Fact-checked by Grok 2 weeks ago

Three-dimensional space

Three-dimensional space, also known as 3D space, is a geometric model in which the position of any point is uniquely determined by three mutually coordinates, typically denoted as (x, y, z) in a Cartesian system, representing length, width, and height. This framework, known as three-dimensional space or \mathbb{R}^3, consists of all ordered triples of real numbers and assumes a flat, isotropic structure where distances and angles follow the principles of , such as the extended to three dimensions. It serves as the foundational setting for describing the arrangement and motion of physical objects in everyday experience, distinct from the one- or two-dimensional spaces used for lines or planes. In , three-dimensional space enables the study of vectors, which are quantities with magnitude and direction representable as arrows from the , and facilitates calculations of distances between points using the \sqrt{(x_2 - x_1)^2 + (y_2 - y_1)^2 + (z_2 - z_1)^2}. Beyond the standard Cartesian coordinates, alternative systems like cylindrical coordinates (r, θ, z) and spherical coordinates (ρ, θ, φ) are employed to simplify descriptions of rotationally symmetric objects or surfaces, such as cylinders or spheres. These tools underpin , where functions of three variables model volumes, surfaces, and gradients, and linear algebra, where subspaces and transformations preserve the space's dimensionality. From a physical perspective, three-dimensional space forms the arena for , , and , where phenomena like and propagation are analyzed assuming spatial and homogeneity on macroscopic scales. Cosmological models suggest that the universe's large-scale structure stabilized into three spatial dimensions during its early evolution, favoring this dimensionality over others for stable atomic and planetary formations. In , while embeds three-dimensional space within four-dimensional , the spatial component remains fundamentally three-dimensional for describing observable matter and fields.

History

Ancient and medieval perspectives

In , space was conceptualized not as an abstract void but as an integral aspect of the physical world, serving as a container for material bodies. , in his Physics (Book IV), defined place () as the innermost boundary of the containing body, emphasizing that is relational and dependent on the presence of bodies rather than an independent entity. This view portrayed three-dimensional as a filled with substances, where extension arises from the arrangement of matter, influencing later understandings of spatial containment without invoking empty voids. Euclid's Elements (c. 300 BCE) further shaped early intuitions of three-dimensional through synthetic methods, describing solids such as polyhedra and spheres via axioms and postulates without coordinate systems or algebraic formalism. Books XI–XIII of the Elements establish properties of planes and volumes intuitively, treating as a continuous medium for geometric constructions observable in everyday objects like buildings and celestial bodies. These works provided a foundational framework for visualizing spatial relations, prioritizing empirical deduction over measurement. Contributions from Indian and Islamic scholars expanded observational approaches to three-dimensional space, particularly through astronomy. , in his (499 CE), developed for modeling celestial motions, treating the and heavens as embedded in a three-dimensional spherical framework to compute planetary positions and eclipses. In the 11th century, advanced geodetic measurements by determining the Earth's radius using trigonometric observations from mountain elevations, confirming its sphericity and curvature with an accuracy close to modern values, thus refining conceptions of global spatial extent. During the medieval European period, scholastic thinkers synthesized these ideas with . , drawing on in works like , integrated the notion of space as a bounded container into a cosmological hierarchy where the finite, three-dimensional reflects divine order, with heavenly spheres encompassing earthly bodies in a . This reconciliation portrayed space as a created medium, harmonious with , bridging philosophical inquiry and religious . A pivotal development bridging medieval and Renaissance views occurred in the 15th century through artistic innovations, exemplified by Brunelleschi's experiments in around 1420. Using mirrors and peepholes to project the Baptistery's facade onto a painted panel, Brunelleschi demonstrated linear perspective, enabling two-dimensional representations that mimicked three-dimensional depth and spatial recession, thus enhancing perceptual understanding of volume and distance. These techniques, while artistic, laid groundwork for later mathematical formalizations of .

Modern mathematical development

The modern mathematical development of three-dimensional space began in the 17th century with ' introduction of Cartesian coordinates in his 1637 work , which allowed for the algebraic representation of points, lines, and surfaces in 3D space through ordered triples of numbers, transforming into an analytic discipline. This innovation enabled the precise description of spatial relationships using equations, bridging algebra and and laying the foundation for subsequent advancements in vector analysis and coordinate-based modeling. In the , Leonhard Euler advanced the study of polyhedra and space-filling structures, culminating in his 1752 formulation of the relation V - E + F = 2 for convex polyhedra, where V denotes vertices, E edges, and F faces, providing a topological invariant that characterizes the connectivity of 3D polyhedral forms. Euler's explorations also included analyses of regular polyhedra and tessellations, contributing to understandings of how shapes fill 3D space without gaps or overlaps. The 19th century saw significant progress in projective and relevant to 3D space. Carl Friedrich Gauss's 1827 demonstrated that the of a surface embedded in 3D space is an intrinsic property, independent of its embedding, which was later generalized to surfaces in higher dimensions. , in his 1827 Der barycentrische Calcül, introduced barycentric coordinates, facilitating projective treatments of points and figures in 3D by expressing positions as weighted combinations relative to reference points. extended projective geometry to lines in 3D space through his line coordinates, introduced in works from the 1840s and elaborated in 1868's Theorie der Flächen, representing lines via six and enabling algebraic studies of line complexes. Bernhard Riemann's 1854 lecture Über die Hypothesen, welche der Geometrie zu Grunde liegen developed the framework of , describing curved 3D spaces via metrics on manifolds and providing the mathematical basis for non-Euclidean geometries. Entering the 20th century, Henri Poincaré's foundational work in topology, particularly his 1895 Analysis Situs and subsequent papers, analyzed 3D manifolds as abstract spaces, introducing concepts like fundamental groups to classify their connectivity and homology, which distinguished simply connected spaces and influenced the study of 3D topological structures.

Euclidean geometry

Coordinate systems

In three-dimensional Euclidean space, the Cartesian coordinate system provides a standard framework for locating points using three mutually perpendicular axes intersecting at the origin. A point is represented by an ordered triple (x, y, z), where x, y, and z denote the signed distances from the origin along the respective axes, typically oriented as the x-axis (horizontal), y-axis (depth), and z-axis (vertical). This system, introduced by René Descartes in his 1637 work La Géométrie, extends the two-dimensional plane to allow precise positioning in space. The distance between two points (x_1, y_1, z_1) and (x_2, y_2, z_2) in this system is given by the Euclidean metric: \sqrt{(x_2 - x_1)^2 + (y_2 - y_1)^2 + (z_2 - z_1)^2}, which generalizes the Pythagorean theorem to three dimensions. Alternative coordinate systems, such as cylindrical and spherical, simplify representations when symmetry about an axis or radial structure is present. Cylindrical coordinates (r, \theta, z) describe a point by its radial distance r \geq 0 from the z-axis in the xy-plane, the azimuthal angle \theta (measured from the positive x-axis), and the height z along the z-axis. The conversion to Cartesian coordinates is: x = r \cos \theta, \quad y = r \sin \theta, \quad z = z. For volume integrals, the Jacobian determinant yields the volume element r \, dr \, d\theta \, dz, accounting for the scaling in the radial direction. Spherical coordinates (r, \theta, \phi) use the radial distance r \geq 0 from the origin, the polar angle \theta (from the positive z-axis, $0 \leq \theta \leq \pi), and the azimuthal angle \phi (from the positive x-axis in the xy-plane, $0 \leq \phi < 2\pi). The conversion formulas are: x = r \sin \theta \cos \phi, \quad y = r \sin \theta \sin \phi, \quad z = r \cos \theta. This system is particularly useful for problems exhibiting radial symmetry, such as those involving spheres or isotropic fields. Coordinate transformations, such as rotations, preserve distances and angles in Euclidean space and are represented by orthogonal matrices with determinant 1. For a counterclockwise rotation by angle \theta around the z-axis, the transformation matrix applied to a point's Cartesian coordinates is: \begin{pmatrix} \cos \theta & -\sin \theta & 0 \\ \sin \theta & \cos \theta & 0 \\ 0 & 0 & 1 \end{pmatrix}. Such matrices facilitate changing the orientation of axes or objects while maintaining the underlying geometry.

