates, and relates to the so-called haversine. Spherical coordinates z=Rsinθ y=Rcosθsin φ x=Rcosθcos φ. R z.
  54. [54]
    Computing Distances - NYU Computer Science
    Haversine Formula (from R.W. Sinnott, "Virtues of the Haversine", Sky and Telescope, vol. 68, no. 2, 1984, p. 159): dlon = lon2 - lon1 dlat = lat2 - lat1 a ...
  55. [55]
    [PDF] Spherical Trigonometry - UCLA Mathematics
    Definition 8.1 (Spherical Excess): The spherical excess of a spherical triangle is the sum of its angles minus π radians. Theorem 8.2 (Girard's Theorem): The ...
  56. [56]
    [TeX] Introduction to the "geosphere" package (Version fooVersion ...
    Bearing changes continuously when traveling along a Great Circle. The final ... They were used in navigation because it is easier to follow a constant compass ...
  57. [57]
    [PDF] Long and short-range air navigation on spherical Earth
    Jan 1, 2017 · Generally, GC calculations on spherical Earth are sufficient for reliable air navigation flight planning purposes, considering all other ...
  58. [58]
    [PDF] Principal Curvatures∗
    Dec 5, 2024 · A sphere bends the same amount in every direction. Take the unit sphere in Example 9 in the notes “Surfaces”, for instance, with the ...
  59. [59]
    [PDF] Basics of the Differential Geometry of Surfaces - CIS UPenn
    There are surfaces of constant Gaussian curvature. For example, a cylinder or a cone is a surface of Gaussian curvature K = 0. A sphere of radius R has positive.Missing: orientation | Show results with:orientation
  60. [60]
    [PDF] 7. THE GAUSS-BONNET THEOREM - Penn Math
    Mar 29, 2012 · State and prove the Gauss-Bonnet Theorem for a spherical polygon with geodesic sides. Page 32. 32. Gaussian curvature of polyhedral surfaces in ...
  61. [61]
    [PDF] the gauss-bonnet theorem
    Aug 29, 2022 · It also generalizes a number of mathematical facts in ele- mentary geometry, such as the sum of exterior angles for a polygon and the surface.
  62. [62]
    [PDF] The Gauss-Bonnet Theorem and its Applications - UC Berkeley math
    In this paper we survey some developments and new results on the proof and applications of the Gauss-Bonnet theorem. Our special emphasis is the relation of ...
  63. [63]
    [PDF] Riemannian Geometry
    The formula (1.35) is the formula for induced Riemannian metric on the surface ... We consider already an example of induced Riemannian metric on sphere in.
  64. [64]
    [PDF] Algebraic Topology - Cornell Mathematics
    This book was written to be a readable introduction to algebraic topology with rather broad coverage of the subject. The viewpoint is quite classical in ...
  65. [65]
  66. [66]
    [PDF] INTRODUCTION TO DIFFERENTIAL GEOMETRY - ETH Zürich
    ... stereographic projection. This chart is conformal (which means that it ... smooth atlas on M is a collection s of charts on M any two of which are ...
  67. [67]
    [PDF] Riemannian geometry - Imperial College London
    It could have one element or could be uncountably infinite. Definition 2.6 (Maximal atlases). A smooth atlas ... Stereographic projection is a local conformal ...
  68. [68]
    [PDF] Partitions of Unity
    Sep 25, 2013 · Corollary. A (second countable Hausdorff) manifold is paracompact. 13. Corollary. (Cr Urysohn's Lemma) Let A and B be disjoint closed subsets.
  69. [69]
    [PDF] Chapter 9 Partitions of Unity, Covering Maps ~ - CIS UPenn
    Proposition 9.3 implies that a second-countable, lo- cally compact (Hausdorff) topological space is the union of countably many compact subsets. Thus, X is ...
  70. [70]
    [PDF] lecture 9: the whitney embedding theorem
    The Whitney embedding theorem states that any smooth manifold can be embedded into a Euclidean space with dimension at least twice the manifold's dimension.
  71. [71]
    [PDF] To motivate the definition of a vector bundle let us consider tangent ...