Lines, planes, and distances

In three-dimensional Euclidean space, a line can be defined using parametric equations that describe its position as a function of a parameter t. The parametric form passing through a point (x_0, y_0, z_0) with direction vector \langle a, b, c \rangle is given by \begin{align*} x &= x_0 + a t, \\ y &= y_0 + b t, \\ z &= z_0 + c t, \end{align*} where t \in \mathbb{R}./01%3A_Vectors_and_Geometry_in_Two_and_Three_Dimensions/1.05%3A_Equations_of_Lines_in_3d) The direction vector \langle a, b, c \rangle indicates the orientation and scaling of the line, and any scalar multiple of it yields an equivalent representation. To determine if two lines intersect, their parametric equations are set equal to solve for parameters t and s. For lines \mathbf{r}_1 = \mathbf{p}_1 + t \mathbf{d}_1 and \mathbf{r}_2 = \mathbf{p}_2 + s \mathbf{d}_2, intersection occurs if there exist scalars t and s such that \mathbf{p}_1 + t \mathbf{d}_1 = \mathbf{p}_2 + s \mathbf{d}_2, which rearranges to ( \mathbf{p}_2 - \mathbf{p}_1 ) = t \mathbf{d}_1 - s \mathbf{d}_2; a unique solution implies intersection at that point, while no solution indicates skew or parallel non-intersecting lines. If the direction vectors are parallel (one is a scalar multiple of the other) and the lines do not coincide, they are parallel and do not intersect unless the vector between points on each lies in the span of the direction. A plane in three-dimensional space is defined by the general equation a x + b y + c z + d = 0, where \langle a, b, c \rangle is the normal vector perpendicular to the plane. This normal vector determines the plane's orientation, and the equation can be derived from a point on the plane and the normal via a (x - x_0) + b (y - y_0) + c (z - z_0) = 0./01%3A_Vectors_and_Geometry_in_Two_and_Three_Dimensions/1.05%3A_Equations_of_Lines_in_3d) The distance from a point (x_0, y_0, z_0) to the plane a x + b y + c z + d = 0 is the length of the perpendicular from the point to the plane, calculated as \frac{|a x_0 + b y_0 + c z_0 + d|}{\sqrt{a^2 + b^2 + c^2}}. This formula arises from projecting the vector from a point on the plane to (x_0, y_0, z_0) onto the unit . The angle between two lines is found using the cosine of the angle \theta between their direction vectors \mathbf{d}_1 and \mathbf{d}_2, given by \cos \theta = \frac{|\mathbf{d}_1 \cdot \mathbf{d}_2|}{|\mathbf{d}_1| |\mathbf{d}_2|}, where the acute angle is considered. Similarly, the angle between two planes is the angle between their \mathbf{n}_1 and \mathbf{n}_2, with \cos \phi = \frac{|\mathbf{n}_1 \cdot \mathbf{n}_2|}{|\mathbf{n}_1| |\mathbf{n}_2|}. For the angle between a line with direction \mathbf{d} and a plane with \mathbf{n}, the setup involves the complement of the angle between \mathbf{d} and \mathbf{n}, using \sin \psi = \frac{|\mathbf{d} \cdot \mathbf{n}|}{|\mathbf{d}| |\mathbf{n}|} for the acute angle \psi./01%3A_Vectors_and_Geometry_in_Two_and_Three_Dimensions/1.05%3A_Equations_of_Lines_in_3d) For two skew lines (non-intersecting and non-parallel) with parametric forms \mathbf{r}_1 = \mathbf{p}_1 + t \mathbf{d}_1 and \mathbf{r}_2 = \mathbf{p}_2 + s \mathbf{d}_2, the shortest distance is the length of the common perpendicular, given by \frac{|(\mathbf{p}_2 - \mathbf{p}_1) \cdot (\mathbf{d}_1 \times \mathbf{d}_2)|}{|\mathbf{d}_1 \times \mathbf{d}_2|}. This expression uses the cross product \mathbf{d}_1 \times \mathbf{d}_2 to find the direction perpendicular to both lines, and the scalar triple product to project the vector between points onto that direction.

Spheres, balls, and polytopes

In three-dimensional Euclidean space, a sphere is defined as the set of all points equidistant from a fixed center point, with that distance being the radius r. Using Cartesian coordinates, the equation of a sphere centered at (a, b, c) is given by (x - a)^2 + (y - b)^2 + (z - c)^2 = r^2. This locus represents the surface of the sphere. The surface area of the sphere is $4\pi r^2, derived by considering the sphere as the limit of polyhedral approximations or through integration in spherical coordinates. The volume enclosed by the sphere, known as the ball of radius r, is \frac{4}{3}\pi r^3, which can be obtained via triple integration over the region or by the method of Cavalieri's principle. A ball in three dimensions is the solid object comprising the sphere and its interior, defined as the set of points whose distance from the center is at most r. On the sphere's surface, great circles—formed by the intersection of the sphere with any plane passing through its center—represent the geodesics, or shortest paths connecting two points along the surface. These curves are the three-dimensional analogues of straight lines and have constant curvature equal to that of the sphere. Convex polytopes in three dimensions are polyhedra, bounded by flat polygonal faces, straight edges, and vertices. The regular convex polyhedra, called , are classified into five types: the (4 triangular faces), (6 square faces), (8 triangular faces), (12 pentagonal faces), and (20 triangular faces); each has congruent regular polygonal faces and the same number meeting at every vertex. For any convex polyhedron that is topologically equivalent to a sphere, the satisfies \chi = V - E + F = 2, where V, E, and F denote the numbers of vertices, edges, and faces, respectively; this relation holds due to the polyhedron's spherical topology and can be verified inductively by decomposition. Some convex polyhedra admit space-filling tessellations, partitioning three-dimensional space without gaps or overlaps. The cubic honeycomb, consisting of identical cubes arranged in a lattice, is a prominent example, with each cube sharing faces with six neighbors. Volumes of Platonic solids provide concrete measures of their spatial extent; for instance, a cube of side length a has volume a^3, while a regular tetrahedron of side length a has volume \frac{\sqrt{2}}{12} a^3, computed by dividing the tetrahedron into pyramids or using vector cross products for the enclosed space.

Quadric surfaces and surfaces of revolution

Quadric surfaces in three-dimensional Euclidean space are defined by second-degree polynomial equations in the variables x, y, and z. The general equation takes the form Ax^2 + By^2 + Cz^2 + axy + bxz + cyz + d_1 x + d_2 y + d_3 z + E = 0, where the coefficients determine the specific type of surface through the eigenvalues of the associated quadratic form or by completing the square and translating coordinates. These surfaces are classified into non-degenerate types—ellipsoids, hyperboloids of one or two sheets, elliptic and hyperbolic paraboloids—and degenerate cases such as cones, cylinders, and pairs of planes, based on the signs and ranks of the quadratic terms after canonical reduction. Among these, the ellipsoid represents a bounded, closed surface analogous to an ellipse stretched in three dimensions, with the standard equation \frac{x^2}{a^2} + \frac{y^2}{b^2} + \frac{z^2}{c^2} = 1, where a, b, and c are positive semi-axes lengths; when a = b = c, it reduces to a . The hyperbolic paraboloid, a ruled surface known for its saddle-like shape, features hyperbolic cross-sections and is given by z = \frac{x^2}{a^2} - \frac{y^2}{b^2} in canonical form, exhibiting both positive and negative curvatures along principal directions. Surfaces of revolution arise by rotating a curve in a plane around an axis lying in that plane but not intersecting the curve, producing rotationally symmetric surfaces in three dimensions. For instance, revolving a semicircle about its diameter yields a , while rotating a circle offset from the axis generates a , whose implicit equation is \left( \sqrt{x^2 + y^2} - R \right)^2 + z^2 = r^2, with R > r > 0 denoting the major and minor radii, respectively. Pappus's centroid theorem provides a method to compute areas and volumes of such surfaces without integration: the lateral surface area equals the arc length of the generating curve times the circumference described by its centroid (i.e., $2\pi times the centroid's distance to the axis), and the enclosed volume equals the area under the curve times the same circumferential distance. This theorem, attributed to Pappus of Alexandria in the 4th century CE, relies on the centroid's definition as the average position weighted by arc length or area.

Linear algebra

Vectors, dot product, and norms

In three-dimensional Euclidean space, vectors are commonly represented as ordered triples of real numbers, \vec{v} = (v_x, v_y, v_z), where v_x, v_y, and v_z are the components along the respective Cartesian axes. This representation corresponds to the displacement from the origin to a point in \mathbb{R}^3. Vector addition is performed component-wise: \vec{u} + \vec{v} = (u_x + v_x, u_y + v_y, u_z + v_z), which geometrically corresponds to the parallelogram rule. Scalar multiplication scales the vector by a real number k, yielding k\vec{v} = (k v_x, k v_y, k v_z), altering its magnitude while preserving direction (or reversing it if k < 0). The dot product of two vectors \vec{u} = (u_x, u_y, u_z) and \vec{v} = (v_x, v_y, v_z) in three dimensions is defined algebraically as \vec{u} \cdot \vec{v} = u_x v_x + u_y v_y + u_z v_z. Geometrically, it equals \|\vec{u}\| \|\vec{v}\| \cos \theta, where \theta is the angle between the vectors and \|\cdot\| denotes the ; this relation links the algebraic form to the spatial orientation. Two nonzero vectors are orthogonal if their dot product is zero, as \cos \theta = 0 implies \theta = 90^\circ. The Euclidean norm, or length, of a vector \vec{v} is given by \|\vec{v}\| = \sqrt{\vec{v} \cdot \vec{v}} = \sqrt{v_x^2 + v_y^2 + v_z^2}, providing a measure of magnitude invariant under rotations. A unit vector, with norm 1, is obtained by normalizing: \hat{v} = \vec{v} / \|\vec{v}\| for \vec{v} \neq \vec{0}. The vector projection of \vec{v} onto \vec{u} (nonzero) is \operatorname{proj}_{\vec{u}} \vec{v} = \left( \frac{\vec{v} \cdot \vec{u}}{\|\vec{u}\|^2} \right) \vec{u}, representing the component of \vec{v} parallel to \vec{u}. These concepts find applications in determining the angle between vectors via \cos \theta = \frac{\vec{u} \cdot \vec{v}}{\|\vec{u}\| \|\vec{v}\|}, essential for geometric computations. In physics, the dot product computes work as W = \vec{F} \cdot \vec{d}, where \vec{F} is force and \vec{d} is displacement, capturing only the component of force along the path.

Cross product and orientations

In three-dimensional Euclidean space, the cross product of two vectors \vec{u} = (u_x, u_y, u_z) and \vec{v} = (v_x, v_y, v_z) is a vector defined by the determinant-like formula \vec{u} \times \vec{v} = \begin{vmatrix} \mathbf{i} & \mathbf{j} & \mathbf{k} \\ u_x & u_y & u_z \\ v_x & v_y & v_z \end{vmatrix} = (u_y v_z - u_z v_y, \, u_z v_x - u_x v_z, \, u_x v_y - u_y v_x). /01:_Vectors_in_Euclidean_Space/1.04:_Cross_Product) This operation yields a vector perpendicular to both \vec{u} and \vec{v}, with magnitude \|\vec{u} \times \vec{v}\| = \|\vec{u}\| \|\vec{v}\| \sin \theta, where \theta is the angle between them; geometrically, this magnitude equals the area of the parallelogram formed by \vec{u} and \vec{v} as adjacent sides./01:_Vectors_in_Euclidean_Space/1.04:_Cross_Product) Key properties of the cross product include anticommutativity, \vec{u} \times \vec{v} = -\vec{v} \times \vec{u}, and orthogonality to its input vectors, \vec{u} \cdot (\vec{u} \times \vec{v}) = 0 and \vec{v} \cdot (\vec{u} \times \vec{v}) = 0./01:_Vectors_in_Euclidean_Space/1.04:_Cross_Product) The direction follows the right-hand rule: aligning the fingers of the right hand with \vec{u} and curling them toward \vec{v} points the thumb in the direction of \vec{u} \times \vec{v}. These attributes make the cross product useful for determining a normal vector to the plane spanned by \vec{u} and \vec{v}, essential in applications like surface parameterization./01:_Vectors_in_Euclidean_Space/1.04:_Cross_Product) In physics, the cross product computes torque as \vec{\tau} = \vec{r} \times \vec{F}, where \vec{r} is the position vector from the pivot to the force application point and \vec{F} is the force, yielding a vector whose magnitude is r F \sin \theta and direction indicates the rotation axis. For volumes, the scalar triple product \vec{a} \cdot (\vec{b} \times \vec{c}) gives the signed volume of the parallelepiped spanned by \vec{a}, \vec{b}, and \vec{c}, with the absolute value representing the actual volume. The cross product inherently encodes orientations through its handedness, distinguishing chiral (handed) structures in 3D space via the right-hand rule, which selects one of two possible perpendicular directions. This vector-valued binary operation is unique to three dimensions; in higher dimensions, analogous constructions yield higher-rank tensors or subspaces rather than vectors.

Abstract vector spaces

In the context of three-dimensional space, the algebraic structure can be abstracted to a finite-dimensional over the real numbers \mathbb{R}, providing a foundation for linear operations independent of specific geometric embeddings. A V over \mathbb{R} is a set equipped with two operations: vector addition +\colon V \times V \to V and scalar multiplication \cdot\colon \mathbb{R} \times V \to V, satisfying the following axioms: for all \mathbf{u}, \mathbf{v}, \mathbf{w} \in V and \alpha, \beta \in \mathbb{R},
  • Associativity of addition: (\mathbf{u} + \mathbf{v}) + \mathbf{w} = \mathbf{u} + (\mathbf{v} + \mathbf{w}),
  • Commutativity of addition: \mathbf{u} + \mathbf{v} = \mathbf{v} + \mathbf{u},
  • Existence of zero vector: there exists \mathbf{0} \in V such that \mathbf{u} + \mathbf{0} = \mathbf{u},
  • Additive inverses: for each \mathbf{u}, there exists -\mathbf{u} such that \mathbf{u} + (-\mathbf{u}) = \mathbf{0},
  • Distributivity over vector addition: \alpha (\mathbf{u} + \mathbf{v}) = \alpha \mathbf{u} + \alpha \mathbf{v},
  • Distributivity over scalar addition: (\alpha + \beta) \mathbf{u} = \alpha \mathbf{u} + \beta \mathbf{u},
  • Compatibility: \alpha (\beta \mathbf{u}) = (\alpha \beta) \mathbf{u},
  • Identity for scalar multiplication: $1 \cdot \mathbf{u} = \mathbf{u}.
A subspace W \subseteq V is a subset that forms a vector space under the induced operations, closed under addition and scalar multiplication, and containing the zero vector; examples include the origin \{\mathbf{0}\}, the entire space V, and lines or planes through the origin in \mathbb{R}^3. The dimension of a vector space is the cardinality of a basis, a maximal linearly independent set that spans V; linear independence means that the only linear combination yielding the zero vector is the trivial one with all coefficients zero. In three-dimensional space modeled as \mathbb{R}^3, the dimension is 3, so no set of four vectors can be linearly independent, as any such set is dependent by the properties of finite-dimensional spaces over \mathbb{R}. A standard basis for \mathbb{R}^3 is \{\mathbf{e}_1, \mathbf{e}_2, \mathbf{e}_3\}, where \mathbf{e}_1 = (1,0,0), \mathbf{e}_2 = (0,1,0), and \mathbf{e}_3 = (0,0,1), allowing any vector to be uniquely expressed as \alpha \mathbf{e}_1 + \beta \mathbf{e}_2 + \gamma \mathbf{e}_3 for \alpha, \beta, \gamma \in \mathbb{R}. Affine spaces extend vector spaces by distinguishing points from vectors, modeling translations without a fixed origin; an affine space A over \mathbb{R}^3 consists of points P such that differences P - Q form vectors in the associated vector space \vec{A}, with translations defined by adding vectors to points. Barycentric coordinates represent a point P in a three-dimensional affine space as an affine combination P = \alpha A + \beta B + \gamma C where \alpha + \beta + \gamma = 1 and \alpha, \beta, \gamma \geq 0 for points inside the simplex formed by basis points A, B, C, generalizing mass distributions at vertices. Inner product spaces refine vector spaces by introducing a bilinear form that captures angles and lengths; specifically, an inner product \langle \cdot, \cdot \rangle\colon V \times V \to \mathbb{R} satisfies bilinearity (\langle \alpha \mathbf{u} + \beta \mathbf{v}, \mathbf{w} \rangle = \alpha \langle \mathbf{u}, \mathbf{w} \rangle + \beta \langle \mathbf{v}, \mathbf{w} \rangle and similarly for the second argument), positive-definiteness (\langle \mathbf{u}, \mathbf{u} \rangle > 0 for \mathbf{u} \neq \mathbf{0}, and \langle \mathbf{0}, \mathbf{0} \rangle = 0), and symmetry (\langle \mathbf{u}, \mathbf{v} \rangle = \langle \mathbf{v}, \mathbf{u} \rangle). Orthogonality holds when \langle \mathbf{u}, \mathbf{v} \rangle = 0. The three-dimensional Euclidean space \mathbb{R}^3 is an inner product space with the standard dot product \langle \mathbf{u}, \mathbf{v} \rangle = u_1 v_1 + u_2 v_2 + u_3 v_3, generalizing the concrete operations to abstract settings while preserving geometric intuition.