    As we shall show in §2.2, Sn has a nonvanishing vector field iff n is odd. From this it follows that the tangent bundle of Sn is not isomorphic to the trivial ...<|control11|><|separator|>
  72. [72]
    Loxodrome -- from Wolfram MathWorld
    If the surface is a sphere, the loxodrome is a spherical spiral. The loxodrome is the path taken when a compass is kept pointing in a constant direction.
  73. [73]
    [PDF] A Comparative Analysis of Rhumb Lines and Great Circles
    May 13, 2016 · Definition 2.1. [Great Circle] A great circle is the shortest distance on Earth from one place to the next. Definition 2.2. [Meridian] A ...
  74. [74]
    [PDF] Mapping the sphere - Rice Math Department
    The immediate result of these three properties is that the Mercator projection maps rhumb lines on the sphere into straight lines, and vice versa. As a ...Missing: historical | Show results with:historical
  75. [75]
    [PDF] Abstract Shape Synthesis From Linear Combinations of Clelia Curves
    This article outlines several families of shapes that can be produced from a linear combination of Clelia curves. We present parameters required to generate ...Missing: φ) = θ
  76. [76]
    Spherical ellipse - MATHCURVE.COM
    The spherical ellipse is the locus of the points on a sphere for which the sum of the distances (taken on the sphere) to two fixed points F and F' on the sphere ...Missing: properties | Show results with:properties
  77. [77]
    The Focal Circles of Spherical Conics - jstor
    sphere inside a tangent right cone. In addition, some of the properties of the Focal Circles of Spherical Conics were derived from the same idea. It is now ...
  78. [78]
    Viviani's Curve -- from Wolfram MathWorld
    Viviani's curve is the intersection of a cylinder and a sphere, studied by Viviani in 1692. It appears as a lemniscate-like curve, circle, and parabolic ...
  79. [79]
    intersection of sphere and plane - PlanetMath
    Mar 22, 2013 · . The intersection curve of a sphere and a plane is a circle. PQ=ϱ=√r2−OQ2= constant.Missing: formula | Show results with:formula<|control11|><|separator|>
  80. [80]
    Circle, Cylinder, Sphere - Paul Bourke
    A line can intersect a sphere at one point in which case it is called a tangent. It can not intersect the sphere at all or it can intersect the sphere at two ...
  81. [81]
    Power of a Point
    The Power of a Point · Definition: The power of A with respect to c = |OA|2 - r · pc = d2 - r · pc(A) = |OA|2 - r · Notation Note: Sved uses the notation P(c) for ...<|control11|><|separator|>
  82. [82]
    Circle Power -- from Wolfram MathWorld
    "Power of a Point with Respect to a Circle." §300-303 in An Elementary Treatise on Modern Pure Geometry. London: Macmillian, pp. 183-185, 1893. Pedoe, D.<|control11|><|separator|>
  83. [83]
    [PDF] Geometric Approaches to Nonplanar Quadric Surface Intersection ...
    Two primary approaches to the representation of quadric surfaces have evolved: an algebraic one and a geometric one [4]. The algebraic approach is summarized in ...
  84. [84]
    Degree of intersection curve of two quadrics - MathOverflow
    Jun 21, 2012 · The intersection of two quadrics has always degree 4, if you count the intersection with multiplicities.Missing: Bézout's projective space<|control11|><|separator|>
  85. [85]
    Cylinder-Sphere Intersection -- from Wolfram MathWorld
    The curve formed by the intersection of a cylinder and a sphere is known as Viviani's curve. The problem of finding the lateral surface area of a cylinder ...Missing: line | Show results with:line
  86. [86]
    [PDF] Equations of the spherical conics
    We know that spherical ellipse is the intersection of the unit sphere with a quadratic cone whose vertex is at the center of the sphere [2]. Spherical ellipse ...
  87. [87]
    [PDF] Polynomial Curves and Surfaces - UT Computer Science
    Sep 8, 2010 · algebraic intersection curve of two algebraic surfaces may be as large as the product of the geometric degrees of the two surfaces [38]. The ...