Calculus

Vector calculus operators

In three-dimensional Euclidean space, vector calculus operators such as the , , , and Laplacian provide essential tools for analyzing scalar and vector fields, capturing local properties like rates of change, sources, rotations, and diffusion. These operators, collectively denoted using the (or nabla) symbol ∇, are defined primarily in Cartesian coordinates but extend to other systems like cylindrical and spherical coordinates, facilitating computations in diverse geometric contexts. The gradient of a f(x, y, z), denoted \nabla f, is a that points in the direction of the steepest ascent of f and whose magnitude |\nabla f| equals the rate of that ascent. In Cartesian coordinates, it is expressed as \nabla f = \left( \frac{\partial f}{\partial x}, \frac{\partial f}{\partial y}, \frac{\partial f}{\partial z} \right). This operator transforms a scalar into a , highlighting directional derivatives aligned with the field's increase. The divergence of a vector field \vec{F} = (F_x, F_y, F_z), denoted \nabla \cdot \vec{F}, quantifies the net flux emanating from a point, positive for sources and negative for sinks, representing the rate at which the field's density exits a local volume. In Cartesian coordinates, it takes the form \nabla \cdot \vec{F} = \frac{\partial F_x}{\partial x} + \frac{\partial F_y}{\partial y} + \frac{\partial F_z}{\partial z}. This scalar operator measures expansion or contraction within the field. The of a vector field \vec{F}, denoted \nabla \times \vec{F}, is a vector field whose magnitude indicates the maximum (circulation per unit area) at a point and whose direction aligns with the axis of that , following the . A field is irrotational if \nabla \times \vec{F} = \vec{0}. In Cartesian coordinates, it is given by \nabla \times \vec{F} = \left( \frac{\partial F_z}{\partial y} - \frac{\partial F_y}{\partial z}, \frac{\partial F_x}{\partial z} - \frac{\partial F_z}{\partial x}, \frac{\partial F_y}{\partial x} - \frac{\partial F_x}{\partial y} \right). This antisymmetric operator detects local . The Laplacian of a f, denoted \nabla^2 f or \Delta f, is the of the , \nabla \cdot (\nabla f), and serves as a measure of the field's variation or , appearing in equations for , and potentials. Functions satisfying \nabla^2 f = 0 are , exhibiting mean-value properties over spheres. In Cartesian coordinates, \nabla^2 f = \frac{\partial^2 f}{\partial x^2} + \frac{\partial^2 f}{\partial y^2} + \frac{\partial^2 f}{\partial z^2}. This second-order scalar operator is fundamental in many physical laws. These operators adapt to curvilinear coordinates for problems with symmetry. The table below summarizes their expressions in Cartesian, cylindrical (r, \theta, z), and spherical (r, \theta, \phi) coordinates, where scale factors account for the geometry.
OperatorCartesian (x, y, z)Cylindrical (r, \theta, z)Spherical (r, \theta, \phi)
Gradient \nabla f\left( \frac{\partial f}{\partial x}, \frac{\partial f}{\partial y}, \frac{\partial f}{\partial z} \right)\left( \frac{\partial f}{\partial r}, \frac{1}{r} \frac{\partial f}{\partial \theta}, \frac{\partial f}{\partial z} \right)\left( \frac{\partial f}{\partial r}, \frac{1}{r} \frac{\partial f}{\partial \theta}, \frac{1}{r \sin \theta} \frac{\partial f}{\partial \phi} \right)
Divergence \nabla \cdot \vec{F}\frac{\partial F_x}{\partial x} + \frac{\partial F_y}{\partial y} + \frac{\partial F_z}{\partial z}\frac{1}{r} \frac{\partial (r F_r)}{\partial r} + \frac{1}{r} \frac{\partial F_\theta}{\partial \theta} + \frac{\partial F_z}{\partial z}\frac{1}{r^2} \frac{\partial (r^2 F_r)}{\partial r} + \frac{1}{r \sin \theta} \frac{\partial (F_\theta \sin \theta)}{\partial \theta} + \frac{1}{r \sin \theta} \frac{\partial F_\phi}{\partial \phi}
Curl \nabla \times \vec{F}\left( \frac{\partial F_z}{\partial y} - \frac{\partial F_y}{\partial z}, \frac{\partial F_x}{\partial z} - \frac{\partial F_z}{\partial x}, \frac{\partial F_y}{\partial x} - \frac{\partial F_x}{\partial y} \right)\left( \frac{1}{r} \frac{\partial F_z}{\partial \theta} - \frac{\partial F_\theta}{\partial z}, \frac{\partial F_r}{\partial z} - \frac{\partial F_z}{\partial r}, \frac{1}{r} \left( \frac{\partial (r F_\theta)}{\partial r} - \frac{\partial F_r}{\partial \theta} \right) \right)\frac{1}{r \sin \theta} \left( \frac{\partial (F_\phi \sin \theta)}{\partial \theta} - \frac{\partial F_\theta}{\partial \phi} \right) \hat{r} + \frac{1}{r} \left( \frac{1}{\sin \theta} \frac{\partial F_r}{\partial \phi} - \frac{\partial (r F_\phi)}{\partial r} \right) \hat{\theta} + \frac{1}{r} \left( \frac{\partial (r F_\theta)}{\partial r} - \frac{\partial F_r}{\partial \theta} \right) \hat{\phi}
Laplacian \nabla^2 f\frac{\partial^2 f}{\partial x^2} + \frac{\partial^2 f}{\partial y^2} + \frac{\partial^2 f}{\partial z^2}\frac{1}{r} \frac{\partial}{\partial r} \left( r \frac{\partial f}{\partial r} \right) + \frac{1}{r^2} \frac{\partial^2 f}{\partial \theta^2} + \frac{\partial^2 f}{\partial z^2}\frac{1}{r^2} \frac{\partial}{\partial r} \left( r^2 \frac{\partial f}{\partial r} \right) + \frac{1}{r^2 \sin \theta} \frac{\partial}{\partial \theta} \left( \sin \theta \frac{\partial f}{\partial \theta} \right) + \frac{1}{r^2 \sin^2 \theta} \frac{\partial^2 f}{\partial \phi^2}

Integrals over curves, surfaces, and volumes

In three-dimensional , integrals over curves, surfaces, and volumes provide essential tools for computing accumulated quantities such as arc lengths, work, surface areas, fluxes, and total masses or volumes of regions. These integrals extend the concepts of one-dimensional definite integrals to higher-dimensional domains, often requiring parameterizations of the domains to evaluate them explicitly. Line integrals arise when integrating scalar or vector fields along a curve C in \mathbb{R}^3. For a scalar field f: \mathbb{R}^3 \to \mathbb{R}, the line integral \int_C f \, ds measures the accumulation of f weighted by arc length along C, defined as \int_C f \, ds = \int_a^b f(\vec{r}(t)) \|\vec{r}'(t)\| \, dt, where \vec{r}: [a, b] \to \mathbb{R}^3 is a smooth parameterization of C with \vec{r}'(t) \neq \vec{0}. For a vector field \vec{F}: \mathbb{R}^3 \to \mathbb{R}^3, the line integral \int_C \vec{F} \cdot d\vec{r} computes quantities like the work done by \vec{F} along C, given by \int_C \vec{F} \cdot d\vec{r} = \int_a^b \vec{F}(\vec{r}(t)) \cdot \vec{r}'(t) \, dt. This form arises from approximating the integral as a sum of \vec{F} dotted with small displacements d\vec{r} \approx \vec{r}'(t) \, dt. A representative example is the work done by the force field \vec{F}(x, y, z) = \langle y, -x, z \rangle on a particle moving along the helix C parameterized by \vec{r}(t) = \langle \cos t, \sin t, t \rangle for $0 \leq t \leq 2\pi. Substituting yields \int_C \vec{F} \cdot d\vec{r} = \int_0^{2\pi} (\sin t \cdot (-\sin t) + (-\cos t) \cdot \cos t + t \cdot 1) \, dt = \int_0^{2\pi} (t - 1) \, dt = 2\pi^2 - 2\pi, illustrating how the integral captures the net work despite the helical path's twist. Surface integrals extend these ideas to two-dimensional surfaces S in \mathbb{R}^3. For a scalar field f: \mathbb{R}^3 \to \mathbb{R}, the surface integral \int_S f \, dS accumulates f over the surface area, expressed using a parameterization \vec{r}(u, v): D \to S (where D \subset \mathbb{R}^2) as \int_S f \, dS = \iint_D f(\vec{r}(u, v)) \|\vec{r}_u(u, v) \times \vec{r}_v(u, v)\| \, du \, dv. The cross product \vec{r}_u \times \vec{r}_v provides the magnitude of the normal vector, approximating surface area elements. For a vector field \vec{F}, the flux integral \int_S \vec{F} \cdot d\vec{S} measures the net flow through S, defined as \int_S \vec{F} \cdot d\vec{S} = \iint_D \vec{F}(\vec{r}(u, v)) \cdot (\vec{r}_u(u, v) \times \vec{r}_v(u, v)) \, du \, dv, with the orientation determined by the direction of \vec{r}_u \times \vec{r}_v. An example is the flux of the radial field \vec{F}(x, y, z) = \langle x, y, z \rangle through the unit sphere S: x^2 + y^2 + z^2 = 1, oriented outward. Using spherical coordinates \vec{r}(\theta, \phi) = \langle \sin\phi \cos\theta, \sin\phi \sin\theta, \cos\phi \rangle for $0 \leq \theta \leq 2\pi, $0 \leq \phi \leq \pi, the integral simplifies to \iint_D 1 \cdot \sin\phi \, d\theta \, d\phi = 4\pi, reflecting the field's divergence from the origin. Volume integrals compute accumulations over three-dimensional regions V \subset \mathbb{R}^3. For a scalar field f: \mathbb{R}^3 \to \mathbb{R}, the triple integral in Cartesian coordinates is \iiint_V f(x, y, z) \, dV = \int_c^d \int_{a(x)}^{b(x)} \int_{g(x,y)}^{h(x,y)} f(x, y, z) \, dz \, dy \, dx, where the limits describe V. To evaluate in other coordinate systems, such as spherical or cylindrical, a \vec{x} = \vec{g}(\vec{u}) (with \vec{u} = (u, v, w)) transforms the integral via the determinant: \iiint_V f(\vec{x}) \, dV = \iiint_{V^*} f(\vec{g}(\vec{u})) \left| \det \left( \frac{\partial(x,y,z)}{\partial(u,v,w)} \right) \right| \, du \, dv \, dw, where V^* is the image of V under the inverse map, and the absolute value of the accounts for scaling under the transformation. This is crucial for regions with , ensuring the integral remains under coordinate changes.

Integral theorems

Integral theorems in three-dimensional form a cornerstone of , providing powerful relationships between local differential operators—such as the , , and —and global integrals over curves, surfaces, and volumes. These theorems enable the transformation of difficult integrals into more tractable forms, facilitating computations in and , particularly for fields like and where conservation principles manifest through such equivalences. By linking infinitesimal changes to overall accumulations, they underpin the understanding of how fields behave across extended regions. The , also known as the , applies to conservative vector fields, which are the gradients of scalar potentials. For a f that is continuously differentiable on a simply connected domain and a smooth curve C from point P to Q, the theorem states: \int_C \nabla f \cdot d\vec{r} = f(Q) - f(P). This result generalizes the one-dimensional to paths in space, showing that the depends only on the endpoints for conservative fields, motivated by the path-independence of work done by conservative forces like . In two dimensions, serves as a special case, relating a around a positively oriented, , closed C enclosing a region D to a double integral over D: \int_C (P \, dx + Q \, dy) = \iint_D \left( \frac{\partial Q}{\partial x} - \frac{\partial P}{\partial y} \right) dA, for continuously differentiable P and Q. This theorem, originally stated by George Green in 1828, motivates the extension to higher dimensions by equating circulation around a boundary to the enclosed "vorticity," applicable under conditions of orientability and smoothness of the region. Stokes' theorem generalizes Green's theorem to three dimensions, connecting the surface integral of the curl over an oriented surface S with boundary curve \partial S to the line integral around that boundary: \int_S (\nabla \times \vec{F}) \cdot d\vec{S} = \int_{\partial S} \vec{F} \cdot d\vec{r}, for a vector field \vec{F} with continuous partial derivatives. First posed as an exam question by George Gabriel Stokes in 1854 and proved by Hermann Hankel in 1861, this theorem motivates the interpretation of curl as measuring rotation, allowing the flux of rotation through a surface to equal circulation along its edge, assuming the surface is piecewise smooth, orientable, and the field is defined on a suitable domain. The , also known as , relates the volume of the over a bounded region V with piecewise smooth boundary surface S to the through that surface: \iiint_V (\nabla \cdot \vec{F}) \, dV = \iint_S \vec{F} \cdot d\vec{S}, for a \vec{F} with continuous first partial derivatives. Formulated by around 1835 and published posthumously in 1867, it motivates the view of as a or density, equating total outflow to net sources inside, under conditions that V is closed, orientable, and bounded with sufficient smoothness. These theorems share common assumptions, including the continuous differentiability of the fields involved, the and piecewise smoothness of the domains, and the existence of suitable parameterizations, ensuring the integrals are well-defined. A key application is computing through complex surfaces without direct evaluation; for instance, the simplifies calculations over closed volumes by converting them to easier volume integrals of , as seen in deriving conservation laws for or charge in fluid or electromagnetic contexts.

Topology and geometry

Topological properties

A 3-manifold is defined as a Hausdorff topological space that is locally homeomorphic to Euclidean 3-space \mathbb{R}^3. This means every point in the space has a neighborhood homeomorphic to an open ball in \mathbb{R}^3, capturing the local flatness essential to three-dimensional . Euclidean 3-space itself serves as the standard example of an 3-manifold, where ensures the existence of a consistent choice of orientation across the space, allowing for a global distinction between left-handed and right-handed coordinate systems without inconsistencies. The \pi_1(\mathbb{R}^3) is trivial, consisting solely of the \{e\}, indicating that every closed in \mathbb{R}^3 can be continuously contracted to a point. This trivial structure underscores the simply connected nature of \mathbb{R}^3, distinguishing it from spaces with non-trivial loops, such as those encircling obstacles. In the context of within 3-dimensional space, embeddings of circles into \mathbb{R}^3 give rise to knots. Homology theory provides further invariants for 3-manifolds, with the Betti numbers of \mathbb{R}^3 being b_0 = 1 (one connected component), b_1 = 0 (no 1-dimensional holes), b_2 = 0 (no 2-dimensional voids), and b_3 = 0 (no 3-dimensional cavities), reflecting its contractible topology. For closed orientable 3-manifolds, establishes an isomorphism H_k(M; \mathbb{Z}) \cong H^{3-k}(M; \mathbb{Z}), linking in k to in the complementary , which facilitates the study of manifold duality and has profound implications for understanding their . Embeddings of 3-manifolds into higher-dimensional Euclidean spaces are governed by the , which guarantees that any smooth can be into \mathbb{R}^6, providing a realization of abstract 3-dimensional topologies within a finite-dimensional ambient space while preserving topological properties. A striking of embeddings in 3-space is , where the 2-sphere \mathbb{S}^2 in \mathbb{R}^3 can be continuously deformed to its mirror image through a regular , without self-intersections at the boundary, demonstrating the flexibility of surfaces in three dimensions. The , positing that every simply connected closed is homeomorphic to the 3-sphere \mathbb{S}^3, was resolved affirmatively by in 2003 using techniques, confirming \mathbb{S}^3 as the sole simply connected up to homeomorphism.

Non-Euclidean and finite geometries

Non-Euclidean geometries in three dimensions extend the principles of by incorporating constant , leading to spaces where the parallel postulate does not hold in its classical form. In 3-space, denoted \mathbb{H}^3, the geometry features constant negative , resulting in infinitely many through a point not on a given line. This space can be modeled using the upper half-space or embeddings, where distances expand exponentially, contrasting with the linear growth in . 3-space satisfies Euclid's first four postulates but replaces the parallel postulate with one allowing multiple parallels, a development originating from the independent work of Lobachevsky and Bolyai. Elliptic 3-space, in contrast, possesses constant positive curvature, where no exist; any two lines intersect. This geometry identifies antipodal points on a , forming a closed, finite without , and the sum of angles in a exceeds 180 degrees. Like , it adheres to Euclid's first four postulates except for the parallel postulate, which is replaced by the assertion that there are no ; through a point not on a given line, every line through that point intersects the given line. The transition between these geometries is captured by the curvature parameter k in spatial metrics. A general form for the line element in such 3D spaces of constant is the Robertson-Walker spatial : ds^2 = \frac{dr^2}{1 - k r^2} + r^2 d\Omega^2, where d\Omega^2 = d\theta^2 + \sin^2\theta \, d\phi^2 is the on the 2-sphere, r is a radial coordinate, and k = -1, 0, +1 corresponds to , , and elliptic geometries, respectively. For k = -1, the space is infinite with negative ; for k = +1, it is finite and positively curved. This describes the hypersurfaces in cosmological models, highlighting how alters volume growth and behavior compared to flat space. Spherical geometry in three dimensions is realized on the S^3, the set of points (w,x,y,z) \in \mathbb{R}^4 satisfying w^2 + x^2 + y^2 + z^2 = 1, which has constant positive . The unit 3-sphere can be parameterized using unit , where points correspond to quaternions of norm 1, forming a under quaternion multiplication. Geodesics on S^3 are great circles, the shortest paths analogous to straight lines, but the space's compactness means all geodesics close after finite length, precluding infinite parallels. This structure underpins elliptic 3-space when quotiented appropriately. Finite geometries provide discrete analogs of continuous 3D spaces over finite fields \mathbb{F}_q, where q = p^k for prime p and k \geq 1. The finite \mathrm{PG}(3,q) consists of points as 1-dimensional subspaces and lines as 2-dimensional subspaces of \mathbb{F}_q^4, yielding \theta_3(q) = \frac{q^4 - 1}{q - 1} points and a similar number of lines, with every two points determining a unique line. This space captures projective properties without infinity, useful in . The \mathrm{AG}(3,q) is the \mathbb{F}_q^3, with q^3 points and parallel classes of lines, serving as a finite counterpart to 3-space where translations preserve structure. In finite 3D geometries, block designs emerge as combinatorial structures, such as 2-(v,k,1) designs where blocks are subsets covering every pair of points exactly once, often realized as subspaces in \mathrm{PG}(3,q) or \mathrm{AG}(3,q). These include projective planes embedded in higher dimensions or affine resolvable designs, with applications to error-correcting codes. Polytopes in these settings, like finite analogs of simplices or cubes, are defined via lattice points over \mathbb{F}_q, enabling discrete approximations of convex bodies for optimization. The failure of the parallel postulate is central to non- and finite cases: in and elliptic spaces, multiple or no parallels exist, respectively, while finite geometries limit lines to discrete sets, preventing infinite extensions altogether. In , 3D spatial slices of often exhibit non-Euclidean , as in Friedmann-Lemaître-Robertson-Walker models where the metric's [k](/page/K) parameter dictates open, flat, or closed universes, influencing cosmic expansion. Voxel-based representations in approximate continuous 3D spaces with finite cubic grids, discretizing for rendering and , akin to affine finite approximations but extended to irregular shapes via unions of voxels.