  88. [88]
    [PDF] Marching cubes: A high resolution 3D surface construction algorithm
    Using the index to tell which edge the surface intersects, we can interpolate the surface intersection along the edge. We use linear interpolation, but have ...
  89. [89]
    [PDF] Numerical Methods for Ray Tracing Implicitly Defined Surfaces
    Sep 25, 2014 · Ray marching with fixed steps has three advantages over analytic ray intersections: ... Marching cubes: A high resolution 3d surface.
  90. [90]
    A Minimal Ray-Tracer: Ray-Sphere Intersection - Scratchapixel
    The geometric solution to the ray-sphere intersection test relies on simple math, mainly geometry, trigonometry, and the Pythagorean theorem.
  91. [91]
    Genus of intersection of two surfaces in $\mathbb{P}^3
    Oct 23, 2014 · As a sanity check, if say d1=1, then C is actually a plane curve of degree d2, and the formula above gives its genus as 12(d2(d2−3)+2)=12(d2−1)( ...
  92. [92]
    [PDF] FOUNDATIONS OF ALGEBRAIC GEOMETRY CLASS 46
    Apr 17, 2008 · ... intersection. By. Bezout's theorem (the degree of a complete intersection of hypersurfaces is the product of the degrees of the hypersurfaces) ...
  93. [93]
    Intersections in CAD - SINTEF
    Jun 7, 2005 · You'll find inside the CAD-system extensive use of intersection algorithms to determine which parts of (curves and) surfaces describe the shells of the volume.
  94. [94]
    Types of Ray Tracing, Performance On GeForce GPUs, and More
    RT Cores on GeForce RTX GPUs provide dedicated hardware to accelerate BVH and ray / triangle intersection calculations, dramatically accelerating ray tracing.
  95. [95]
    [PDF] Extending GPU Ray-Tracing Units for Hierarchical Search ...
    Abstract—Specialized ray-tracing acceleration units have be- come a common feature in GPU hardware, enabling real-time ray-tracing of complex scenes for the ...
  96. [96]
    [PDF] Using Hardware Ray Transforms to Accelerate Ray/Primitive ...
    In this paper, we evaluate the use of RTX ray tracing capabilities to accelerate these primitives by tricking the GPU's instancing units into executing a ...
  97. [97]
    [PDF] Lecture 4.9. Positive definite and semidefinite forms - Purdue Math
    Apr 10, 2020 · Definitions. Q and A are called positive semidefinite if Q(x) ≥ 0 for all x. They are called positive definite if Q(x) > 0 for all x 6= 0.Missing: TA | Show results with:TA
  98. [98]
    Why is a positive definite matrix needed in the ellipsoid matrix ...
    Aug 19, 2015 · An ellipsoid centered at the origin is defined by the solutions x to the equation xTMx=1, where M is a positive definite matrix. How can I ...Relationship between "Ellipsoid" and "Quadratic Form"?What is the intuition behind $x^T A x - Math Stack ExchangeMore results from math.stackexchange.comMissing: TA | Show results with:TA
  99. [99]
    [PDF] eigenvalues, eigenvectors, canonical forms 95
    An ellipse has two independent principal axes. An ellipsoid ... not unique. We will now show that the principal axes of an ellipsoid are the eigenvectors.
  100. [100]
    Ellipsoid gaussian curvature - Applied Mathematics Consulting
    Oct 7, 2019 · The Gaussian curvature of an ellipsoid is given by K(x,y,z) = \frac{1}{a^2 b^2 c^2 \left(\frac{x^2}{a^4} + \frac{y^2}{b^4} + \frac{z^2}{c^4} \ ...
  101. [101]
    Ellipsoid -- from Wolfram MathWorld
    , the ellipsoid is a sphere. There are two families of parallel circular cross sections in every ellipsoid. However, the two coincide for spheroids (Hilbert ...Missing: properties | Show results with:properties
  102. [102]
    [PDF] Superquadrics and Angle-Preserving Transformations
    Superquadrics and Angle-Preserving Transformations · A. Barr · Published in IEEE Computer Graphics and… 1981 · Computer Science, Mathematics.Missing: original | Show results with:original
  103. [103]
    [PDF] the surface area are and the volume of n-dimensional sphere
    May 5, 2017 · n-dimensional sphere. Spring 2017. The surface area and the volume of the unit sphere are related as following: v(n) = s(n) n . (5). Consider ...