References

  1. [1]
    [PDF] 6.1 Three-Dimensional Space
    in three-dimensional space can be described by three coordinates x, y, and z. The geometric meaning of these coordinates is shown in Figures 1 and 2.
  2. [2]
    [PDF] n-Dimensional Euclidean Space and Matrices
    Definition of n space. As was learned in Math 1b, a point in Euclidean three space can be thought of in any of three ways: (i) as the set of triples (x, y, ...
  3. [3]
    Aspects of Three Dimensional Space with Dennis Sullivan
    Sep 1, 2023 · Three space is special because our physical reality including biology takes place there. In three dimensions mathematical mysteries related to ...
  4. [4]
    Linear Algebra, Part 5: Euclidean Vector Spaces (Mathematica)
    In three-dimensional space, the Euclidean distance is the length of a line segment between the two points. It can be calculated from the Cartesian ...
  5. [5]
    Three-Dimensional Coordinate Systems - UTSA
    Nov 2, 2021 · The polar coordinate system is extended into three dimensions with two different coordinate systems, the cylindrical and spherical coordinate systems.
  6. [6]
    Calculus III - 3-Dimensional Space - Pauls Online Math Notes
    Nov 16, 2022 · We will be looking at the equations of graphs in 3-D space as well as vector valued functions and how we do calculus with them.
  7. [7]
    [PDF] Three Dimensions
    We now want to talk about three-dimensional space; to identify every point in three dimensions we require three numerical values. The obvious way to make ...
  8. [8]
    Physicists say universe evolution favored three and seven dimensions
    Sep 28, 2005 · They found that as the branes diluted, the ones that survived displayed three dimensions or seven dimensions. In our universe, everything we see ...
  9. [9]
    Does our 3-D world hold six other dimensions? - Cornell Chronicle
    Mar 28, 2006 · Brane-world theory, which proposes that our three-dimensional universe lies inside higher spatial dimensions, and we are no more aware of them.
  10. [10]
    Topoi on Topos: The Development of Aristotle's Concept of Place
    Aristotle's notion of place receives its fullest development in Physics A 1-5, where he argues that place is the inner limit of a containing body. The.
  11. [11]
    [PDF] 1 15 space and place
    Aristotle takes this definition of place to be a refinement of a common-sense notion of location, illustrated by the paradigmatic example of water contained in ...
  12. [12]
    [PDF] Euclid's Elements of Geometry - Richard Fitzpatrick
    The Elements consists of thirteen books. Book 1 outlines the fundamental propositions of plane geometry, includ- ing the three cases in which triangles are ...
  13. [13]
    Epistemology of Geometry - Stanford Encyclopedia of Philosophy
    Oct 14, 2013 · Physical space was the naïve, three-dimensional version of the space of Euclid's Elements and of Cartesian coordinatised three-dimensional ...Missing: scholarly | Show results with:scholarly
  14. [14]
    [PDF] Aryabhata and Axial Rotation of Earth - Indian Academy of Sciences
    The J:tg-Veda contains verses suggesting that the Earth was considered spherical; the Satapatha BrahmaJta de-.
  15. [15]
    [PDF] Al-Biruni and the Mathematical Geography - HAL
    Aug 7, 2019 · Using it, he measured the Earth's circumference finding a value quite close to the modern one. As we will see in this paper, Al-Biruni also ...
  16. [16]
    Thomas Aquinas - Stanford Encyclopedia of Philosophy
    Dec 7, 2022 · Thomas Aquinas (ca. 1225–1274). The greatest figure of thirteenth-century Europe in the two preeminent sciences of the era, philosophy and theology.Life and Works · Cognitive Theory · Will and Freedom · Ethics
  17. [17]
    ST. THOMAS AQUINAS'S COMMENTARY ON ARISTOTLE'S ... - jstor
    philosophy of nature and of man, his own cosmology, metaphysics, and ethics. In this essay I shall not deal extensively with the sources of. Aquinas's Expositio ...
  18. [18]
    Linear Perspective: Brunelleschi's Experiment - Smarthistory
    An introduction to Filippo Brunelleschi's experiment regarding linear perspective, c. 1420, in front of the Baptistry in Florence.Missing: source | Show results with:source
  19. [19]
    René Descartes (1596 - 1650) - Biography - MacTutor
    René Descartes was a French philosopher whose work, La géométrie, includes his application of algebra to geometry from which we now have Cartesian geometry. His ...
  20. [20]
    Chronology for 1740 - 1760 - MacTutor History of Mathematics
    Euler states his theorem V − E + F = 2 V - E + F = 2 V−E+F=2 for polyhedra. 1753. Simson notes that in the Fibonacci sequence the ratio between adjacent numbers ...
  21. [21]
    Leonhard Euler (1707 - 1783) - Biography - MacTutor
    By 1739 Euler had found the rational coefficients C C C in ζ ( 2 n ) = C π 2 n \zeta(2n) = C\pi^{2n} ζ(2n)=Cπ2n in terms of the Bernoulli numbers. Other work ...
  22. [22]
    August Möbius (1790 - 1868) - Biography - MacTutor
    August Ferdinand Möbius ... He introduced a configuration now called a Möbius net, which was to play an important role in the development of projective geometry.
  23. [23]
    Riemann's Bases of Geometry - MacTutor - University of St Andrews
    William Kingdon Clifford translated Bernhard Riemann's article on geometry ... curvature of every measurable portion of space does not differ sensibly from zero.
  24. [24]
    Cartesian coordinate system - Science, civilization and society
    In his work La Geéometrie ("Geometry") René Descartes introduced the rectangular system of coordinates now known as the Cartesian coordinate system.Missing: history | Show results with:history
  25. [25]
    Vectors in two- and three-dimensional Cartesian coordinates
    In three-dimensional space, there is a standard Cartesian coordinate system (x,y,z). Starting with a point which we call the origin, construct three mutually ...
  26. [26]
    11.1 Introduction to Cartesian Coordinates in Space
    Each point P in space can be represented with an ordered triple, P=(a,b,c), where a, b and c represent the relative position of P along the x-, y- and z-axes, ...Missing: history | Show results with:history
  27. [27]
    Cylindrical and spherical coordinates
    Cylindrical Coordinates: When there's symmetry about an axis, it's convenient to take the z-axis as the axis of symmetry and use polar coordinates in the ...
  28. [28]
    The Jacobian for Polar and Spherical Coordinates
    The Jacobian for Polar and Spherical Coordinates. We first compute the Jacobian for the change of variables from Cartesian coordinates to polar coordinates.
  29. [29]
    Spherical coordinates - Math Insight
    In summary, the formulas for Cartesian coordinates in terms of spherical coordinates are x=ρsinϕcosθy=ρsinϕsinθz=ρcosϕ.Missing: source | Show results with:source
  30. [30]
    [PDF] Three-Dimensional Rotation Matrices
    The most general three-dimensional rotation matrix represents a counterclockwise rotation by an angle θ about a fixed axis that lies along the unit vector n.
  31. [31]
    Calculus III - Equations of Lines - Pauls Online Math Notes
    Aug 15, 2023 · In 3D, a line's equation needs a point and a vector parallel to it. The vector form is →r=→r0+t→v, where t moves along the line.
  32. [32]
    1.5 Equations of Lines in 3d
    A 3D line is defined by a point and a direction vector. Parametric equations are derived, and solving for the parameter gives symmetric equations.
  33. [33]
    Normal Vector -- from Wolfram MathWorld
    The normal vector, often simply called the "normal," to a surface is a vector which is perpendicular to the surface at a given point.
  34. [34]
    Point-Plane Distance -- from Wolfram MathWorld
    The (signed) distance from a point x_0 to the plane containing the three points is given by D_i=n^^·(x_0-x_i), where x_i is any of the three points.
  35. [35]
    Distance from point to plane - Math Insight
    The distance from point P to plane Ax+By+Cz+D=0 is calculated as d=|Ax1+By1+Cz1+D|√A2+B2+C2. The shortest distance is along a perpendicular line.
  36. [36]
    Lesson Explainer: Angle between Two Straight Lines in Space
    The angle 𝜃 between two lines in space is the angle between their direction vectors ⃑ d and ⃑ d . ... If the lines are perpendicular, then ⃑ d ⋅ ⃑ d = 0 ...<|control11|><|separator|>
  37. [37]
    3D Coordinate Geometry - Skew Lines | Brilliant Math & Science Wiki
    Distance between Skew Lines​​ The (shortest) distance between a pair of skew lines can be found by obtaining the length of the line segment that meets ...
  38. [38]
    11-01 3-D Coordinate System - Andrews University
    In three dimensions, all the points a given distance from the central point is a sphere, and its equation is (x − h)2 + (y − k)2 + (z − k)2 = r2 where (h, k, j) ...Missing: Cartesian | Show results with:Cartesian
  39. [39]
    [PDF] AREA AND VOLUME WHERE DO THE FORMULAS COME FROM?
    By multiplying both sides of the equation by 3 and dividing both sides by r the formula for the surface area of the sphere,. A = 4πr2 is derived. SURFACE AREA ...
  40. [40]
    [PDF] Geometry in Very High Dimensions
    The n-dimensional ball, or n-ball, of radius r is the subset of Euclidean n-space defined by: Bn(r) = {x ∈ Rn : kxk ≤ r}. Here, x = (x1,...,xn) is an element of ...Missing: three- | Show results with:three-
  41. [41]
    Chapter 3: Section 7: Part 4
    Thus, a geodesic on a sphere is a great circle, which is a circle formed by the intersection of the sphere with a plane through the sphere's center.
  42. [42]
    [PDF] There are 5 convex regular 3- polytopes. Euler's polyhe
    Dec 6, 2009 · There are 5 platonic solids, two-dimensional convex polyhedra, for which all faces and all vertices are the same and every face is a regular ...
  43. [43]
    [PDF] Polyhedra and Euler Characteristics - MIT Mathematics
    The value of the Euler Characteristic χ(S) for a 2-sphere S equals to 2, as defined by the number of 0-cells (1) subtracted by the number of 1-cells (1) and ...
  44. [44]
    [PDF] A Reversible Material for Folding Three-Dimensional Lattice Structures
    Mar 9, 2017 · To date, cubes are the only regular polyhedron known to be able tessellate in 3D Euclidean space. Since our goal is to use the reconfigurable ...
  45. [45]
    [PDF] The Volume of a Platonic Solid - UNL Digital Commons
    Jul 7, 2007 · Inscribing a Regular Tetrahedron in a Cube to Find Its Volume. Inscribing a regular tetrahedron in a cube may be done by letting each edge of ...
  46. [46]
    [PDF] 11.5: Quadric surfaces
    The most general second-degree equation in three variables x, y and z: Ax2 + By2 + Cz2 + axy + bxz + cyz + d1x + d2y + d3z + E = 0,. (1) where A, B, C, a, b ...
  47. [47]
    Calculus III - Quadric Surfaces - Pauls Online Math Notes
    Nov 16, 2022 · Ellipsoid. Here is the general equation of an ellipsoid. Here is a sketch of a typical ellipsoid. If a=b=c a = b = c then we will have a sphere.
  48. [48]
    The hyperbolic paraboloid - Math Insight
    Because it's such a neat surface, with a fairly simple equation, we use it over and over in examples. Hyperbolic paraboloids are often referred to as “saddles,” ...
  49. [49]
    15 Surfaces of Revolution. The Torus - The Geometry Center
    The area of the surface of revolution on a curve C is equal to the product of the length of C and the length of the path traced by the centroid of C (which is 2 ...
  50. [50]
    The Mathematics of Torus Triptych - Brown Math
    Oct 8, 2000 · T(q,f) = ((a + bcos q) cos f, (a + bcos q) sin f,bsin q). In "Torus Triptych" we used a = sqrt(2) and b = 1. This basic torus was rotated to ...
  51. [51]
    [PDF] Geometry of Masses I - UCLA Math Circle
    Oct 11, 2024 · Pappus's First Centroid Theorem allows you to determine the area of a surface of revolution using information about the line segment and the ...
  52. [52]
    Vectors - Calculus II - Pauls Online Math Notes
    Nov 16, 2022 · In this section we will introduce some common notation for vectors as well as some of the basic concepts about vectors such as the magnitude of a vector and ...
  53. [53]
    [PDF] Chapter 3 Vectors
    Cartesian coordinates consist of a set of mutually perpendicular axes, which intersect at a common point, the origin O. We live in a three-dimensional spatial ...
  54. [54]
    Dot Product - World Web Math: Vector Calculus - MIT
    The algebraic definition of the dot product is that. (v1, v2, v3) · (w1, w2, w3) = v1 w1 + v2 w2 + v3 w3. The miracle is that these two definitions, the ...
  55. [55]
    Dot Product - Calculus II - Pauls Online Math Notes
    Nov 16, 2022 · The dot product is called the scalar product. The dot product is also an example of an inner product and so on occasion you may hear it called an inner product.
  56. [56]
    Dot product examples - Math Insight
    Examples of calculating the dot product of two- and three-dimensional vectors.
  57. [57]
    ALAFF The vector 2-norm (Euclidean length) - UT Computer Science
    The length of a vector is most commonly measured by the square root of the sum of the squares of the elements, also known as the Euclidean norm.
  58. [58]
    Vector Mathematics in 3D - Computer Science
    The normalization of a vector \bs{p} is \frac{\bs{p}}{||\bs{p}||} (i.e., the vector divided by its norm). An Observation: ||\bs{p}|| is a scalar so ...
  59. [59]
    Projections and components
    Since v⋅v=‖v‖‖v‖cos(0)=‖v‖2, we often write this as dashed red vector = (u⋅vv⋅v)v. This red vector is called the (vector) projection of u ...
  60. [60]
    1.3 The Dot Product
    The dot product can help us understand the angle between two vectors. For instance, if we are given two non-zero vectors ...
  61. [61]
    11.3 The Dot Product‣ Chapter 11 Vectors ‣ Part Calculus III
    Application to Work. In physics, the application of a force F to move an object in a straight line a distance d produces work; the amount of work W is W=Fd ...
  62. [62]
    The Cross Product - Department of Mathematics at UTSA
    Jan 20, 2022 · The cross product, a × b (read "a cross b"), is a vector that is perpendicular to both a and b, and thus normal to the plane containing them.
  63. [63]
    10.6 Torque – General Physics Using Calculus I - UCF Pressbooks
    We calculate each torque individually, using the cross product, and determine the sign of the torque. Then we sum the torques to find the net torque.
  64. [64]
    Parallelepiped -- from Wolfram MathWorld
    A parallelepiped is a prism whose faces are all parallelograms. Let A, B, and C be the basis vectors defining a three-dimensional parallelepiped.
  65. [65]
    [PDF] 6.801/6.866: Machine Vision, Lecture 18 - MIT OpenCourseWare
    ... cross products: Although a cross product produces a vector in 3D, in higher dimensions the result of a cross product is a subspace, rather than a vector.
  66. [66]
    Vector Space -- from Wolfram MathWorld
    A consequence of the axiom of choice is that every vector space has a vector basis. A module is abstractly similar to a vector space, but it uses a ring to ...
  67. [67]
    Finite-Dimensional Vector Spaces - SpringerLink
    Download chapter PDF · Spaces. Paul R. Halmos. Pages 1-54. Transformations. Paul R ... Finite-Dimensional Vector Spaces. Authors: Paul R. Halmos. Series Title ...
  68. [68]
    affine space in nLab
    Jan 13, 2025 · An affine space or affine linear space is a vector space that has forgotten its origin. An affine linear map (a morphism of affine spaces) is a linear map.
  69. [69]
    Barycentric Coordinates -- from Wolfram MathWorld
    Barycentric coordinates are triples of numbers corresponding to masses placed at the vertices of a reference triangle . These masses then determine a point , ...
  70. [70]
    Inner Product -- from Wolfram MathWorld
    A vector space together with an inner product on it is called an inner product space. This definition also applies to an abstract vector space over any field.
  71. [71]
    Del -- from Wolfram MathWorld
    The upside-down capital delta symbol del , also called "nabla" used to denote the gradient and other vector derivatives.<|separator|>
  72. [72]
    Gradient -- from Wolfram MathWorld