  104. [104]
    [PDF] Chapter 14 Curvature in Riemannian Manifolds - CIS UPenn
    We find that the sectional curvature of the sphere is con- stant and equal to +1. Let us now consider the hyperbolic space H n . This time the geodesic from p ...
  105. [105]
    [PDF] The Quaternions and the Spaces S3 , SU(2), SO(3), and RP
    The quaternions of norm 1, also called unit quaternions, are in bijec- tion with points of the real 3-sphere S. 3 . It is easy to verify that the unit.
  106. [106]
    [PDF] Lectures in Geometric Functional Analysis Roman Vershynin
    The course is a systematic introduction to the main techniques and results of geometric functional analysis. 1. Preliminaries on Banach spaces and linear ...<|control11|><|separator|>
  107. [107]
    [PDF] METRIC SPACES 1. Introduction As calculus developed, eventually ...
    Every closed ball in Rm is compact, since closed balls are closed subsets and are clearly bounded. Theorem 6.7. Every compact subset of a metric space is closed ...
  108. [108]
    [PDF] Hilbert Space Methods for Partial Differential Equations
    This book is an outgrowth of a course which we have given almost pe- riodically over the last eight years. It is addressed to beginning graduate.
  109. [109]
    [PDF] NOTES FOR MATH 5510, FALL 2017, V 0 1. Metric Spaces 1 1.1 ...
    Jul 23, 2017 · One way to visualize these metrics is by looking at their unit spheres, that is, {x ∈ Rn : d(0,x)=1}. First, define the terminology: Definition ...
  110. [110]
    Plato's Timaeus - Stanford Encyclopedia of Philosophy
    Oct 25, 2005 · In the Timaeus Plato presents an elaborately wrought account of the formation of the universe and an explanation of its impressive order and beauty.Overview of the Dialogue · Relation of the Timaeus to... · Physics · Bibliography
  111. [111]
    Aristotle's Natural Philosophy
    May 26, 2006 · Aristotle had a lifelong interest in the study of nature. He investigated a variety of different topics, ranging from general issues like motion, causation, ...
  112. [112]
    Archimedes - Biography - MacTutor - University of St Andrews
    In the first book of On the sphere and cylinder Archimedes shows that the surface of a sphere is four times that of a great circle, he finds the area of any ...
  113. [113]
    Apollonius (262 BC - 190 BC) - Biography - MacTutor
    The mathematician Apollonius was born in Perga, Pamphylia which today is known as Murtina, or Murtana and is now in Antalya, Turkey. Perga was a centre of ...
  114. [114]
    Descartes' Mathematics - Stanford Encyclopedia of Philosophy
    Nov 28, 2011 · To speak of René Descartes' contributions to the history of mathematics is to speak of his La Géométrie (1637), a short tract included with ...Descartes' Early Mathematical... · La Géométrie (1637) · Book One: Descartes...
  115. [115]
    Who calculated for the first time the volume (and surface area) of the ...
    Jan 24, 2015 · The principal step was no doubt made by Archimedes in On Sphere and Cylinder, where he proved rigorously that VS:VC=2:3, where VS is the volume ...<|separator|>
  116. [116]
    Gauss's Theorema Egregium -- from Wolfram MathWorld
    Gauss's theorema egregium states that the Gaussian curvature of a surface embedded in three-space may be understood intrinsically to that surface.Missing: sphere 1820s
  117. [117]
    Henri Poincaré - Stanford Encyclopedia of Philosophy
    Sep 3, 2013 · Poincaré is also famous for his 1904 conjecture concerning the topology of three-dimensional spheres which remained one of the major ...
  118. [118]
    Machine learning approaches for the optimization of packing ...
    Jul 20, 2023 · Here we discuss how machine learning methods can support the search for optimally dense packing shapes in a high-dimensional shape space.