    The term "gradient" has several meanings in mathematics. The simplest is as a synonym for slope. The more general gradient, called simply "the" gradient in ...
  73. [73]
    Divergence -- from Wolfram MathWorld
    The physical significance of the divergence of a vector field is the rate at which density exits a given region of space.
  74. [74]
    Curl -- from Wolfram MathWorld
    The curl of a vector field, denoted curl(F) or del xF (the notation used in this work), is defined as the vector field having magnitude equal to the maximum ...
  75. [75]
    Laplacian -- from Wolfram MathWorld
    The Laplacian is extremely important in mechanics, electromagnetics, wave theory, and quantum mechanics, and appears in Laplace's equation.
  76. [76]
    Cylindrical Coordinates -- from Wolfram MathWorld
    The gradient operator in cylindrical coordinates is given by. del =r^^partial/(partialr)+theta^^1/. (32). so the gradient components become. del _rr^^, = 0. (33).
  77. [77]
    Spherical Coordinates -- from Wolfram MathWorld
    A system of curvilinear coordinates that are natural for describing positions on a sphere or spheroid.
  78. [78]
    Calculus III - Line Integrals - Pauls Online Math Notes
    Nov 16, 2022 · In this section we will give the fundamental theorem of calculus for line integrals of vector fields. This will illustrate that certain kinds of line integrals ...
  79. [79]
    Calculus III - Line Integrals - Part I - Pauls Online Math Notes
    Nov 16, 2022 · With line integrals we will start with integrating the function f(x,y) f ( x , y ) , a function of two variables, and the values of x x and y y ...
  80. [80]
    Calculus III - Line Integrals of Vector Fields - Pauls Online Math Notes
    Nov 16, 2022 · In this section we will define the third type of line integrals we'll be looking at : line integrals of vector fields.
  81. [81]
    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.Missing: flux | Show results with:flux
  82. [82]
    Calculus III - Surface Integrals of Vector Fields
    Nov 16, 2022 · In this section we will introduce the concept of an oriented surface and look at the second kind of surface integral we'll be looking at ...
  83. [83]
    Calculus III - Change of Variables - Pauls Online Math Notes
    Nov 16, 2022 · In order to change variables in a double integral we will need the Jacobian of the transformation. Here is the definition of the Jacobian.
  84. [84]
    [PDF] A History of the Divergence, Green's, and Stokes' Theorems
    Cauchy wrote his first paper in 1811 and by 1816 had solved a claim by Pierre ... mathematicians, he would still be the first to provide a valid proof of Stokes' ...
  85. [85]
    [PDF] The History of Stokes' Theorem - Harvard Mathematics Department
    The first published proof of the theorem seems to have been in a monograph of Hermann Hankel in 1861 [10]. Hankel gives no credit for the theorem, only a ...
  86. [86]
    [PDF] Divergence-measure fields: Gauss-Green formulas and Normal Traces
    Carl Gauss (30. April 1777 – 23 February 1855). The formula that would be later known as the divergence theorem was first discovered by. Lagrange2 in 1762 (see ...
  87. [87]
    [PDF] 4 The Integral Theorems - DAMTP
    This means that the magnetic vector field can't pile up anywhere: at any given point in space, there is as much magnetic field coming in as there is going out.
  88. [88]
    3-manifold - Planet Math
    Mar 22, 2013 · A 3-manifold is a Hausdorff topological space which is locally homeomorphic to the Euclidean space R3 ℝ 3 . One can see ...
  89. [89]
    [PDF] Notes on Basic 3-Manifold Topology
    By Proposition 1.6, Mϕ is irreducible since its universal cover is R3 . Once again we restrict attention to orientable 3 manifolds for simplicity. For Mϕ this ...
  90. [90]
    [PDF] Poincaré duality in dimension 3 - arXiv
    Sep 17, 2004 · Abstract The paper gives a review of progress towards extending the Thurston programme to the Poincaré duality case. In the first section, ...
  91. [91]
    [PDF] lecture 10: the whitney embedding theorem
    Theorem 1.2 (The Whitney embedding theorem: median version). Any compact manifold M of dimension m can be embedded into R2m+1 and immersed into R2m. Proof.
  92. [92]
    Sphere eversion from the viewpoint of generic homotopy
    Jun 1, 2017 · In this paper, we construct a sphere eversion by lifting a “simple” generic homotopy of S 2 to R 2 to a generic regular homotopy of S 2 to R 3 .
  93. [93]
    [PDF] The Poincaré Conjecture - Clay Mathematics Institute
    As we know the Poincaré conjecture is about characterizing the 3- dimensional sphere in topological terms and its resolution by Perelman, combined with the ...
  94. [94]
    Hyperbolic Geometry -- from Wolfram MathWorld
    In hyperbolic geometry, the sum of angles of a triangle is less than 180 degrees, and triangles with the same angles have the same areas.
  95. [95]
    [PDF] Coding billiards in hyperbolic 3-space - arXiv
    Jul 3, 2023 · The hyperbolic. 3-space, denoted as H3, is a 3-dimensional Riemannian manifold that carries a constant negative curvature. The notion of ...
  96. [96]
    Elliptic Geometry -- from Wolfram MathWorld
    Elliptic geometry is a non-Euclidean geometry with positive curvature which replaces the parallel postulate with the statement "through any point in the plane,
  97. [97]
    Elliptic Space -- from Wolfram MathWorld
    A space endowed with a non-Euclidean elliptic geometry.
  98. [98]
    [PDF] Lecture XXII: The homogeneous, isotropic Universe
    ... Robertson-Walker (FLRW) metric, and the. Einstein equations ... We can assess these spatial metrics by considering the curvature of a 3D slice at a = 1.
  99. [99]
    Understanding the curvature parameter k - Physics Forums
    Oct 20, 2016 · The curvature parameter k in the Robertson-Walker metric is a constant that determines the spatial geometry of the universe, with values of ...Principal and Gaussian curvature of the FRW metric - Physics ForumsFriedman-Robertson-Walker metric - Physics ForumsMore results from www.physicsforums.com
  100. [100]
    [PDF] The Quaternions and the Spaces S3, SU(2), SO(3), and RP
    The quaternions of norm 1, also called unit quaternions, are in bijection with points of the real 3-sphere S3. It is easy to verify that the unit ...
  101. [101]
    Hypersphere -- from Wolfram MathWorld
    The n-hypersphere (often simply called the n-sphere) is a generalization of the circle (called by geometers the 2-sphere) and usual sphere (called by geometers ...
  102. [102]
    [PDF] Introduction to Projective Geometry
    In 1906, Veblen and Bussey gave this finite Projective Geometry the name PG(n,p) and extended it to PG(n,q), where q = pk, p is prime, and k is any positive ...
  103. [103]
    [PDF] FINITE CUBES AND FINITE AFFINE SPACES
    Suppose that F is a finite field. Then the affine geometry AG(3,F) of dimension 3 over F is a finite analogue of ordinary 3-dimensional Euclidean space.
  104. [104]
  105. [105]
    Projective Plane -- from Wolfram MathWorld
    A projective plane, sometimes called a twisted sphere (Henle 1994, p. 110), is a surface without boundary derived from a usual plane by addition of a line ...
  106. [106]
    Parallel Postulate -- from Wolfram MathWorld
    Given any straight line and a point not on it, there exists one and only one straight line which passes through that point and never intersects the first line.
  107. [107]
    Non-Euclidean Geometry -- from Wolfram MathWorld
    The non-Euclidean geometries are called hyperbolic geometry (or Lobachevsky-Bolyai-Gauss geometry) and elliptic geometry (or Riemannian geometry).
  108. [108]
    [PDF] Some Exact Solutions in General Relativity - arXiv
    The fundamental physical postulate of GR is that the presence of matter causes curvature in the spacetime in which it exists. This curvature is taken to be the ...<|separator|>
  109. [109]
    [PDF] Efficient, robust, and provably good approximation of 3D medial axes
    A voxel shape is the interior of the union of a finite set of voxels. By this definition, a voxel shape is an open set that does not include the vertices, edges ...